<<

CHAPTER 19 Supranuclear and Internuclear Ocular Motility Disorders

David S. Zee and David Newman-Toker

OCULAR MOTOR SYNDROMES CAUSED BY LESIONS IN OCULAR MOTOR SYNDROMES CAUSED BY LESIONS OF THE MEDULLA THE SUPERIOR COLLICULUS Wallenberg’s Syndrome (Lateral Medullary Infarction) OCULAR MOTOR SYNDROMES CAUSED BY LESIONS OF Syndrome of the Anterior Inferior Cerebellar Artery THE Skew Deviation and the Ocular Tilt Reaction OCULAR MOTOR ABNORMALITIES AND DISEASES OF THE OCULAR MOTOR SYNDROMES CAUSED BY LESIONS IN THE Parkinson’s Disease Location of Lesions and Their Manifestations Huntington’s Disease Etiologies Other Diseases of Basal Ganglia OCULAR MOTOR SYNDROMES CAUSED BY LESIONS IN OCULAR MOTOR SYNDROMES CAUSED BY LESIONS IN THE THE CEREBRAL HEMISPHERES Lesions of the Internuclear System: Internuclear Acute Lesions Ophthalmoplegia Persistent Deficits Caused by Large Unilateral Lesions Lesions of the Focal Lesions Lesions of the Paramedian Pontine Reticular Formation Ocular Motor Apraxia Combined Unilateral and Internuclear Abnormal Eye Movements and Dementia Ophthalmoplegia (One-and-a-Half Syndrome) Ocular Motor Manifestations of Seizures Slow from Pontine Lesions Eye Movements in Stupor and Coma Saccadic Oscillations from Pontine Lesions OCULAR MOTOR DYSFUNCTION AND MULTIPLE OCULAR MOTOR SYNDROMES CAUSED BY LESIONS IN SCLEROSIS THE MESENCEPHALON OCULAR MOTOR MANIFESTATIONS OF SOME METABOLIC Sites and Manifestations of Lesions DISORDERS Neurologic Disorders that Primarily Affect the Mesencephalon EFFECTS OF DRUGS ON EYE MOVEMENTS

In this chapter, we survey clinicopathologic correlations proach, although we also discuss certain metabolic, infec- for supranuclear ocular motor disorders. The presentation tious, degenerative, and inflammatory diseases in which su- follows the schema of the 1999 text by Leigh and Zee (1), pranuclear and internuclear disorders of eye movements are and the material in this chapter is intended to complement prominent. Details about these conditions can be found in that of Chapters 17 and 23. We emphasize an anatomic ap- the appropriate chapters in other volumes of this text.

OCULAR MOTOR SYNDROMES CAUSED BY LESIONS IN THE MEDULLA Many structures within the medulla are important in the intercalatus nucleus, and ventrally the nucleus of Roller. control of eye movements: the vestibular nuclei, perihypog- These nuclei are interconnected with other ocular motor lossal nuclei, and inferior olive and its outflow pathway structures in the stem and cerebellum. The NPH and through the inferior cerebellar peduncle. The perihypoglos- the adjacent medial vestibular nuclei (MVN) are critical for sal nuclei consist of the nucleus prepositus hypoglossi holding horizontal positions of gaze (the neural integrator) (NPH), which lies in the floor of the fourth ventricle, the (2). These structures also participate in vertical gaze-hold- 907 908 CLINICAL NEURO-OPHTHALMOLOGY ing, although more rostral structures, especially the intersti- neurons and their dendrites contain increased acetylcholines- tial nucleus of Cajal (INC), also contribute. With lesions in terase reaction product (14). Guillain and Mollaret (15) sug- the paramedian structures of the medulla, —com- gested that disruption of connections between the dentate monly upbeat—is the most common finding (3). Upbeat and the contralateral olivary nucleus (which run via the red nystagmus may also reflect involvement of a ventral tegmen- nucleus and central tegmental tract) causes this syndrome. tal pathway for the upward vestibulo-ocular reflex (VOR) Another hypothesis is that the ocular oscillations are caused producing a downward vestibular bias and a consequent up- by instability in circuits that include the projection from the beat nystagmus (4). Wernicke’s disease commonly affects inferior olive to the flocculus, which is thought to be impor- the region of NPH and MVN, which may account for the tant in the adaptive control of the VOR (10). horizontal gaze-evoked nystagmus and spontaneous vertical Occasionally, lesions are restricted to the vestibular nu- nystagmus and loss of vestibular responses that occur with clei. For example, vertigo may be the sole symptom of an this disease. exacerbation of (MS) (16,17) and of brain Lesions of the inferior olivary nucleus or its connections stem ischemia (18–21). Nystagmus caused by lesions in the may produce the oculopalatal myoclonus (oculopalatal vestibular nuclei may be purely horizontal, vertical, or tor- tremor) syndrome (5,6). This condition usually develops sional, or mixed. Moreover, nystagmus from a central vestib- weeks to months after a brain stem or cerebellar infarction, ular lesion can mimic that caused by peripheral vestibular although it may also occur with degenerative conditions disease (22,23). Dolichoectasia of the basilar artery may pro- (7,8). The term myoclonus is misleading because the move- duce a variety of combinations of central and peripheral ves- ments of affected muscles are to and fro and are approxi- tibular syndromes (24,25). Microvascular compression of mately synchronized, typically at a rate of 2–4 cycles/sec. the vestibulocochlear nerve may produce paroxysmal ver- Ocular palatal tremor is the better term (6). The abnormal tigo (26,27). Brandt and Dieterich (28), Bu¨ttner et al. (29), ocular movements consist of pendular oscillations that are and Dieterich (30) provide useful topographic schemes for often vertical but may have a horizontal or torsional compo- localizing central vestibular syndromes and central vestibu- nent. Predominantly vertical oscillations are usually associ- lar nystagmus within the brain stem and cerebellum. ated with symmetric bilateral palatal tremor. Mixed vertical and torsional movements, sometimes disconjugate and with WALLENBERG’S SYNDROME (LATERAL a seesaw quality, are associated with unilateral or asymmet- MEDULLARY INFARCTION) ric palatal tremor (9,10). Occasionally, patients develop the eye oscillations without movements of the palate. Closing Lesions of the vestibular nuclei commonly affect neigh- the eyes may bring out the vertical ocular oscillations (11). boring structures, in particular the cerebellar peduncles and The nystagmus sometimes disappears with sleep, but the perihypoglossal nuclei. The best-recognized syndrome in- palatal movements usually persist. Occasionally, the oscilla- volving the vestibular nuclei is caused by a lateral medullary tions resolve spontaneously. Gabapentin may partially ame- infarction (Wallenberg’s syndrome) (Fig. 19.2). The typical liorate the eye oscillations (12). findings of Wallenberg’s syndrome are impairment of sensa- The main pathologic finding with palatal tremor is hyper- tion of pain and temperature over the ipsilateral face, ipsi- trophy of the inferior olivary nucleus, which is often seen on lateral Horner’s syndrome and limb , and dysarthria magnetic resonance (MR) imaging (13). The olivary nucleus and dysphagia. Sensation of pain and temperature are im- contains enlarged, vacuolated neurons with expanded den- paired over the contralateral trunk and limbs. The ipsilateral drites and enlarged astrocytes (Fig. 19.1). The hypertrophic facial nerve may also be affected if the infarct extends more

Figure 19.1. Pathology of palato-ocular myoclonus. A sec- tion through the cerebellum and medulla shows marked demy- elination of the right dentate nucleus and restiform body (dou- ble arrows). The left inferior olive is hypertrophic and shows mild demyelination (arrow). (From Nathanson M. Palatal my- oclonus: further clinical and pathophysiological observations. Arch Neurol 1956;75Ϻ285–296.) SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 909

and unusual sensations of body and environmental tilt, often so bizarre as to suggest a psychogenic origin (35,36). Pa- tients may report that the whole room appears tilted on its side or even upside down; with their eyes closed, they may feel themselves to be tilted. Such symptoms are occasionally reported in patients without signs of lateral medullary infarc- tion and may be caused by transient brain stem or cerebellar ischemia (37–40) and occasionally with lesions in the thala- mus (41), cerebral hemispheres (42), or peripheral vestibular apparatus (43). Lateropulsion, a compelling sensation of being pulled to- ward the side of the lesion, is often a prominent symptom in patients with Wallenberg’s syndrome and is also reflected Figure 19.2. Wallenberg’s syndrome. Transverse section through the me- in the ocular motor system (44–46). If the patient is asked dulla oblongata showing a unilateral infarction in its dorsolateral region. to fix straight ahead and then gently close the lids, the eyes (From Ongerboer de Visser BW, Kuypers HGJM. Late blink reflex changes deviate conjugately toward the side of the lesion. This is in lateral medullary lesions: an electrophysiological and neuroanatomical reflected in the corrective saccades that the patient must study of Wallenberg’s syndrome. Brain 1978;101Ϻ285–294.) make on eye opening to reacquire the target. Lateropulsion may appear with a blink. Saccadic eye movements are also affected by lateropul- rostrally (Fig. 19.3). Wallenberg’s syndrome is most com- sion (Fig. 19.5) (44,46–49). Horizontal saccades directed monly caused by occlusion of the ipsilateral vertebral artery; toward the side of the lesion usually overshoot the target, occasionally, the posterior inferior cerebellar artery is selec- and saccades directed away from the side of the lesion under- tively involved (31,32) (Fig. 19.4). Dissection of the verte- shoot the target; this is referred to as ipsipulsion of saccades bral artery (either spontaneous or traumatic, such as follow- and should be differentiated from the contrapulsion of sac- ing chiropractic manipulation) may cause Wallenberg’s cades that occurs with lesions in the superior cerebellar pe- syndrome (33). Rarely, demyelinating disease is the cause duncle due, for example, to superior cerebellar artery occlu- (34). sions (discussed later). Quick phases of nystagmus are The symptoms of Wallenberg’s syndrome include vertigo similarly affected in Wallenberg’s syndrome, so that sac-

Figure 19.3. showing the specific neural structures (shaded area) that are commonly damaged in Wallen- berg’s syndrome. 910 CLINICAL NEURO-OPHTHALMOLOGY

phases) can also occur (54). The ocular tilt reaction (dis- cussed later) commonly occurs in Wallenberg’s syndrome (55), including a skew deviation with an ipsilateral hypo- tropia. The eyes counter-roll with the top poles rotated to- ward the side of the lesion, but unequally, so that there is also a cyclodeviation. The lower eye is usually more ex- torted. Some patients show ipsilateral head tilt (56). The skew deviation and head tilt arise from imbalance in path- ways mediating otolith responses. The subjective sensations of tilt or inversion of the world probably also reflect involve- ment of central projections from the graviceptors—the utri- cle and saccule. Smooth pursuit is usually impaired in Wallenberg’s syn- drome, particularly for tracking targets moving away from the side of the lesion (46). Caloric testing usually shows intact horizontal canal function. During both rotational and caloric testing, there is a directional preponderance of slow phases, usually toward the side of the lesion (45,46,57). Head nystagmus also occurs in some patients (44). Many of the findings in Wallenberg’s syndrome, includ- ing the bizarre visual disturbances and the skew deviation, may reflect imbalance of otolith influences caused by direct damage to the caudal aspects of the vestibular nuclei. Dam- age to the restiform body, which carries olivocerebellar pro- jections, may also account for some of the ocular motor findings, especially the steady-state deviation of the eyes toward the side of the lesion and the ipsipulsion of saccades (47–49,58,59). Ipsipulsion of saccades, with deviation of the eyes to the side of the lesion, can be reproduced experimen- Figure 19.4. Neuroimaging of Wallenberg’s syndrome. T2-weighted magnetic resonance image, axial view, in a 51-year-old man with Wallen- tally by lesions of the fastigial nucleus (60). This finding berg’s syndrome characterized in part by lateropulsion of saccades toward supports the hypothesis that in Wallenberg’s syndrome, the the right side, a skew deviation with the right eye being hypotropic, and interruption of climbing fiber input to the dorsal cerebellar the ocular tilt reaction, shows a hyperintense area (arrowhead) consistent vermis releases Purkinje cell inhibition upon the underlying with an infarct on the right side of the medulla. fastigial nucleus, thus leading to the equivalent of a lesion in the fastigial nucleus (58). An analogous increase in Pur- kinje cell inhibition from the flocculus to the vestibular nu- cades directed away from the side of the lesion are smaller cleus may also play a role in the nystagmus (slow phase than those toward the lesion. On attempting a purely vertical toward the side of the lesion) seen in these patients. refixation, an oblique directed toward the side of the lesion is produced. Corrective saccades then bring the SYNDROME OF THE ANTERIOR INFERIOR eyes back to the target (50). Saccades made in total darkness CEREBELLAR ARTERY also show lateropulsion, although in one report the patient The anterior inferior cerebellar artery (AICA) supplies was still able to make corrective saccades to the remembered portions of the vestibular nuclei and the adjacent dorsolateral location of a previously seen target (51). This finding im- brain stem, and the inferior lateral cerebellum. The AICA plied that the central ‘‘knew’’ actual eye is also the origin of the labyrinthine artery in most persons position. With time, vertical saccades may become more and also sends a twig to the cerebellar flocculus in the cerebe- bizarre; S-shaped saccadic trajectories can appear a week or llopontine angle. Consequently, ischemia in the distribution more after the onset of the illness and may reflect an adaptive of the AICA may cause vertigo, vomiting, , fa- strategy to correct the saccadic abnormality. Torsipulsion, cial palsy, and ipsilateral limb ataxia, along with deficits in inappropriate torsional saccades during attempted horizontal gaze-holding and pursuit as well as vestibular nystagmus or vertical saccades, also may occur, sometimes in associa- (61–63) (Fig. 19.6). The ocular motor signs reflect a combi- tion with torsional nystagmus (52,53). nation of involvement of the labyrinth, vestibular nuclei, and Spontaneous nystagmus in Wallenberg’s syndrome, when flocculus (discussed later) . present, is usually horizontal or mixed horizontal and tor- sional with a small vertical component (52). In primary posi- SKEW DEVIATIONANDTHE OCULAR TILT tion, the slow phase is directed toward the side of the lesion, REACTION although it may reverse direction in eccentric positions, sug- gesting that the gaze-holding mechanism is also impaired. Skew deviation is a vertical misalignment of the visual Lid nystagmus (synkinetic lid twitches with horizontal quick axes caused by a disturbance of prenuclear vestibular inputs SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 911

Figure 19.5. Lateropulsion of saccades in a patient with a left Wallenberg’s syndrome. A, On attempted leftward gaze, the patient overshoots the target and must make a corrective saccade. B, On attempted rightward gaze, the patient makes a series of hypometric saccades. C and D, On attempted upward and downward gaze, the eyes move obliquely to the left and must make several refixation movements back to center. (Redrawn from Kommerell G, Hoyt WF. Lateropulsion of saccadic eye movements: electro-oculographic studies in a patient with Wallenberg’s syndrome. Arch Neurol 1973;28Ϻ313–318.) to the oculomotor nuclei. Torsional and horizontal devia- In many patients, skew deviation is associated with ocular tions may be associated findings. The may be torsion and head tilt, the pathologic ocular tilt reaction the same (comitant) in all positions of gaze, or it may vary (OTR), which may be tonic (sustained) (56) or paroxysmal and even alternate (typically with an alternating abducting (70,71). Such patients also show a deviation of the subjective hypertropia—i.e., right hypertropia on right gaze, left hyper- vertical (72,73). The ocular torsion may be dissociated, pro- tropia on left gaze) (64–67). When skew deviation is incomi- ducing a cyclodeviation (74). The OTR is usually attributed tant, it may mimic an isolated superior (68) or inferior (69) to an imbalance in otolith-ocular and otolith-collic reflexes oblique palsy by the Bielschowsky three-step test. In such that are part of a phylogenetically old righting response to cases, however, it appears that the addition of a fourth step, a lateral tilt of the head (75). In patients with more rostral assessment of ocular torsion in both eyes, may be able to lesions, interruption of descending pathways involved with distinguish skew deviation from isolated oblique weakness: controlling head posture may also contribute to the head tilt with skew mimicking superior oblique palsy, the hypertropic of the OTR (75,76). eye is incyclotorted (rather than excyclotorted, as one would Skew deviation occurs with a variety of abnormalities in expect if the hypertropia were the result of superior oblique the vestibular periphery, brain stem, or cerebellum weakness) (68). By analogy, when skew mimics inferior (66,77–83) and with raised from supra- oblique palsy, the hypotropic eye is excyclotorted (rather tentorial tumors or pseudotumor cerebri (84,85). In infants, than incyclotorted, as one would expect if the hypotropia a skew deviation may be the harbinger of a subsequent hori- were the result of inferior oblique weakness) (69). It remains zontal (86). to be seen, however, whether the addition of this fourth step Why does skew deviation occur with lesions at so many will reliably distinguish between superior oblique palsy and sites within the posterior fossa? An imbalance in otolith path- all forms of skew deviation, or whether some cases will ways is the most likely cause (77). An imbalance of posterior simply be indistinguishable at the bedside. semicircular canal inputs may also play a role (56,87), al- 912 CLINICAL NEURO-OPHTHALMOLOGY

Figure 19.6. Neuroimaging of infarct in territory of the left anterior inferior cerebellar artery (AICA). A, T2-weighted axial magnetic resonance (MR) scan shows hyperintense area in the region of the left middle cerebellar peduncle (arrowhead). B, T1-weighted axial MR scan after intravenous injection of paramagnetic contrast material shows diffuse enhancement in the distal distribution of the AICA involving the left cerebellar hemisphere. The 69-year-old patient had, among other manifestations, left-beating gaze-evoked nystagmus.

though if that is the case nystagmus should also be present compensatory response to the perceived tilt of the subjective (88). visual vertical, although it may also reflect direct involve- To understand skew deviation, it is helpful to consider ment of descending projections from the vestibular nuclei physiologic aspects of otolith-ocular reflexes (75). In lateral- or the INC to cervical motoneurons (76). eyed animals, tilting the head laterally around the longitudi- Lesions of the vestibular organ or its nerve can cause both nal (anterior-posterior) axis causes a disjunctive, vertical skew deviation and the OTR by producing an imbalance in (skew) deviation (i.e., one eye goes up, the other down) that utricle inputs (80,93). The ocular tilt reaction may also occur acts to hold the visual axis of each eye close to the horizontal. as a component of Tullio phenomenon, which is character- In human subjects, who are frontal-eyed, a static head tilt ized by sound-induced vestibular symptoms (94–96). It oc- (ear to shoulder) causes sustained, largely conjugate counter- curs in patients with perilymph fistula either at the oval or rolling of the eyes (ocular torsion) that is about 10% of the round window, with other abnormal communications be- head roll (89–91), although the actual amount may be related tween the membranous labyrinth and the perilymph space, to various factors, including the angle of vergence. Thus, or with abnormalities of the ossicular chain and its connec- the static ocular response does not compensate for the head tion with the membranous labyrinth. Dehiscence of the roof tilt and is thought to be vestigial. There may also be a small of the superior semicircular canal is another recently identi- amount of skewing in normal subjects during rotation of the fied cause (96–98). The OTR is consistent with the effects head around its roll (anterior-posterior) axis (92). In contrast, of experimental stimulation of the otoliths (100) and the peripheral or central lesions that disrupt otolith inputs often utricular nerve (101), which cause ipsilateral hypertropia and cause large amounts of skew deviation (e.g., 7Њ) and ocular conjugate ocular counter-rolling (i.e., top pole of the eyes torsion (e.g., 25Њ). Usually, any pathologic head tilt (ear to toward the contralateral ear). shoulder) is contralateral to the hypertropic eye, and the ocu- The utricle projects predominantly to the ipsilateral lateral lar torsion is such that the upper poles of the eyes rotate vestibular nucleus, whereas the saccule projects to the y- toward the lower ear. The contralateral head tilt may be a group of vestibular nuclei (101,102). Thus, lesions of the SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 913 vestibular nuclei (e.g., as part of Wallenberg’s lateral medul- the mesencephalon, in or around the INC, thus may cause lary syndrome) may also cause skew deviation with hypo- skew deviation (78,108,109) and the OTR (56,110). When tropia on the side of the lesion (55). In addition, some pa- the head tilt is sustained (tonic), it is contralateral to the side tients show an ipsilateral head tilt and disconjugate ocular of the lesion; in addition, there is usually a hypertropia that torsion. The latter is an excylotropia, with excyclodeviation is ipsilateral to the lesion and a conjugate cyclotorsion that of the ipsilateral, lower eye, but small or absent incyclodevi- is characterized by intorsion of the ipsilateral eye and extor- ation of the contralateral, higher eye (52,103). sion of the contralateral eye (i.e., ocular counter-rolling away Skew deviation is encountered in patients with cerebellar from the side of the mesencephalic lesion). Defects of verti- lesions (66,67,104–106). Some of these patients show an cal eye movements and oculomotor or func- alternating skew deviation that is characterized by a hyperde- tion are also common in this setting. Combined prenuclear viation of the abducting eye. This abnormality also may be and fascicular or nuclear lesions in the may create analogous to a phylogenetically old, otolith-mediated, right- torsion of one eye and the OTR (111,112). ing reflex present in lateral-eyed animals, which in this case In patients with meso-diencephalic lesions, spontaneous is related to the ocular motor response that compensates for ocular oscillations may also be present that suggest involve- fore-and-aft motion (as opposed to lateral roll) of the head. ment of otolith inputs to the oculomotor system, such as Although the brain stem is likely also involved in some of seesaw nystagmus (113) or, more rarely, slowly alternating these patients, skew deviation can occur in some patients skew deviation (114–116). In the latter case, skew deviation who appear to have pure cerebellar disease. This suggests slowly alternates or varies cyclically in magnitude over the that just as the cerebellum governs the semicircular canal- course of a few minutes. The periodicity of the phenomenon ocular reflex, it also influences the otolith-ocular reflexes. (with each eye typically being hypertropic for about 30 sec- Indeed, downbeat nystagmus, which is sometimes attribut- onds to a minute) is reminiscent of periodic alternating nys- able to disease of the flocculus (discussed later), commonly tagmus, and the two phenomena can coexist (117). coexists with this pattern of skew deviation that alternates In some patients, the skew deviation (with or without a with lateral gaze. head tilt) is not sustained but paroxysmal (70,71). In one Utricular projections from the vestibular nuclei probably patient with a clearly defined lesion close to the right INC, cross the midline and ascend in the medial longitudinal fasci- culus (MLF). Therefore, unilateral internuclear ophthal- episodes of contralateral hypertropia and ipsilateral head tilt moplegia (INO) is often associated with a skew deviation, occurred, suggesting an irritative mechanism (71). This in- presumably because lesions of one MLF cause an imbalance terpretation of the findings in paroxysmal skew deviation is of ascending otolith inputs. The INO is usually on the side supported by the results of electric stimulation near the INC of the hypertropic eye, likely reflecting that utricular projec- in the monkey. This produces an ocular tilt reaction that tions are generally damaged in the MLF after crossing at consists of and extorsion of the ipsilateral eye the level of the mid-pons (i.e., a right MLF lesion damages and elevation and intorsion of the contralateral eye (118). utricular inputs from the left ear, leading to a right hyper- With the head free to move, an ipsilateral head tilt also occurs tropia similar conceptually to that seen with a left lateral (119). In human patients, stimulation in the region of the INC medullary syndrome or right rostral midbrain syndrome) causes an ipsilateral OTR (120). Microvascular compression (107). may also cause a paroxysmal skew deviation with torsional In the midbrain, otolith projections contact the oculomotor nystagmus (121). Skew deviation has also been reported as and trochlear nerve nuclei as well as the INC. Lesions in an epileptic phenomenon (122).

OCULAR MOTOR SYNDROMES CAUSED BY LESIONS IN THE CEREBELLUM Clinicians are appropriately cautious in attributing eye features of each of these syndromes are summarized in Table movement abnormalities specifically to cerebellar dysfunc- 19.1. tion, because the brain stem is so frequently damaged in patients with lesions of the cerebellum. Likewise, brain stem LOCATION OF LESIONS AND THEIR lesions can produce a ‘‘functional’’ cerebellar lesion and MANIFESTATIONS corresponding cerebellar eye signs by virtue of a change in the climbing fiber activity projecting to cerebellar Purkinje Experimental lesions of the dorsal vermis (lobules VI and cells, as occurs in Wallenberg’s syndrome (discussed previ- VII) and of the underlying fastigial nuclei (called the fastigial ously). Holmes (123) and Cogan (77), however, recognized oculomotor region) cause saccadic , typically hy- specific cerebellar eye signs, and Daroff (124) listed more pometria if the vermis alone is involved and hypermetria if than 30 ‘‘cerebellar eye signs’’ in a compendium published the deep nuclei are affected (127). Pursuit, especially the in 1982. Recent clinical and experimental studies provide initial acceleration of the eyes during tracking, is also af- additional evidence that cerebellar lesions alone can cause fected by dorsal vermis lesions (128). Lesions of the deep specific ocular motor abnormalities (104,125,126). In es- nuclei can lead to macrosaccadic oscillations, an extreme sence, three principal syndromes can be identified: the syn- degree of hypermetria. Lesions restricted to the dorsal ver- drome of the dorsal vermis and underlying posterior fastigial mis produce not only saccadic dysmetria but also impaired nuclei, the syndrome of the flocculus and paraflocculus, and initiation of pursuit with a decrease in the acceleration of the syndrome of the nodulus and ventral uvula. The main the eyes during tracking, and disturbances of eye alignment 914 CLINICAL NEURO-OPHTHALMOLOGY

Table 19.1 Localization of Cerebellar Eye Movement Abnormalities

Structure Function Disorder

Flocculus and paraflocculus Retinal-image stabilization (smooth tracking Impaired smooth pursuit, VOR cancellation and fixation with head still or free suppression of suppression of caloric nystagmus; gaze-evoked, inappropriate vestibular nystagmus, holding rebound, centripetal and downbeat nystagmus; positions of gaze, adaptive control of the VOR postsaccadic drift; inappropriate amplitude or direction and pulse-step match) of the VOR Nodulus and ventral uvula Control of low-frequency response of the VOR Periodic alternating nystagmus, impaired tilt suppression of postrotatory nystagmus, positional nystagmus, impaired habituation of the VOR, increased duration of vestibular responses. Dorsal vermis and posterior fastigial Saccade accuracy, smooth-pursuit eye alignment Saccadic dysmetria, impaired pursuit, esodeviations nucleus

VOR, vestibulo-ocular reflex. (Zee DS, Walker MF. Cerebellar control of eye movements. In: Chalupa LM, Werner JS, eds. The Visual Neurosciences. Cambridge, MA: MIT Press, 2003:1485Ð1498.) including the development of an esodeviation. The pattern phases of nystagmus during sustained rotations, a failure of saccadic dysmetria that occurs in cerebellar disease, as of tilt suppression of postrotatory nystagmus (145), loss of well as whether corrective saccades occur, may also vary habituation (146) and positional nystagmus (147). with the type of visual stimulus. Saccades to remembered Syndromes of the dorsal vermis and fastigial nucleus, targets are more dysmetric (129–131). In some cerebellar flocculus and paraflocculus, and nodulus are frequently en- patients, only corrective saccades are dysmetric (132). Sac- countered in patients with degenerative disorders that pre- cadic dysmetria may be present for externally triggered sumably affect only the cerebellum. In some instances, movements to a visual target but not for internally triggered pathologic examination confirms disease restricted to the saccades during scanning of a visual scene (133). Deficits cerebellum (148). Similar findings occur in patients with of pursuit may also be produced by lesions of the dorsal focal structural lesions of the cerebellum. vermis (130), as can defects in motion perception (134). FitzGibbon et al. (149) attributed another ocular motor Bilateral symmetric lesions of the deep nuclei do not lead to sign to a focal cerebellar lesion. Patients with cavernous pursuit deficits during sustained tracking, although unilateral angiomas in the middle cerebellar peduncle show torsional lesions lead to a contralateral deficit (135,136), probably nystagmus during vertical pursuit. The direction of the tor- because of an imbalance in eye acceleration signals. sional nystagmus changes with the direction of the pursuit, Experimental lesions of the flocculus and paraflocculus with the eye velocity of the slow phase of the torsional nys- cause gaze-evoked nystagmus, rebound nystagmus, and tagmus being directly proportional to the eye velocity of the downbeat nystagmus. Such lesions also cause impaired slow phase of pursuit. This finding probably relates to the smooth tracking (either with head still [smooth pursuit] or fact that smooth pursuit is organized in, and superimposed with head moving), postsaccadic drift, and loss of some on, a phylogenetically old vertical ‘‘labyrinthine-optoki- adaptive capabilities, such as the ability to adjust the ampli- netic’’ coordinate system. Thus, for a pure vertical pursuit tude and direction of the VOR or the pulse-step (phasic- movement to occur, opposite torsional components must tonic) match for saccades. Unilateral lesions produce ipsi- cancel (as is the case for pure vertical vestibular nystagmus). lateral deficits in pursuit and gaze-holding (47,137,140). In The middle cerebellar peduncle probably carries information patients with cerebellar disease, pursuit defects with the head to and from the cerebellum (perhaps between the flocculus still, defects in combined eye and head tracking, and gaze- and the nucleus reticularis tegmenti pontis in the pons) and holding deficits frequently occur together, reflecting their contains vertical pursuit signals encoded with a torsional common substrate in the flocculus and vestibular nuclei component (150). (139). Quantitatively, however, pursuit with the head still is Other signs that occur in patients with lesions restricted sometimes relatively more impaired (140,141). Patients with to the cerebellum cannot always be attributed to dysfunction cerebellar disease may show timing errors during tracking of a particular part of the cerebellum. They include square- of periodic targets (141), but there is some preservation of wave jerks (151,152), alternating skew deviation (67,104), predictive capability (142). The ability to generate anticipa- cross-coupled VOR responses (inappropriately directed slow tory smooth eye movements of high speed at the onset of phases) (147,153,154), disconjugate (poorly yoked) sac- tracking is also impaired in some cerebellar patients (143). cades with disconjugate gaze-evoked nystagmus (104,126) Experimental lesions of the nodulus lead to an increase and divergent nystagmus (155), centripetal nystagmus (156), in the duration of vestibular responses that predisposes the primary position upbeating nystagmus (157), positional nys- animal to the development of periodic alternating nystagmus tagmus (158), increased responsiveness of the cervico-ocular (144). Other abnormalities of the ‘‘velocity-storage mecha- reflex (159), and impaired responses to linear translation (L- nism’’ also are present, including abnormally directed slow VOR) (154,160,161). SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 915

The cerebellum is also important in long-term adaptive Table 19.2 functions that keep eye movements appropriate to the visual Eye Signs in the Arnold-Chiari Malformation stimulus (162). Experimentally, it has been shown that con- trol of VOR amplitude and direction, saccade amplitude, Downbeat nystagmus (occasionally with a torsional component), worse on pursuit initiation, and phoria adaptation are all under cerebel- lateral gaze Sidebeat nystagmus (primary position, unidirectional, horizontal nystagmus) lar control. For example, adaptation of the gain of the VOR Periodic alternating nystagmus is impaired in patients with cerebellar lesions (163). This Divergent nystagmus adaptive or ‘‘repair shop’’ function of the cerebellum proba- bly accounts for both the enduring nature of the ocular motor Gaze-evoked nystagmus deficits that accompany diffuse cerebellar lesions and, per- Rebound nystagmus including torsional rebound haps, the somewhat variable effects of cerebellar lesions. Impaired pursuit (and VOR cancellation) Thus, inherent, idiosyncratic abnormalities in brain stem or Impaired OKN with slow build-up of eye velocity in response to a constant peripheral ocular motor mechanisms that are normally ‘‘re- velocity stimulus paired’’ by the cerebellum may reappear after cerebellar le- Convergence nystagmus sions. For example, some patients with cerebellar disease Divergence Skew deviation accentuated or alternating on lateral gaze may not be able to adapt to a phoria induced by wearing Saccadic dysmetria prisms (164,165), and the same is true for monkeys with Internuclear ophthalmoplegia cerebellar vermis lesions (126). Children with cerebellar le- Increased VOR gain sions often make better recoveries than adults (166). If cere- Shortened VOR time constant bellar ablation is performed in neonatal monkeys, almost- Positional nystagmus complete recovery occurs, provided the deep cerebellar nu- clei are left intact; if they are not, gaze-holding and smooth OKN, optokinetic nystagmus; VOR, vestibulo-ocular reflex. pursuit never fully recover (167).

ETIOLOGIES with other lesions located at the craniocervical junction (172). The positional nystagmus of posterior fossa lesions There are many conditions that produce cerebellar eye (173–175) must be differentiated from the more common signs. These include developmental anomalies, degenerative benign paroxysmal positional nystagmus of the labyrinth diseases, vascular diseases, and tumors. (147,158,176–181). The Dandy-Walker syndrome consists of a malformation Developmental Anomalies of the Hindbrain of the cerebellar vermis, a membranous cyst of the fourth The Arnold-Chiari malformation is an abnormality of the hindbrain involving the caudal cerebellum (including the vestibulocerebellum, flocculus, paraflocculus [tonsils], uvula, and nodulus) and the caudal medulla (168,169). In the type I malformation, the cerebellar tonsils are displaced caudally into the foramen magnum and the medulla is elon- gated. A meningomyelocele usually is not present. Such pa- tients often develop symptoms in adult life. In the type II malformation, both the fourth ventricle and the inferior ver- mis extend below the foramen magnum; the brain stem and spinal cord are thin; and a lumbar meningomyelocele is usu- ally present. Patients with type II malformation usually pres- ent in childhood, but in milder cases, the onset of symptoms is delayed until adulthood. Presenting symptoms of the Ar- nold-Chiari malformation include that is brought on or exacerbated by head movements, and dizziness, ver- tigo, cervical pain, and headaches, all of which can be brought on by Valsalva maneuvers. A variety of ocular motor abnormalities, and especially downbeat nystagmus (both spontaneous and positional), occur in patients with the Arnold-Chiari malformation (Table 19.2). Many of these signs are reproducible by vestibulocerebellar lesions in mon- keys (170). Diagnosis is by MR imaging, with sagittal views of the craniocervical junction (Fig. 19.7). Patients often im- Figure 19.7. Neuroimaging of an Arnold-Chiari malformation. T1- prove after suboccipital decompression, although it may take weighted sagittal magnetic resonance image shows herniation of cerebellar months for the eye movement abnormalities to diminish tonsils below the foramen magnum (arrowhead). Note flattening of the (171). A similar ocular motor syndrome can occur in patients brain stem in this region. 916 CLINICAL NEURO-OPHTHALMOLOGY ventricle, and malformations of the cerebellar cortex and cerebellar eye signs, including downbeat nystagmus, and deep cerebellar nuclei. Patients with this condition often many are mapped to the same area of chromosome 19 in show a mild saccadic dysmetria, although some patients which hemiplegic migraine and SCA6 are mapped have normal eye movements (182). Ocular motor abnormali- (200–202). Such syndromes may be associated with mi- ties, including nystagmus and strabismus, also occur in pa- graine, essential tremor, or myokymia; many are channel- tients with agenesis of the vermis (183) or hypoplasia of the opathies (203,204). entire cerebellum (184). Other rare syndromes associated with anomalous cerebellar development include Coffin-Siris Vascular Diseases syndrome (developmental delay, hypotonia, cutaneous The cerebellum is supplied by three branches of the verte- changes, and abnormalities of the roof of the fourth ventri- brobasilar circulation: the posterior inferior cerebellar artery cle) (185) and Joubert’s syndrome (a variable combination (PICA), the AICA, and the superior cerebellar artery (SCA). of episodic tachypnea, psychomotor retardation, retinal dys- Occlusion of one or more of these vessels often produces trophy, torsional nystagmus, skew deviation, ocular motor brain stem infarction too, making precise clinicopathologic apraxia, agenesis of the cerebellar vermis, and fibrosis of correlation difficult. Infarction in the distribution of the dis- the ) (186–188). tal PICA may cause acute vertigo and nystagmus that often simulates an acute peripheral vestibular lesion (205). These Degenerative Diseases symptoms probably reflect a central imbalance in horizontal VOR pathways created by asymmetric infarction of the ves- Many degenerative processes can affect the cerebellum tibulocerebellum. The vestibulocerebellum also influences or its connections and produce cerebellar eye signs. The the gaze-holding networks within the vestibular nuclei, and hereditary show considerable variability in pheno- patients with lesions in the vestibulocerebellum may have typic expression. As a result, there have been difficulties prominent gaze-evoked nystagmus that helps differentiate with classification, a problem that is being partially resolved this cerebellar lesion from an acute peripheral vestibu- with the discovery of genetic markers and the realization lopathy. that phenotype is influenced by both primary and modifier genes (189–195). Moreover, many of these conditions also affect brain stem structures. Thus, other, presumably nonce- rebellar, ocular motor signs may be present (e.g., slow sac- cades, prolonged saccadic latencies, decreased or absent ves- tibulo-ocular responses, and ophthalmoplegia). It remains to be proved whether eye movement abnormalities can be used to reliably detect extracerebellar involvement or early signs of disease in persons at risk for developing hereditary ataxias In general, patients with Friedreich’s ataxia show a decrease of vestibulo-ocular responses and prominent square-wave jerks. Slow saccades point to brain stem involvement, as occurs with SCA2. In one form of a recessively inherited cerebellar degeneration with and prominent square-wave jerks (saccadic intrusions) and sac- cadic dysmetria, the speed of large saccades was actually too fast (196). In general, the gross amount of atrophy of different parts of the cerebellum as shown by MR imaging does not correlate well with the severity of ocular motor dysfunction (197), although more sensitive imaging analysis techniques of individual lobules will likely reveal better clin- ical-anatomic correlations. Paraneoplastic cerebellar degeneration is a rare remote effect of cancer, usually occurring with breast, ovarian, or small-cell lung cancer (198,199). The onset of symptoms is usually acute or subacute, with severe midline and appendic- ular ataxia, dysarthria, and downbeat nystagmus. Autopsies show a total loss of Purkinje cells. Such patients have lost the output from the cerebellar cortex, and the common finding of primary position downbeat nystagmus is compatible with Figure 19.8. Neuroimaging of infarct in the territory of the superior cere- the hypothesis that a preponderance of inhibitory projections bellar artery (SCA) in a 69-year-old man with hypertension. Computed of the cerebellum to the central connections of the superior tomographic scan shows a large hypodense area in the left cerebellar hemi- semicircular canals can cause this nystagmus. sphere corresponding to the distribution of the left SCA. The patient had The episodic vertigo and ataxia syndromes, often respon- a left horizontal gaze palsy from compression of the left side of the brain sive to acetazolamide, may be associated with prominent stem. SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 917

Infarction in the territory of the AICA, the branches of compression is upward or directly forward, respectively. The which often supply the flocculus, may cause vertigo, vomit- oculomotor, trochlear, and abducens nerves may also be af- ing, hearing loss, facial palsy, and ipsilateral limb ataxia, fected. Ocular motor dysfunction may also be caused by along with gaze-holding and pursuit deficits as well as ves- secondary obstructive and increased intracra- tibular nystagmus (61–63,206) (Fig. 19.6). The ocular motor nial pressure. signs reflect a combination of involvement of the labyrinth, Medulloblastomas arising in the posterior medullary vestibular nuclei, and flocculus. velum frequently produce or are associated with positional Infarction in the territory of the SCA causes ataxia, limb nystagmus. Involvement of the nodulus and uvula is presum- dysmetria, and vertigo (206,207) (Fig. 19.8). A characteristic ably responsible for this finding (213) and may also account abnormality is saccadic contrapulsion. Contralateral sac- for the inability to suppress postrotational nystagmus by tilt- cades overshoot and ipsilateral saccades undershoot the tar- ing the head (145,146). Tumors within the fourth ventricle get. Attempted vertical saccades are oblique, with a horizon- may affect the cerebellar nuclei, vestibulocerebellum, and tal component away from the side of the lesion (208–210). dorsal medulla. Upbeating nystagmus may occur in such Thus, this saccadic disorder is the opposite of the saccadic cases. ipsipulsion seen in Wallenberg’s syndrome (discussed previ- Vestibular schwannomas may compress the cerebellar ously) and probably reflects interruption of outputs from the flocculus (which lies in the cerebellopontine angle) and pro- fastigial nucleus running in the uncinate fasciculus next to duce eye signs of vestibulocerebellar lesions (214), including the superior cerebellar peduncle (121,211). Infarction re- Brun’s nystagmus, in which there is a coarse nystagmus stricted to the posterior-inferior vermis can selectively im- beating to the side of the lesion (reflecting a gaze-holding pair pursuit and optokinetic eye movements (212). deficit) and a fine nystagmus beating away from the side of the lesion (reflecting a vestibular imbalance). Head-shaking Mass Lesions nystagmus may be present (215). MR imaging after intrave- Cerebellar hemorrhage, tumors, infarcts, abscesses, cysts, nous injection of contrast is the most sensitive and specific and extra-axial hematomas may all cause cerebellar eye method used to detect small lesions in this region. Acute signs by direct damage to the cerebellar parenchyma. Cere- cerebellar hemorrhage frequently causes nystagmus, gaze bellar lesions, however, may also compress the brain stem palsy (usually toward the side of the lesion), and produce additional signs. Vertical or horizontal gaze palsy, and skew deviation (216–218). These signs are, in disorders can occur, depending on whether the direction of part, caused by compression of the brain stem.

OCULAR MOTOR SYNDROMES CAUSED BY LESIONS OF THE PONS LESIONS OF THE INTERNUCLEAR SYSTEM: speed (determined by the high-frequency ‘‘pulse’’ of inner- INTERNUCLEAR OPHTHALMOPLEGIA vation) is diminished out of proportion to the limitation in range of adduction (determined by the low-frequency Among the fibers that make up the MLF, many carry a ‘‘step’’ of innervation). An adduction lag is brought out clin- conjugate horizontal eye movement command from abdu- ically by asking the patient to make large-amplitude (which cens internuclear neurons in the pons to the medial rectus subdivision of the contralateral oculomotor nuclear complex in the midbrain. Other fibers in the MLF carry signals for holding vertical eye position, for vertical smooth pursuit, and for the vertical VOR.

Manifestations Lesions of the MLF produce INO (219,220) (Fig. 19.9). When the lesion is unilateral, the INO is characterized by weakness of adduction ipsilateral to the side of the lesion (Fig. 19.10). This weakness may vary from a complete loss of adduction beyond the midline to a mild decrease in the velocity or acceleration of adduction without any limitation in range of motion (Fig. 19.11). The fibers subserving hori- zontal gaze in the MLF each carry commands for all types of conjugate eye movements. Hence, vestibular slow phases, pursuit and optokinetic following, and saccades and quick phases of nystagmus are all affected by the MLF lesion. The weakness of adduction, however, may be more obvious for Figure 19.9. Neuroimaging in a patient with a left internuclear ophthal- saccades because damaged axons, especially those that are moplegia. T2-weighted magnetic resonance image, axial view, shows a tiny demyelinated, show a greater defect for carrying high- rather area of hyperintensity consistent with an infarct in the region corresponding than low-frequency impulses. This dissociation is reflected to the location of the left medial longitudinal fasciculus (arrowhead). The in a ‘‘pulse-step mismatch’’ that occurs because saccade patient also had a skew deviation. 918 CLINICAL NEURO-OPHTHALMOLOGY

Figure 19.10. Unilateral, right internuclear ophthalmoplegia in a 32-year-old man with multiple sclerosis. Note complete lack of adduction in the right eye on attempted left horizontal gaze. require the highest speeds) horizontal saccades back and parity that occurs when a unilateral INO is associated with forth across the midline or by using an ‘‘optokinetic’’ tape a skew deviation also may interfere with convergence effort. or drum with repetitive symbols to produce nystagmus that In some patients with an INO, abducting saccades in the allows easy comparison of the movements of the two eyes affected eye may also be slow or ‘‘fractionated’’ (226,227). (221). Quantitative recordings of eye movements in patients This phenomenon may reflect impaired inhibition of the af- with MS show that the most sensitive sign of adduction fected medial rectus, although Kommerell (228) was unable weakness in INO is a decreased ratio either of peak eye to find any evidence of impaired medial rectus inhibition in velocity (222) or of peak acceleration (223) of the adducting one patient with a unilateral INO in whom he performed saccades of the eye on the side of the lesion to the abducting electromyography. Slowing of abducting saccades tends to saccades of the other eye (224). be more prominent in bilateral INO, probably because of When patients with INO are able to converge, despite damage to extra-MLF pathways running through the pontine absence of voluntary adduction, the lesion is located cau- tegmentum (229,230). dally with preservation of the medial rectus subdivision of The second cardinal sign of an INO is nystagmus in the the oculomotor nuclear complex (225). Patients with INO contralateral eye when it is abducted. This nystagmus con- and intact convergence were said to have a posterior INO sists of a centripetal (inward) drift, followed by a corrective by Cogan (77). Although the presence of intact convergence saccade that may be hypermetric, hypometric, or orthomet- is important in such cases, the absence of convergence in ric. It is present in nearly all patients with INO (231). A the setting of an INO (the ‘‘anterior’’ INO of Cogan) does number of mechanisms could account for the abduction nys- not necessarily imply a rostral lesion involving the medial tagmus of INO (220) and may not be mutually exclusive. rectus nuclear subdivision. Some patients simply are not able They include (a) an increase in convergence tone; (b) im- to produce a strong convergence effort, and the vertical dis- paired inhibition of the medial rectus contralateral to the

Figure 19.11. Ocular motor recording of a pa- tient with a mild right internuclear ophthal- moplegia. On attempted left horizontal gaze, the velocity of the adducting right eye is less than the velocity of the abducting left eye, which overshoots the target and shows mild abducting nystagmus. SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 919 lesion (232); (c) interruption of descending internuclear fi- bers that project to the abducens nucleus (233); (d) a super- imposed gaze-evoked nystagmus; and (e) adaptation to the contralateral medial rectus weakness (234–236). Because abduction nystagmus is not observed in acute experimental INO induced by injection of lidocaine into the MLF of mon- keys (237), the cause of abduction nystagmus must relate either to lesions outside the MLF or to an adaptive response to the initial adduction weakness. Patching experiments in patients with INO show that in many cases the abducting nystagmus diminishes if the pa- tient is required to habitually view with the nonparetic eye (235,236). In these patients, the abducting nystagmus relates, at least in part, to a conjugate (because of Hering’s law of equal innervation to the two eyes) adaptive change in innervation that acts to improve adduction of the paretic eye, but at the expense of a disturbance in abducting saccades of the contralateral eye. Abducting saccades overshoot the tar- get and are followed by a centripetal drift (‘‘pulse-step mis- match’’), giving the appearance of an abducting nystagmus. Such nystagmus usually dies out after a beat or two, and the postsaccadic drift is brief. It would bring the eye to a stable eccentric position if it were not interrupted by corrective eccentric saccades. Figure 19.12. Skew deviation in a patient following an operation on the The other common mechanism for abduction nystagmus posterior fossa for a superiorly placed vermis tumor. The deviation persisted is a dissociated gaze-evoked nystagmus that appears more for 3 weeks. prominent in the abducting eye because the adducting eye is weak. In contrast to the abducting nystagmus produced by adaptation, such gaze-evoked abduction nystagmus is more sustained, and the postsaccadic drift would bring the eye to patients with bilateral INO cannot fix steadily. In such pa- the primary position if it were not interrupted by corrective tients, sporadic bursts of monocular abducting saccades may eccentric saccades. Interruption of the paramedian tracts that occur in each eye (243). run near the MLF and carry fibers to and from the flocculus Patients with bilateral INO have bilateral adduction weak- may be responsible for the gaze-evoked nystagmus that is ness and abducting nystagmus (Figs. 19.13 and 19.14). Such associated with INO (238). patients also have impaired vertical vestibular and pursuit Skew deviation commonly occurs with unilateral INO but eye movements and impaired vertical gaze-holding with rarely with bilateral INO. When skew deviation is associated gaze-evoked nystagmus on looking up or down (244–247). with INO, the higher eye is usually on the side of the lesion Many patients with an INO have no visual symptoms, (Fig. 19.12). It is usually easy to differentiate skew deviation particularly when there is no limitation of adduction (231). of an INO from trochlear nerve palsy because of the ad- In other cases, the presence of either limitation of adduction ducting weakness and abducting nystagmus, but at times or skew deviation may cause that is horizontal, ver- skew deviation and trochlear nerve palsy may be hard to tical, or oblique. Patients with INO occasionally complain separate (68). Skew deviation may reflect imbalance of oto- of oscillopsia (248). Horizontal oscillopsia usually occurs lith inputs that cross in the medulla and ascend in the MLF from either the adduction lag or the abduction nystagmus, (discussed previously). whereas vertical oscillopsia occurs during head movements Dissociated vertical nystagmus (downbeat in the ipsi- and is caused by a deficient vertical VOR (249). In many lateral eye, torsional in the contralateral eye) may occur with patients, visual symptoms become less bothersome or re- an INO (239). This pattern of dissociated nystagmus reflects solve completely, either because of recovery of function of the fact that posterior semicircular canal pathways mediating MLF axons or because of more central adaptive mecha- excitation pass through the MLF, but some anterior semicir- nisms. cular canal pathways do not (240). Experimental INO pro- Lesions that damage the MLF may also damage the abdu- duced by lidocaine blockade causes ipsilateral hypertropia cens nucleus, fascicle, or both on either side of the brain and unilateral downbeating nystagmus (237). Patients with stem. Lesions that damage the MLF on one side and the a unilateral INO may also have an ipsiversive torsional nys- ipsilateral abducens nucleus produce the one-and-a-half syn- tagmus (top poles of the eyes cyclorotate so as to beat toward drome (discussed later), whereas lesions that damage the the side of the lesion) (241,242). The torsional nystagmus ipsilateral abducens fascicle produce horizontal ophthal- is sometimes dissociated and is usually but not always asso- moplegia in the ipsilateral eye from the combination of an ciated with a skew deviation. It may relate to interruption of INO and an abducens nerve palsy. Lesions that damage the pathways between the vestibular nuclei and the INC. Some MLF on one side and the paramedian pontine reticular for- 920 CLINICAL NEURO-OPHTHALMOLOGY

Figure 19.13. Bilateral internuclear ophthalmoplegia in a young woman with multiple sclerosis. The have been dilated with mydriatics.

mation (PPRF) or abducens nucleus on the opposite side Etiologies produce a horizontal gaze palsy toward the side of the dam- aged PPRF or abducens nucleus. In such cases, the INO Table 19.3 summarizes some causes of INO. In general, cannot be diagnosed because of the overriding horizontal a unilateral INO is most commonly caused by ischemia, gaze palsy. Damage to the MLF on one side and to the although there is often subtle involvement of the other side contralateral abducens nerve fascicle will produce abduction (Fig. 19.15). Bilateral INO is commonly caused by demye- lination associated with MS, but even in MS, the INO may be weakness of the contralateral eye combined with adduction quite asymmetric (250) (Fig. 19.16). Although MR imaging weakness of the ipsilateral eye. In this setting, there will be frequently shows a lesion in the MLF in patients with INO a ‘‘pseudo-horizontal gaze palsy’’ on attempted horizontal (250a) (Fig. 19.9), there are many exceptions (251,252). gaze away from the side of the MLF lesion. The diagnosis Thin cuts and sagittal T2-weighted imaging are sometimes may be suspected in a patient who appears to have a horizon- required. tal gaze palsy that is asymmetric, with one eye (usually the adducting eye) being more limited than the other. Posterior INO of Lutz As with patients having conjugate gaze palsy from dam- age to the abducens nucleus, patients with the one-and-a- In 1923, Lutz (253) described a condition in which abduc- half syndrome often have an associated facial nerve paraly- tion (not adduction) was impaired with respect to saccades sis. This paralysis is ipsilateral to the horizontal gaze palsy and pursuit but not with respect to vestibular stimulation. and occurs from damage to the fascicle of the facial nerve As noted above, Cogan (77) described an INO as one in as it loops around the abducens nucleus (discussed later). which there was slowed or limited adduction of one eye,

Figure 19.14. Ocular motor recording showing the saccadic velocities of a pa- tient with bilateral internuclear ophthal- moplegia. On attempted right horizontal gaze, the velocity of the adducting left eye is less than the velocity of the abducting right eye. The right eye overshoots the tar- get and shows abducting nystagmus. On at- tempted left horizontal gaze, the velocity of the adducting right eye is less than the velocity of the abducting left eye. The left eye overshoots the target and shows ab- ducting nystagmus. SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 921

Figure 19.15. Pathology of unilateral internuclear ophthalmoplegia. A, Section through the pons in a patient with a left internuclear ophthalmoplegia shows a small infarction in the region of the left medial longitudinal fasciculus (arrows). Sections are numbered from rostral to caudal. B, Area from the infarct indicated by the arrow in A and the surrounding region. The large arrowhead points to the infarct involving the left medial longitudinal fasciculus. The small arrowhead points to a small infarct just lateral to the right medial longitudinal fasciculus. C, Section through the rostral pons in a patient with a left internuclear ophthalmoplegia. There is a septic infarction involving the left medial longitudinal fasciculus (arrowheads). Many axons of the left medial longitudinal fasciculus have been destroyed, but the right medial longitudinal fasciculus is entirely normal. (A and B, From Cogan DG, Kubik CS, Smith JL. Internuclear ophthalmoplegia: a review of 58 cases. Arch Ophthalmol 1950;44Ϻ783–796. C, From Ross AT, DeMyer WE. Isolated syndrome of the medial longitudinal fasciculus in man. Arch Neurol 1966;15Ϻ203–205.)

associated with overshoot of the abducting eye. Because this to understand because the abducens nucleus contains neu- condition appeared to be the opposite of the INO that Lutz rons that innervate both lateral and medial rectus muscles for (253) had described, some authors began to refer to the eye all conjugate eye movements (219,257,258). Patients with movement abnormality observed by Lutz as the ‘‘posterior abducens palsy thus may show nystagmus in the contralat- INO of Lutz’’ and the eye movement abnormality observed eral adducting eye, if the weak abducting eye is used prefer- by Cogan (77) as the ‘‘anterior INO of Cogan.’’ This resulted entially for fixation (256). This nystagmus could reflect the in confusing terminology because Cogan (77) had his own same type of mechanisms (an adaptive response or a disso- anterior and posterior types of INO that were based on ciated nystagmus) that accounts for the abducting nystagmus whether convergence was present. In any event, the INO that develops in patients with typical INO. described by Cogan (77) is quite common, whereas the pos- C¸ elebisoy and Akyu¨rekli (259) described a patient in terior INO of Lutz is rare (254–256). whom a posterior INO of Lutz was mimicked by a rostral The pathogenesis of the posterior INO of Lutz is difficult brain stem infarction that caused a horizontal gaze paresis 922 CLINICAL NEURO-OPHTHALMOLOGY

Table 19.3 to one side and a pseudo-abducens paresis on attempted gaze Etiology of Internuclear Ophthalmoplegia to the other side. The patient also had bilateral vertical gaze palsy and other evidence of midbrain dysfunction. The Multiple sclerosis (commonly bilateral [223]); postirradiation demyelination mechanisms of posterior INO using examples with more Brain stem infarction (commonly unilateral), including complication of rostral brain stem lesions have been discussed (227). In other arteriography and hemorrhage (250) Brain stem and fourth ventricular tumors cases an infranuclear cause has been suggested (260). Arnold-Chiari malformation, and associated hydrocephalus and syringobulbia Infection: bacterial, viral, and other forms of meningoencephalitis LESIONS OF THE ABDUCENS NUCLEUS Hydrocephalus, subdural hematoma, supratentorial arteriovenous mal- Lesions of the abducens nucleus cause an ipsilateral palsy formation of horizontal conjugate gaze because the abducens nucleus Nutritional disorders: Wernicke’s encephalopathy and pernicious anemia (864) contains two groups of neurons: abducens motoneurons that Metabolic disorders: hepatic encephalopathy innervate the ipsilateral lateral rectus muscle, and abducens Drug intoxication: phenothiazines, tricyclic , narcotics, internuclear neurons that innervate the contralateral medial propranolol, lithium, , D-penicillamine, rectus motor neurons via the MLF. Vergence movements of Cancer: due to carcinomatous infiltration or remote effect the eyes are spared, however, so that adduction is possible Head trauma, and cervical hyperextension or manipulation with a near stimulus. Such a localized lesion can be produced Degenerative conditions: progressive supranuclear palsy experimentally (261), but it occurs rarely in humans Syphilis (262–264) except in Mo¨bius syndrome, a congenital brain Pseudo-internuclear ophthalmoplegia of myasthenia gravis and Fisher’s stem anomaly with horizontal gaze disturbances, often with syndrome an associated facial palsy (265). This condition occurs in at least some cases from aplasia or hypoplasia of both abducens nuclei. More often, however, the abducens nucleus is af- fected in association with adjacent tegmental structures, par- ticularly the genu of the facial nerve, the MLF, and the PPRF (229,266). Lesions restricted to the abducens nucleus can often be distinguished from those in the adjacent caudal PPRF (see also below), because only in the latter may pursuit and vestibular movements be spared, and only in the former may ipsilateral saccades in the contralateral field be spared by virtue of intact inhibition upon the contralateral abducens nucleus. Gaze-evoked nystagmus on contralateral gaze also occurs in patients with presumed abducens nucleus lesions (263). Possible mechanisms for the gaze-evoked nystagmus include damage to adjacent vestibular or NPH pathways that are involved in neural integration for gaze-holding, or dam- age to the paramedian cells and tracts that lie in part in the rostral abducens nucleus and have reciprocal connections with the cerebellar flocculus, a structure also involved in gaze-holding (238).

LESIONS OF THE PARAMEDIAN PONTINE RETICULAR FORMATION(PPRF) The PPRF, which corresponds principally to medial por- tions of the nucleus reticularis pontis caudalis, contains burst neurons that are important in the generation of saccades, and the paramedian nucleus raphe interpositus contains pause neurons that inhibit burst neurons at all times except during saccades. Destructive lesions of the PPRF, such as infarction and hemorrhage, tend to affect all cell groups, along with fibers of passage that convey pursuit and vestibular signals to the Figure 19.16. Brain stem pathology in a patient with a bilateral in- ipsilateral abducens nucleus. Unilateral destructive lesions ternuclear ophthalmoplegia and severe cerebrovascular and cardiovascular cause an ipsilateral, conjugate, horizontal, gaze palsy disease. Section at pontomesencephalic junction shows paramedian necrosis with areas of coalescing cavitation in the regions normally occupied (233,267,268). With acute lesions, the eyes may be deviated by both medial longitudinal fasciculi. This lesion involves both sides contralaterally. Nystagmus occurs when gaze is directed into of the midline. (From Gonyea EF. Bilateral internuclear ophthalmoplegia: the intact contralateral field of movement with quick phases association with occlusive cerebrovascular disease. Arch Neurol 1974; directed away from the lesioned side; this is usually accentu- 31Ϻ168–173.) ated in darkness. Ipsilaterally directed saccades and quick SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 923 phases are small and slow and do not carry the eye past the tal eye movements, but, particularly with chronic lesions midline. The degree of slowing of saccades directed toward such as tumors, reflexive eye movements may be spared. the lesioned side, when made in the intact field of gaze, Bilateral pontine lesions may impair vertical eye move- probably depends on whether projections to inhibitory burst ments (284–286). Signals for vertical vestibular and smooth neurons, which project to the contralateral abducens nucleus, pursuit eye movements ascend in the MLF and other path- are damaged. Recall that excitatory saccadic inputs reach ways through the pons; consequently, vestibular and pursuit the abducens nucleus from the ipsilateral population of exci- eye movements can be impaired with pontine lesions. Ab- tatory burst cells in the PPRF, whereas inhibitory saccadic normal (usually slow) vertical saccades occur in patients inputs originate from contralateral inhibitory burst cells in with discrete, bilateral pontine lesions (278,287–289) and the medulla. If the lesion is restricted to the ipsilateral abdu- in monkeys with bilateral neurotoxic lesions of the PPRF cens nucleus, saccades from the opposite field of gaze to the (280,290). Pause cells project rostrally to vertical burst neu- midline may be relatively rapid, because inhibition of the rons located in the midbrain. Pontine lesions, therefore, may antagonist is intact. If the PPRF is extensively damaged, lead to desynchronization of vertical (and horizontal) burst particularly in its more caudal part, inhibition is also affected neuron discharge and consequently to slow vertical (and hor- so that saccades beginning from the field of gaze opposite izontal) saccades. Ocular flutter has been reported with a to the lesioned side and moving to the midline are extremely PPRF lesion (291). slow. Rapid eye movements directed to the side opposite the lesion appear normal. Vertical saccades may be slightly COMBINED UNILATERAL CONJUGATE GAZE slow, and an inappropriate horizontal component, directed PALSY AND INTERNUCLEAR OPHTHALMOPLEGIA away from the side of the lesion, may occur during attempted (ONE-AND-A-HALF SYNDROME) vertical saccades (269). Smooth-pursuit movements and slow phases of optoki- Combined lesions of the abducens nucleus or PPRF and netic nystagmus may be preserved in both directions within adjacent MLF on one side of the brain stem cause an ipsi- the intact field of movement of a patient with a lesion of the lateral horizontal gaze palsy and INO, so that the only pre- PPRF, but they usually cannot bring the eyes across the served horizontal eye movement is abduction of the contra- midline. Sometimes horizontal pursuit is asymmetrically im- lateral eye; hence the name one-and-a-half syndrome (Figs. paired, more so for contralateral target motion (270). This 19.17–19.19) (18,229,282,292,293). Such patients may is in contrast to the effects of more basal lesions in the pons show an when attempting to look straight ahead; that impair ipsilateral or bilateral pursuit (271–273). Be- the eye opposite the side of the lesion is deviated outward. cause of the confluence of pursuit pathways in the brain This misalignment of the eyes is thought to be caused by stem and its cerebellar connections, the direction of a pursuit the unopposed drives of the intact pontine gaze center to deficit with a brain stem lesion is not reliable for determining the spared abducens nucleus and is called paralytic pontine the side of the lesion (274). In some patients with PPRF exotropia lesions, vestibular stimuli drive the eyes past the midline. (294). Many patients, however, have an esotropia Presumably, in such individuals, either the ipsilateral abdu- or no deviation in primary position (as with many cases of cens nucleus or its direct vestibular input is intact, or the INO), though it may be always present when fixation is PPRF lesion is more rostral (269,275). With more restricted eliminated (295). The spared abduction saccades of the con- lesions, usually hemorrhages, both smooth pursuit and the tralateral eye are followed by centripetal drift, so that a nys- VOR may be preserved, whereas saccades are absent or slow tagmus similar to that of the abducting eye in INO is present. (276). Occasionally in such patients, vestibular stimuli can Occasionally, the ipsilateral horizontal vestibular responses drive only the contralateral adducting eye and not the ipsi- are preserved when voluntary gaze is abolished (275, lateral abducting eye into the ipsilateral field. This finding 282,293), suggesting that the pontine lesion is more rostral implies a lesion of one PPRF and the ipsilateral abducens in the PPRF or more discrete in the caudal PPRF (296), thus nerve but sparing the abducens nucleus (277). sparing the vestibular projections to the abducens nucleus. Bilateral lesions restricted to the PPRF are uncommon. Although attempts at conjugate movements elicit no adduc- Discrete infarction (278) or tumor (279) can cause a selective tion, vergence movements may be preserved. Ocular bob- loss of saccades, leaving smooth pursuit and the VOR rela- bing may accompany the one-and-a-half syndrome (297). tively preserved. Such a selective deficit implies loss or dys- The one-and-a-half syndrome may result from brain stem function of saccadic burst neurons but sparing of fibers of ischemia (298–302), MS (292), tumor (303,304), hemor- passage conveying smooth pursuit and the VOR. Experimen- rhage (305), trauma (306), and other miscellaneous causes, tal lesions of the PPRF in monkeys, using neurotoxins that including HIV infection with tuberculosis, (302). Bilateral spare fibers of passage, may cause a similar deficit (280). INO with associated abducens nerve palsy also may mimic Patients with horizontal gaze palsies may substitute conver- the one-and-a-half syndrome, as can myasthenia gravis gence for impaired conjugate adduction and then cross-fixate (299). A ‘‘one-and-a-half syndrome’’ has been described in to extend their range of view (281). Furthermore, during which only adduction in one eye was spared (307). The pa- the recovery phase from bilateral gaze palsies from vascular tient had mucormycosis that caused an abducens nerve palsy lesions, involuntary synkinetic divergence and convergence from cavernous sinus infiltration on one side and a contralat- movements may appear with horizontal or vertical gaze eral horizontal gaze palsy from simultaneous carotid artery (282,283). Bilateral pontine lesions may abolish all horizon- occlusion. 924 CLINICAL NEURO-OPHTHALMOLOGY

Figure 19.17. One-and-a-half syndrome in a 23-year-old man with multiple sclerosis. Arrows indicate direction of attempted gaze. All ocular motor signs cleared within 3 months after onset.

Figure 19.18. Location of lesions producing the one-and-a-half syndrome. A, The important structures involved in the produc- tion of horizontal gaze. MLF, medial longitudinal fasciculus; PPRF, paramedian pontine reticular formation. Neurons project from the PPRF to the abducens nucleus and neurons in the abducens nucleus are both motoneurons whose axons represent the abducens nerve and internuclear neurons whose axons ascend in the contralateral MLF. B, The areas that may be involved when a one-and-a-half syndrome is present. Involvement of either the abducens nucleus or the PPRF can cause the horizontal gaze palsy. Damage to the ipsilateral medial longitudinal fasciculus produces the internuclear ophthalmoplegia. (Modified from Sharpe JA, Rosenberg MA, Hoyt WF, et al. Paralytic pontine exotropia: a sign of acute unilateral pontine gaze palsy and internuclear ophthalmoplegia. 1974;24Ϻ1076–1081.) SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 925

A

Figure 19.19. Brain stem findings in a 48- year-old hypertensive man with a recent myo- cardial infarction who developed a complete right horizontal gaze palsy and a complete B right internuclear ophthalmoplegia (one-and- a-half syndrome). Sections through the pons demonstrate an extensive area of hemorrhage. The drawings on the right illustrate the rela- tionship of the hemorrhage to the fourth ven- tricle, superior cerebellar peduncles, medial longitudinal fasciculi, medial tectospinal tracts, medial lemnisci, dentate nuclei (all outlined in heavy black lines), and the abdu- cens nuclei (black). A, Rostral pons. B, Mid- pons. C, Caudal pons. (From Pierrot-Deseil- ligny C, Chain F, Serdaru M, et al. The ‘‘one- and-a-half’’ syndrome: electro-oculographic analyses of five cases with deductions about the physiological mechanisms of lateral gaze. C Brain 1981;104Ϻ665–699.)

SLOW SACCADES FROM PONTINE LESIONS tients is not established. Conversely, if the saccades are so slow that they cannot bring the eye to the moving target, Slow saccades are characteristic of many degenerative and metabolic diseases (Table 19.4). Horizontal saccades may then even if pursuit is intact, it may not appear so because be slowed in patients with spinocerebellar degenerations; the eye seems to lag the target. Vestibular stimulation elicits vertical saccades are often relatively less affected in such normal compensatory slow phases, but quick phases of nys- patients. In diseases that principally affect the midbrain, such tagmus are slow and show approximately the same abnormal as progressive supranuclear palsy, vertical saccades are the relationship between amplitude and peak velocity, as do vol- first to become slow. In some patients with slow saccades, untary saccades. Patients with SCA2 usually make saccades blinks of the may actually speed up the movements of normal amplitude despite their low velocity. Progressive (308). The explanation for this phenomenon is uncertain, supranuclear palsy, however, causes both slow and small but it probably reflects the effects of blinks upon pause and horizontal saccades (discussed later). Patients with slow sac- burst neurons rather than a momentary deprivation of vision. cades may use a variety of strategies of combining eye and Some patients with slow saccades can generate smooth-pur- head movements to move their gaze more quickly to the suit movements of up to about 20Њ/sec. It is difficult to distin- target. guish a considerably slowed saccade from pursuit, however, Slow horizontal saccades may be caused by disease of so that whether pursuit function is truly intact in such pa- saccadic burst cells in the PPRF. Autopsy examinations of 926 CLINICAL NEURO-OPHTHALMOLOGY

Table 19.4 the of patients with SCA2 with slow saccades demon- Etiology of Slow Saccades strate a marked loss of neurons from the nucleus reticularis pontis caudalis and the adjacent raphe nuclei (309). Alterna- Olivopontocerebellar atrophy and related spinocerebellar degenerations tively, slowing of saccades may reflect abnormal inputs to Huntington’s disease the paramedian reticular formation from the cerebral hemi- Progressive supranuclear palsy Parkinson’s (advanced cases) and related diseases; Lytico-Bodig and spheres, the superior colliculus, or the cerebellum. Lesions Caribbean Parkinson’s in the pause cell region also may produce slow saccades Whipple’s disease (290). Lipid storage diseases Wilson’s disease SACCADIC OSCILLATIONS FROM PONTINE Drug intoxication: anticonvulsants, LESIONS Tetanus In dementia: Alzheimer’s disease (stimulus-dependent) and in association Saccadic oscillations that lack an intersaccadic interval, with AIDS such as opsoclonus and ocular flutter, probably result from Lesions of the paramedian pontine reticular formation pontine lesions in the pause cell region (310). A number Internuclear ophthalmoplegia of other mechanisms are possible, however. Macrosaccadic Peripheral nerve palsy, diseases affecting the neuromuscular junction and oscillations (an extreme form of saccadic dysmetria) also extraocular muscle, restrictive ophthalmopathy Paraneoplastic syndromes occur with pontine lesions (311), as can ocular flutter (291) (see Chapter 23).

OCULAR MOTOR SYNDROMES CAUSED BY LESIONS IN THE MESENCEPHALON SITES AND MANIFESTATIONS OF LESIONS tion is known by a variety of names: Parinaud’s syndrome, Koerber-Salus-Elschnig syndrome, pretectal syndrome, dor- Disturbances of vertical eye movements from midbrain sal midbrain syndrome, and the sylvian aqueduct syndrome lesions are usually caused by damage to one or more of (313,314). Unilateral midbrain lesions can also create the three main structures: the posterior commissure, the rostral same ocular motor syndrome, possibly by interrupting the interstitial nucleus of the medial longitudinal fasciculus, or afferent and efferent connections of the posterior commis- the INC (312). sure, or possibly by disturbing supranuclear inputs from the Posterior Commissure nucleus of the posterior commissure directly to the midbrain oculomotor nuclei (discussed later). Although paralysis of Lesions of the posterior commissure cause a syndrome upward gaze in this syndrome has, in the past, been ascribed characterized by loss of upward gaze and other associated to destruction of the superior colliculi, this is not the case. findings (Table 19.5 and Figs. 19.20 and 19.21). The condi- Experimental lesions restricted to the superior colliculus in nonhuman primates produce no limitation of upward gaze (315). Furthermore, the selective upgaze paralysis seen in Table 19.5 this syndrome cannot be explained on the basis of different Features of the Syndrome of the Posterior Commissure (aka Dorsal crossing patterns of premotor burst neurons originating in Midbrain Syndrome) the riMLF (downward premotor neurons are uncrossed to the Limitation of upward eye movements (Parinaud’s syndrome) third and fourth cranial nerve nuclei, while upward premotor Saccades neurons are both uncrossed and crossed), since none of these Smooth pursuit neurons appear to cross via the posterior commissure Vestibulo-ocular reflex (312,316). Although sparse data are available, damage to Bell’s phenomenon premotor neurons in the nucleus of the posterior commissure Lid retraction (Collier’s sign); occasionally that selectively subserve upward eye movements (or their Disturbances of downward eye movements axons, which appear to traverse the posterior commissure to Downward gaze preference (“setting sun” sign) arrive at their oculomotor targets) may be involved in the Downbeating nystagmus genesis of the selective loss of upward eye movements in Downward saccades and smooth pursuit may be impaired, but vestibular movements are relatively preserved. the syndrome of the posterior commissure (317). Disturbances of vergence eye movements The vertical gaze deficit caused by lesions of the posterior Convergence-retraction nystagmus (Koerber-Salus-Elschnig syndrome) commissure usually affects all types of eye movements, al- Paralysis of convergence though the VOR and Bell’s phenomenon are sometimes Spasm of convergence spared. Below the horizontal meridian, vertical saccades can Paralysis of divergence be made but are usually slow. Acutely, the eyes may be “A” or “V” pattern exotropia tonically deviated downward (setting-sun sign); this finding Pseudo-abducens palsy is prominent in premature infants with intraventricular hem- Fixation instability (square-wave jerks) orrhage (318). Transient downward deviation of the eyes Skew deviation Pupillary abnormalities (light–near dissociation) occasionally occurs in healthy neonates, but in such cases, the eyes can be easily driven above the horizontal meridian SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 927

Figure 19.20. Appearance of a patient with Parinaud’s dorsal midbrain syndrome. Above left, The eyes are straight in primary position. Note bilateral upper retraction. Above right, Downward gaze is normal. Below center, There is marked limitation of upward gaze bilaterally, worse on the left, producing a right hypertropia. (From Bajandas FJ, Aptman M, Stevens S. The sylvian aqueduct syndrome as a sign of thalamic vascular malformation. In: Smith JL, ed. Neuro-Ophthalmology Focus 1980. New York: Masson, 1979Ϻ401–406.) by the vertical doll’s head maneuver (86). Tonic upward The dorsal midbrain syndrome is also characterized by deviation of the eyes occurs in some patients with midbrain disturbances of horizontal eye movements. In some patients lesions (319). Oculogyric crises may also occur in patients convergence is paralyzed, whereas in others it is excessive with midbrain lesions and are discussed below. Episodic and causes convergence spasm. During horizontal saccades, (paroxysmal) tonic up-gaze may also occur in otherwise nor- the abducting eye may move more slowly than its adducting mal infants, in patients with other neurologic deficits, includ- fellow. This finding is called pseudo-abducens palsy (313) ing ataxia, as a familial dominantly inherited disorder, and, and may reflect an excess of convergence tone (328). It may occasionally, with structural lesions in the midbrain be responsible for an early symptom of evolving posterior (320–327). commissure lesions—reading difficulty caused by a tran- sient inability to find, and to focus both eyes on, the begin- ning of the next line when a horizontal saccade is made. Convergence-retraction nystagmus is another sign often evident in patients with disease affecting the dorsal midbrain (329,330) (Fig. 19.22). Experimental lesions suggest that this eye movement disturbance results from damage to the posterior commissure (331). Originally thought to be a sac- cadic disorder consisting of asynchronous, opposed sac- cades, recent quantitative measures suggest that the ad- ducting movements are abnormal convergence movements (332). Abnormalities of the eyelids occur in patients with dorsal midbrain lesions. The most common is eyelid retraction (Collier’s tucked lid sign), but ptosis occasionally occurs. In some patients, the abnormalities of the lids may be more prominent than those of the eye movements (333,334). Pupillary reactions are also commonly affected in patients Figure 19.21. Dorsal midbrain syndrome. A, Gaze straight ahead. B,On with lesions in the region of the posterior commissure. The attempted upward gaze, the patient develops convergence-retraction nys- pupils are usually large and react better to an accommodative tagmus. stimulus than to light (i.e., light-near dissociation). 928 CLINICAL NEURO-OPHTHALMOLOGY

Figure 19.22. Dorsal midbrain syndrome. A, Position of eyes in primary position. B, Retraction on attempted upward gaze. C, Lateral view of right eye when the patient is looking straight ahead. D, On attempted upward gaze, obvious retraction of the occurs. (From Lyle DJ, Mayfield FH. Retraction nystagmus: a case report. Am J Ophthalmol 1954;37Ϻ177–182.)

Many disease processes can affect the region of the poste- duce this syndrome by enlarging the aqueduct and third ven- rior commissure (Table 19.6 and Figs. 19.23–19.25). Pineal tricle or the suprapineal recess, thus stretching or compress- tumors (335,336) produce the dorsal midbrain syndrome ing the posterior commissure (337,338). Shunt dysfunction either by direct pressure on the posterior commissure or by may produce Parinaud’s syndrome before dilation of the causing obstructive hydrocephalus. Hydrocephalus may pro- ventricles is apparent on neuroimaging or measures of intra-

Table 19.6 Etiology of Disorders of Vertical Gaze

Tumor: Classically, pineal germinoma or teratoma in an adolescent male; also pineocytoma, pineoblastoma, glioma, metastasis Hydrocephalus: Usually aqueductal stenosis leading to dilatation of the third ventricle and aqueduct or enlargement of the suprapineal recess with pressure on the posterior commissure Vascular: Midbrain or thalamic hemorrhage or infarction Metabolic: Lipid storage disease: Niemann-Pick variants Drug-induced Degenerative: Progressive supranuclear palsy, cortical basal degeneration, Lytico-Bodig diffuse Lewy body disease; miscellaneous degenerations Miscellaneous: Multiple sclerosis, Whipple’s disease, hypoxia, , Figure 19.23. Pathology of dorsal midbrain syndrome. Contusion hemor- syphilis, , neurosurgical procedure, mesencephalic clefts, rhages are present in the region of the oculomotor nucleus and posterior tuberculoma, trauma, benign transient form of childhood commissure after a fall that resulted in an impact to the right side of the vertex. (Courtesy of Dr. Richard Lindenberg.) SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 929

Figure 19.24. Metastatic malignant mela- noma causing a dorsal mid-brain syndrome. Horizontal section through the inferior third ventricle (V) shows replacement of posterior commissure (PC) by metastatic melanoma. CC, . (From Keane JR, Davis RL. Pretectal syndrome with metastatic mela- noma to the posterior commissure. Am J Ophthalmol 1976;82Ϻ910–914.)

Figure 19.25. Locations of lesions pro- ducing isolated paralysis of upward gaze. Left, Four lesions giving rise to upward gaze paralysis have been superimposed on a sagittal section of the human brain stem. The enlargement of the lesions on the right shows that the common areas involved (stippled regions) include the in- terstitial nucleus of Cajal (iC), the nucleus of Darkschewitsch (nD), and fibers of the posterior commissure (PC). IC, inferior colliculus; MT, mammillothalamic tract; riMLF, rostral interstitial nucleus of the medial longitudinal fasciculus; RN, red nucleus; SC, superior colliculus; T, thala- mus; TR, tractus retroflexus; III, oculo- motor nerve. (Redrawn from Bu¨ttner-En- never JA, Bu¨ttner U, Cohen B, et al. Vertical gaze paralysis and the rostral in- terstitial nucleus of the medial longitudi- nal fasciculus. Brain 1982;105Ϻ125– 149.) 930 CLINICAL NEURO-OPHTHALMOLOGY cranial pressure are consistently high (339). An intermittent by individual burst neurons within the left riMLF nucleus Parinaud’s syndrome dependent on head position was de- only). scribed in a patient with a posterior fossa subdural hematoma Because infranuclear fibers innervating the superior (340). oblique (fourth) and superior rectus (third) muscles are crossed projections, each riMLF burst neuron for downward Rostral Interstitial Nucleus of the Medial Longitudinal saccades can control the movement of both eyes without Fasciculus sending axon collaterals across the midline. Thus, unilateral midbrain lesions in the riMLF can produce selective loss of The rostral interstitial nucleus of the medial longitudinal downward saccades. fasciculus (riMLF), which lies in the prerubral fields of the The same burst neurons that lack right–left asymmetry mesencephalon, contains the burst neurons that generate ver- with respect to vertical saccades are ‘‘lateralized’’ with re- tical and torsional saccades (analogous to the PPRF in the spect to ‘‘directionality’’ of torsional saccades; that is to say, pons for horizontal saccades). The riMLF lies dorsomedial all of the right riMLF vertical-torsional burst neurons drive to the rostral half of the red nucleus, medial to the fields of saccades toward the right shoulder (top pole rotating to the Forel, lateral to the periventricular gray and the nucleus of right), and all of the left riMLF vertical-torsional burst neu- Darkschewitsch, and immediately rostral to the INC. The rons drive saccades toward the left shoulder. Thus, unilateral riMLF projects to ipsilateral more than contralateral third midbrain lesions in the riMLF can also produce selective loss and fourth nerve nuclei, the INC (important for vertical gaze- of ipsilesional torsional saccades. The associated left–right holding [discussed later]), and caudal targets, including the asymmetry in tonic firing rates between the two riMLF nu- cell groups of paramedian tract, that relay ocular motor sig- clei may explain both the contralesional-beating torsional nals to the cerebellum. Although the right and left riMLF nystagmus (a sign of vertical VOR imbalance), and the tonic nuclei are interconnected, the projections are now believed ocular torsion toward the contralesional side (a sign of oto- to course ventral to the cerebral aqueduct rather than via the lith ocular imbalance). posterior commissure, as thought previously (312,316). Clinically, bilateral lesions of the riMLF are more com- Given this information, it is not unexpected that bilateral mon than unilateral lesions. They generally cause either loss experimental lesions of the riMLF abolish all vertical and of downward saccades or loss of all vertical saccades (345) torsional saccadic movements (341,342). However, unilat- (Figs. 19.26–19.28). Lesions of the riMLF are usually in- eral experimental lesions of the riMLF produce slowed farcts in the distribution of a small perforating vessel (the downward saccades, loss of ipsilesional torsional quick posterior thalamo-subthalamic paramedian artery) that arises phases of the VOR, contralesional beating torsional nystag- between the bifurcation of the basilar artery and the origin of mus, and tonic contralesional ocular torsion (343). An ana- the posterior communicating artery (Fig. 19.28). This vessel tomic explanation for these findings follows. may be paired or single (346,347); it supplies structures that Burst neurons within the riMLF have mixed vertical and include the riMLF, the rostromedial red nucleus, the adjacent torsional vectors rather than subpopulations of neurons with subthalamus, the posterior inferior portion of the dorsome- either vertical or torsional vectors, echoing the anatomic dial nucleus, and the parafascicular nucleus of the thalamus. structure of the vertical semicircular canals (presumably Some investigators think that paralysis of downward sac- since the vestibulo-ocular system is the phylogenetically- cades results from lateral lesions of both riMLFs, whereas older substrate underlying the saccadic system (344)). The medial bilateral lesions cause a total palsy of vertical sac- net vector sum of various burst neurons firing may, however, cades (Figs. 19.28–19.32) (348). Others (349), however, lead to purely vertical or purely torsional saccadic eye move- have observed that downward gaze paresis is more often ments. Burst neurons within the riMLF subserving upward- associated with medial lesions than with lateral ones. Al- torsional saccades appear to be redundant (i.e., project to though vertical smooth pursuit and the VOR may be affected both ipsilateral and contralateral third and fourth nerve nu- with lesions in this area, this probably reflects damage to clei), while those subserving downward-torsional saccades nearby structures, such as the MLF and INC (350). Deleu are not (i.e., project to the ipsilateral third and fourth nerve et al. (351) reported a patient with a vertical one-and-a-half nuclei only (312,316)). Each vertical-torsional saccade burst syndrome. The patient had loss of all downward movements neuron sends axon collaterals to yoke muscle pairs that sub- and selective loss of upward movements in one eye. A verti- serve the same pulling direction in each eye; specifically, cal one-and-a-half syndrome in which there was impairment these pairings are as follows: (a) right superior rectus and of all upward eye movements and a selective deficit of down- left inferior oblique (serving upward-leftward torsional sac- ward saccades in the eye on the side of the lesion has also cades, innervated by individual burst neurons within each been reported (352,353,354). Unilateral lesions of the mid- riMLF nucleus); (b) left superior rectus and right inferior brain can also produce combined complete up-gaze and oblique (serving upward-rightward torsional saccades, in- down-gaze palsy, isolated up-gaze palsy, or monocular ele- nervated by individual burst neurons within each riMLF nu- vator palsy (353). cleus); (c) right inferior rectus and left superior oblique Clinically, unilateral lesions of the riMLF generally pro- (serving downward-rightward torsional saccades, innervated duce a down-gaze palsy, mainly affecting saccades or, more by individual burst neurons within the right riMLF nucleus rarely, a complete vertical gaze palsy for all eye movement only); and (d) left inferior rectus and right superior oblique types (268,348,355,356). As seen with experimental lesions, (serving downward-leftward torsional saccades, innervated unilateral lesions of the riMLF also produce a contralaterally SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 931

Figure 19.27. Isolated paralysis of downward gaze in a 42-year-old alco- holic man. Above, Upward gaze is normal. Below, The patient cannot make voluntary downward eye movements. Bilateral simultaneous caloric stimu- lation with warm water induced downward nystagmus, but the excursion of the fast phases was truncated at the level of direct forward gaze. (From Jacobs L, Anderson PJ, Bender MB. The lesions producing paralysis of downward but not upward gaze. Arch Neurol 1973;28Ϻ319–323.)

Interstitial Nucleus of Cajal The INC, which lies adjacent and caudal to the riMLF in the midbrain reticular formation, contains premotor neurons that subserve vertical gaze-holding (analogous to the perihy- Figure 19.26. Neuroimaging of a lesion in a patient with bilateral conju- poglossal nucleus in the medulla for horizontal gaze) (312). gate vertical saccadic gaze palsy and cognitive dysfunction. T2-weighted The INC projects to contralateral (more than ipsilateral) third axial magnetic resonance image shows a hyperintense midline lesion (ar- rowhead) affecting the region of the rostral interstitial nucleus of the medial and fourth nerve nuclei, the riMLF (important for vertical longitudinal fasciculus. and torsional saccades [discussed previously]), and caudal targets, including the cell groups of paramedian tract, which relay ocular motor signals to the cerebellum, as well as the ipsilateral vestibular and perihypoglossal nuclei (subserving beating torsional nystagmus, a tonic torsional deviation to horizontal gaze-holding), and the ipsilateral C1-C2 ventral the contralateral side, and a deficit in generating ipsilaterally horns (involved in eye–head coordination), among others directed torsional quick phases (i.e., top pole rolling toward (312,316,360). The right and left INC are interconnected via the side of the lesion) (357,358). As pointed out above, the the posterior commissure. selective impairment of downward saccades with unilateral Given this information, it is not surprising that bilateral riMLF lesions appears to result from a lack of redundancy experimental lesions of the INC severely impair both vertical in the premotor pathways for down-gaze—burst neurons and torsional gaze-holding and disrupt normal vertical and subserving downward(-torsional) saccades project only to torsional VOR responses (361). Unilateral experimental le- the ipsilateral third and fourth nerve nuclei, while those sub- sions of the INC produce similar (though less profound) serving upward(-torsional) saccades project both ipsi- vertical and torsional gaze-holding deficits (vertical gaze- laterally and contralaterally (312,316). These apparent facts evoked nystagmus, and spontaneous, ipsilesional beating notwithstanding, Ranalli et al. (359) described a patient with torsional nystagmus), along with tonic contralesional ocular a unilateral lesion of the riMLF (which partially involved torsion (362). the adjacent INC) that led to a loss of saccades above the Lesions restricted to the INC are not well studied in hu- primary position and slow and limited saccades below. mans, apparently because isolated lesions of INC are rare Smooth pursuit and the VOR were also affected in the verti- (due to frequent involvement of riMLF). Nevertheless, avail- cal plane, being restricted in range and reduced in gain. The able evidence suggests that such lesions produce a clinical anatomic underpinnings of this clinical picture remain uncer- syndrome characterized by vertical gaze-evoked nystagmus, tain. impaired vertical smooth pursuit, tonic torsional deviation 932 CLINICAL NEURO-OPHTHALMOLOGY

Figure 19.28. Pathology of isolated paralysis of downward gaze. A, In a 58-year-old man who died of septic shock and was noted to have selective paralysis of downward gaze, the cut surface of the rostral mesencephalon reveals bilateral cavitary infarcts (arrows). The posterior commissure (PC) is intact. B and C, In a 42-year-old man who suffered an infarction of the brain stem and was noted to have selective paralysis of downward gaze, sections through the rostral mesencephalon show zones of bilateral infarction and necrosis involving portions of the tractus retroflexus and red nuclei on both sides. Zones of necrosis also involve portions of the intralaminar, dorsomedial, lateral, and posteromedial thalamic nuclei. D, Section through the mesencephalon at the level of the posterior commissure shows bilateral zones of infarction (arrows) in the matter. The posterior commissure itself and other portions of the pretectal region of the mesencephalon are normal. (A, From Trojanowski JQ, Wray SH. Vertical gaze ophthalmoplegia: selective paralysis of downgaze. Neurology 1980;30Ϻ605–610. B–D, From Jacobs L, Anderson PJ, Bender MB. The lesions producing paralysis of downward but not upward gaze. Arch Neurol 1973;28Ϻ319–323.) to the contralateral side, and torsional nystagmus tion of the INC in a human produced torsional nystagmus (110,358,363). The torsional nystagmus has been variably and the OTR (120). reported as contralesional beating or ipsilesional beating (110,358). Recent studies suggest that the apparent confu- Periaqueductal Gray Matter and Mesencephalic sion may stem from the comorbid presence of riMLF damage Reticular Formation in those cases where contralesional-beating nystagmus is present (358). A jerk seesaw nystagmus may also occur in The effects of lesions localized to mesencephalic struc- patients with lesions that damage the INC (364), although tures other than the posterior commissure, INC, and the seesaw nystagmus has not been reproduced in the monkey riMLF are less clear. The periaqueductal gray matter of the with INC lesions (365). The INC may also contribute to mesencephalon contains neurons that stop discharging dur- dynamic properties of the VOR, especially its low-frequency ing saccades. A patient with bilateral lesions of the periaque- behavior, by virtue of the effect on the vestibular velocity ductal gray matter, caudal to the posterior commissure and storage mechanism (366). Lesions in the midbrain produce ventral to the superior colliculus, showed a selective defect a change in the phase of the VOR (367), and electric stimula- of down-gaze, with tonic upward deviation of the eyes (368). SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 933

Figure 19.29. Locations of lesions producing isolated paralysis of down- ward gaze. Left, Sagittal section through the human brain stem on which the bilaterally destroyed areas in the six autopsied cases of isolated downward gaze paralysis are super- imposed. Right, An enlargement of the lesions shows that the common area destroyed in these cases (stippled region) lies dorsal to the red nucleus at the level of the tractus retroflexus (TR)—that is, around the rostral in- terstitial nucleus of the medial longi- tudinal fasciculus (riMLF) and the nu- cleus of Darkschewitsch (nD). iC, interstitial nucleus of Cajal; IC, infe- rior colliculus; MT, mammillotha- lamic tract; PC, posterior commis- sure; RN, red nucleus; SC, superior colliculus; T, thalamus; III, oculomo- tor nerve; nIII, oculomotor nucleus. (Redrawn from Bu¨ttner-Ennever JA, Bu¨ttner U, Cohen B, et al. Vertical gaze paralysis and the rostral intersti- tial nucleus of the medial longitudinal fasciculus. Brain 1982;105Ϻ125– 149.)

Figure 19.30. Smallest area critical for the selective mediation of downgaze (stip- pled area). Left, Sagittal view of the brain stem as seen from the midline. Right, Coro- nal view of the critical region as seen from a rostral approach. The involved region is rostral to the posterior commissure (PC) in a region that has been designated as the ros- tral interstitial nucleus of the medial longi- tudinal fasciculus (riMLF). iC, interstitial nucleus of Cajal; MLF, medial longitudinal fasciculus; MT, mammillothalamic tract; nD, nucleus of Darkschewitsch; nIII, oculo- motor nucleus; RN, red nucleus; SC, supe- rior colliculus; T, thalamus; TR, tractus re- troflexus; V, third ventricle. (Redrawn from Pierrot-Deseilligny C, Chain F, Gray F, et al. Parinaud’s syndrome: electro-oculo- graphic analysis of six cases with deduc- tions about vertical gaze organization in the premotor structures. Brain 1982;105Ϻ667– 696.) 934 CLINICAL NEURO-OPHTHALMOLOGY

Figure 19.31. Locations of brain stem lesions that produce paralysis of both upward and downward gaze. Left, Five unilateral or bilateral le- sions producing upward and down- ward gaze paralysis are superimposed on a sagittal section of the human brain. The enlargement on the right shows the common area involved (stippled region). This area includes parts of the rostral interstitial nucleus of the medial longitudinal fasciculus (riMLF), the interstitial nucleus of Cajal (iC), and the nucleus of Darkschewitsch (nD). IC, inferior colliculus; MT, mammillothalamic tract; PC, posterior commissure; RN, red nucleus; SC, superior colliculus; T, thalamus; TR, tractus retroflexus; III, oculomotor nerve; nIII, oculomo- tor nucleus. (Redrawn from Bu¨ttner- Ennever JA, Bu¨ttner U, Cohen B, et al. Vertical gaze paralysis and the ros- tral interstitial nucleus of the medial longitudinal fasciculus. Brain 1982; 105Ϻ125–149.)

Another reported lesion of the periaqueductal gray matter of both elevator muscles (the superior rectus and inferior caused persistent up-gaze palsy (369). Thus, it seems that oblique muscles) of one eye. This double elevator palsy is lesions of the periaqueductal gray matter of the midbrain thought to be supranuclear in origin, because in the primary may cause an imbalance of the vertical gaze-holding mecha- position, the eyes are nearly straight, and only on looking nism. Areas of the mesencephalic reticular formation outside upward does a vertical disconjugacy become evident (376). of the riMLF and INC have also been implicated in control of This condition occurs in patients with midbrain infarction saccades. Lesions there affect horizontal and vertical saccade and tumor; the lesion may be ipsilateral or contralateral accuracy; the pattern depends on the exact location within (353,377,378). Rarely, it is congenital (379,380). Because the mesencephalic reticular formation (370,371). the superior rectus is a stronger elevator than the inferior oblique, it is not always possible to be sure of inferior oblique Other Sites and Manifestations weakness. If, however, the inferior oblique is weak in addi- Unilateral, paramedian midbrain lesions sometimes cause tion to the superior rectus, then a nuclear lesion is unlikely, impairment of ipsilateral, horizontal smooth pursuit by af- because these two muscles are supplied by the ipsilateral fecting the descending smooth-pursuit pathway; contralat- and contralateral oculomotor subnuclei, respectively. More eral saccades may also be affected (372), but the horizontal likely, prenuclear inputs to the oculomotor nuclei are im- VOR tends to be spared (Roth-Bielschowsky phenomenon) paired in such patients. The VOR is variably affected. Mono- (313). Paramedian midbrain lesions often damage the oculo- cular elevator palsy may be associated with a contralateral motor nerve nucleus, thereby producing a combination of down-gaze palsy (381). Occasionally monocular elevator nuclear and prenuclear deficits (372,373). Lesions that dam- palsy is caused by a lesion selectively involving the inferior age the oculomotor nucleus are characterized by bilateral oblique and superior rectus muscle fascicles of the oculomo- elevator and eyelid weakness, because the superior rectus tor nerve as it exits the brain stem (382). In most instances, subnucleus is located contralaterally, and the levator sub- however, a monocular paresis of elevation results from nucleus is located in the midline. Abnormal pupils are com- causes other than midbrain disease, such as thyroid ophthal- mon in patients with such lesions. Occasionally, large mid- mopathy, blowout fracture of the , myasthenia gravis, brain lesions also cause complete loss of horizontal eye or restrictive ophthalmopathies. MR scanning of the orbit movements (374,375). may be helpful in differentiating restrictive eye muscle dis- Rarely, midbrain lesions selectively impair the function eases causing a similar clinical picture (383). SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 935

Figure 19.32. Supranuclear vertical ophthal- moplegia in a patient with a metabolic storage dis- ease (DAF syndrome). Top, In primary position, the patient’s eyes are straight. She is also able to make normal voluntary horizontal eye movements. Center, The patient cannot make voluntary upward or downward vertical eye movements. Bottom, On oculocephalic testing, full vertical ocular excur- sions are elicited. (From von Noorden GK, Mau- menee AE. Atlas of Strabismus, 2nd ed. St Louis: CV Mosby, 1973Ϻ161.)

NEUROLOGIC DISORDERS THAT PRIMARILY of PSP occur and appear to be inherited in an autosomal- AFFECT THE MESENCEPHALON dominant fashion (392,393), the disease is usually sporadic. The initial ocular motor deficit of PSP is usually impair- Two neurologic disorders produce profound disturbances ment of vertical saccades and quick phases, either down or of ocular motility, primarily because of their effects on cells up (245). The saccades are at first slow and later also small, in the mesencephalon: progressive supranuclear palsy and with eventual complete loss of voluntary vertical refixations. Whipple’s disease. Vertical smooth pursuit is usually relatively preserved, and the VOR is intact until later in the disease, although a charac- Progressive Supranuclear Palsy teristic nuchal rigidity may make the vertical doll’s head maneuver difficult to perform. A large-field visual stimulus Progressive supranuclear palsy (PSP) is a degenerative often elicits much better visual tracking than does a small disease of later life characterized by disturbances of tone target (394). Bell’s phenomenon is usually absent. There and posture leading to falls, difficulties with swallowing and are many types of eyelid abnormalities in PSP, including speech, and mental slowing (384–389). PSP is one of the blepharospasm, apraxia of eyelid closing or opening, lid re- tauopathies (385,390). Abnormal eye movements are usually traction, and lid lag (395). present early in the course of the disease, but occasionally Horizontal eye movements also show characteristic ab- they are noted late or not at all (391). Although familial cases normalities in patients with PSP, including impaired fixation 936 CLINICAL NEURO-OPHTHALMOLOGY with square-wave jerks, impaired pursuit, impaired vestibu- The differential diagnosis of PSP includes a similar syn- lar cancellation, and saccades and quick phases that are small drome caused by multiple infarcts that affect the basal gan- and eventually slow (396–398). In some patients, the abnor- glia, , and midbrain (418–420). Hydrocepha- mality of voluntary horizontal eye movements resembles an lus may also mimic PSP (421), as can Whipple’s disease INO; however, vestibular stimulation may overcome the lim- (422) and a delayed syndrome following aortic valve re- itation of adduction in such cases (399). Convergence eye placement (423). At the bedside, slowing of vertical sac- movements are commonly impaired in patients with PSP, cades, the inability to suppress blinks to a bright light (a and many patients have severe convergence insufficiency visual ‘‘glabellar sign’’), the ‘‘applause sign’’ (persistent (400). Late in the disease, the ocular motor deficit may clapping after instructed to clap just three times), and tonic progress to a complete ophthalmoplegia. deviation of the eyes and head following sustained rotation Studies of eye movements in patients with PSP may also of the body in a vestibular chair are helpful diagnostic find- reflect the disturbances of attention that occur in this disor- ings (424). der. The latency (reaction time) of horizontal saccades in PSP should be differentiated from cortical-basal gangli- PSP is prolonged in some patients, but in others, saccadic onic degeneration, which may affect vertical and horizontal latency (using a gap paradigm in which the fixation target gaze but which also causes focal , ideomotor goes out before the new peripheral target appears) is reduced apraxia, alien hand syndrome, myoclonus, and an asymmet- so that patients show short-latency or express saccades ric akinetic-rigid syndrome with late onset of gait or balance (401–403). Patients with PSP also make errors with an antis- disturbances (402,403,425–429). A syndrome similar to PSP accade task in which they are required to look away from a has been reported from Guadeloupe and has been attributed presented target. Both the express saccades and errors on to exposure to mitochondrial complex 1 inhibitors such as the antisaccade task suggest defects in function quinolones, acetogenins, and rotenoids (430). It also is a tauopathy. The Parkinson-dementia complex of Guam (Lyt- (401) and, although neuropathologic changes in the frontal ico-Bodig variant) also resembles PSP (431,432). PSP lobe are mild in patients with PSP, positron emission tomo- should also be differentiated from Parkinson’s disease (dis- graphic (PET) scanning indicates profound frontal hypome- cussed later). Although upward gaze may be limited in PSP tabolism in such patients (404). Defects of visual attention and in Parkinson’s disease, impaired downward gaze, slow are more prominent in the vertical than in the horizontal vertical saccades, abnormal horizontal saccades, and square- plane (405), and poor performance on visual search tasks wave jerks are much more characteristic of PSP. Diffuse may be related in part to loss of voluntary saccades and Lewy body disease may present a problem in differential disruption of steady fixation by square-wave jerks (406). diagnosis and mimic PSP (433) or Parkinson’s disease (434). PSP is a diffuse brain stem disorder, although cortical Other basal ganglia disorders that may mimic PSP include involvement also occurs (407). Both computed tomographic idiopathic striopallidodentate calcification (435), autosomal- (CT) scanning and MR imaging show atrophy of the mid- dominant parkinsonism and dementia with pallidopontoni- brain and dilation of the quadrigeminal cisterns, cerebral gral degeneration (436), multisystem atrophy, and olivopon- aqueduct, and third and fourth ventricles (408). There is a tocerebellar atrophy (398). Some of the neurologic deficits characteristic ‘‘hummingbird sign’’ due to the midbrain atro- of PSP, but generally not the eye movements, have been phy (409). Histologically, neuronal loss, neurofibrillary tan- reported to improve modestly with methysergide therapy gles, and gliosis principally affect the brain stem reticular (437), tricyclic antidepressants (438), or formation and the ocular motor nuclei (384,410). The mid- such as (388,439,440). brain may bear the brunt of the early pathology, accounting for the relative vulnerability of vertical saccades (411). Thus, Whipple’s Disease the slowing of vertical saccades probably results from dys- function of burst neurons in the riMLF, whereas the neck Whipple’s disease is a rare multisystem disorder charac- stiffness, often with dorsiflexion, may be caused by damage terized by weight loss, diarrhea, arthritis, lymphadenopathy, to the INC, which contributes to the control of head move- and fever that may involve and even be confined to the cen- ments (412). Abnormalities are also present in the nucleus tral nervous system (441,442). This disease can cause a de- raphe interpositus in the pons in which omnipause neurons, fect of ocular motility that may mimic PSP (443). Initially, important for control of velocity and amplitude of both hori- vertical saccades and quick phases are abnormal (444); zontal and vertical saccades, are located (413). Recent stud- eventually, all eye movements may be lost (445). A highly ies suggest that PSP likely causes the abnormal saccade ve- characteristic finding is pendular vergence oscillations and locity by virtue of its direct effect on premotor saccade burst concurrent contractions of the masticatory muscles: oculo- neurons within the midbrain (414,415). Pathology within masticatory myorhythmia (446–448). The pendular ver- the substantia nigra pars reticulata may also play a role gence oscillations are always associated with vertical sac- (416). Pursuit abnormalities may reflect damage to the cade palsy. Rhythmic palatal tremor may also be present deep pontine nuclei (417). The pedunculopontine nucleus (449). Rajput and Mchattie (450) described ophthalmoplegia is also specifically affected and may be associated with with myorhythmia of the leg, but not of the eyes or jaw. secondary damage to a number of other cortical and sub- Whipple’s disease can be diagnosed using molecular analy- cortical structures (407). sis and can be treated with antibiotics (442). SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 937

OCULAR MOTOR SYNDROMES CAUSED BY LESIONS OF THE SUPERIOR COLLICULUS Lesions restricted to the superior colliculi are rare in hu- normal latency but were hypometric. These findings are sim- mans. One patient (297) underwent removal of a cavernous ilar to those seen after experimental ablation of the superior angioma from the right superior colliculus. Following sur- colliculi in monkeys (451). Pierrot-Deseilligny et al. (452) gery, he had persistently impaired upward gaze, implying reported a patient who had a hematoma largely restricted to pretectal damage, but a full range of horizontal eye move- the right superior colliculus. The patient showed defects in ments. Systematic testing of horizontal saccades demon- latency and accuracy for contralateral saccades and in- strated a paucity of spontaneous refixations contralateral to creased numbers of inappropriate saccades during the anti- the side of the lesion. Saccades to the left occurred after a saccade task.

OCULAR MOTOR SYNDROMES CAUSED BY LESIONS OF THE THALAMUS Thalamic lesions are characterized by disturbances of both ward deviation of the eyes may represent either an irritant horizontal and vertical gaze (453). Conjugate deviation of effect of the hemorrhage on structures responsible for down- the eyes contralateral to the side of the lesion (also called ward gaze or an imbalance created by an acute up-gaze palsy. wrong-way deviation) may occur with hemorrhage affecting Resolution of the downward deviation occurs after treatment the medial thalamus (Fig. 19.33) (18,454). The reason for of raised intracranial pressure; thus, traction on mesence- this contraversive deviation is unclear. The descending path- phalic structures or hydrocephalus may cause this condition. ways from the frontal eye fields to the pons have not yet Esotropia, sometimes quite marked, occurs in patients crossed at this level, although the notion of a precise ocular with caudal thalamic lesions. Although it is usually associ- motor decussation is less certain now than in the past. Dam- ated with downward gaze deviation, it may occur as an iso- age to the descending pathway for smooth pursuit might lated finding (458,459). The esotropia that occurs with thala- lead to a paretic, contraversive deviation of the eyes (305). mic lesions may reflect a disturbance of inputs to premotor Nevertheless, a patient with a defect of smooth pursuit di- vergence neurons in the midbrain (328). Combined lesions rected toward the side of a small hemorrhage in the posterior of the thalamus and midbrain may cause paresis of conver- thalamus and adjacent internal capsule still showed an ipsi- gence (460). versive gaze preference (455). Another possibility is that Patients with posterolateral thalamic infarctions may have wrong-way deviation may be an irritative phenomenon, be- disturbances of the subjective visual vertical (either ipsi- cause neurophysiologic studies show that saccade-related lateral or contralateral) (72). The ocular tilt reaction is not neurons in the intralaminar thalamic nuclei mainly discharge present, however, unless the rostral midbrain is also dam- for contralateral saccades. aged. Forced downward deviation of the eyes, with convergence Infarction of the caudal thalamus caused by occlusion of and , is another common feature of thalamic hemor- the proximal portion of the posterior cerebral artery or its rhage; affected patients appear to peer at their noses perforator branch, the posterior thalamosubthalamic par- (456,547). In autopsied cases, the hemorrhage usually ex- amedian artery, is reported to cause paralysis of downgaze. tends into or compresses the midbrain. Hence, forced down- In fact, this deficit is probably caused not by damage to

Figure 19.33. Contralateral (‘‘wrong-side’’) gaze de- viation with supratentorial, thalamic-basal ganglia hem- orrhage. A, top, The patient’s eyes are deviated to the right. Bottom, Coronal section immediately posterior to the mammillary bodies shows a left intracerebral hem- orrhage involving the thalamus, internal capsule, and basal ganglia. B, top, In another patient, the eyes are deviated down and left. Bottom, Horizontal section through the mid-diencephalon reveals a right intracere- bral hemorrhage involving the pretectum, thalamus, posterior limb of the internal capsule, , and . In both patients, the hemorrhage also in- volved the lateral midbrain tegmentum. (From Keane JR. Contralateral gaze deviation with supratentorial hemorrhage: three pathologically verified cases. Arch Neurol 1975;32Ϻ119–122.) 938 CLINICAL NEURO-OPHTHALMOLOGY the thalamus but by damage to the adjacent riMLF or its grammed using extraretinal information about eye position immediate premotor inputs (461,462). Some patients have must pass through the thalamus and possibly the internal impairment of horizontal gaze, perhaps from interruption of medullary lamina (471). Lesions in the posterior parietal descending pathways (463,464). Associated disturbances of cortex and in the supplementary eye field in the frontal lobes arousal and short-term memory occur in some patients with create similar deficits in the double-step paradigm. Patients thalamic lesions and may be caused by damage to specific with thalamic lesions also show defects in saccade adapta- thalamic nuclei (465,466). Experimentally, combined le- tion, likely reflecting interruption in pathways that carry in- formation to and from the cerebellum (472). sions of the superior colliculus and caudal thalamus in mon- Patients with lesions of the pulvinar develop difficulties keys produce an enduring saccadic hypometria without cor- in shifting attention and gaze into the contralateral hemifield, rective saccades (467), implying a role for these structures manifested by a paucity and prolonged latency of visually in generating efference copy signals (468–470). guided saccades (473,474). These results are consistent with Patients with central thalamic lesions show defects in dou- some reported effects of pharmacologic and destructive le- ble-step saccade paradigms (making successive accurate sac- sions of the pulvinar in the monkey and indicate the impor- cades after a target has quickly jumped twice to different tance of this thalamic nucleus in directing visual attention locations), suggesting that information for saccades pro- (475,476).

OCULAR MOTOR ABNORMALITIES AND DISEASES OF THE BASAL GANGLIA PARKINSON’S DISEASE disease, implicating the basal ganglia in ocular motor learn- ing (506). Patients with Parkinson’s disease may show a number of Pallidotomy causes no improvement in ocular motor per- ocular motor findings (477). Steady fixation is often dis- formance and may actually cause mild deficits in internally rupted by square-wave jerks (478). Upward gaze is often generated saccades and square-wave jerks (507,508). In con- moderately restricted (479), although this abnormality fre- quently is observed in normal, elderly persons (480). Con- trast, stimulation of the subthalamic nucleus has been re- vergence insufficiency is a common and often symptomatic ported to improve saccade performance in Parkinson’s pa- disturbance (481,481a). tients (403). Saccades made reflexively to novel visual stimuli in Par- Clinically, the saccadic initiation defect to command or kinson’s disease are usually of relatively normal amplitude to continuously visible targets appears to be more marked (426,428,482). In contrast, saccades in Parkinson’s disease in the vertical plane. Upward saccades especially may be become hypometric in more complicated tasks such as when hypometric. In contrast, vertical saccades to randomly ap- patients are asked to perform rapid, self-paced refixations pearing visual targets are normal (496). If downward sac- between two stationary targets (478,482–486). During such cades are abnormal, or the velocity of vertical saccades in a test, which relies on an internally generated, predictive either direction is decreased, PSP is a more likely diagnosis strategy, intersaccadic intervals increase above values dur- than Parkinson’s. ing unpredictable tracking to suddenly appearing targets Combined movements of the eye and head may be abnor- (482,483). This effect is not, however, related simply to the mal in Parkinson’s disease. Affected patients tend not to persistence of the target light, because saccades made in move their heads unless instructed to do so (509–511). Dur- anticipation of the appearance of a target light or to a remem- ing rapid eye-head gaze shifts, in response to either predict- bered target location are also hypometric (482,487–491), able or unpredictable step displacements of the target, pa- and patients with Parkinson’s disease have difficulty in gen- tients show increased latency and slowing of head or eye erating sequences of memory-guided saccades to all types movements (510). of stimuli (492,493). Despite this hypometria, patients can Smooth-pursuit movements are usually impaired, al- still shift their gaze, using a series of saccades, to the location though minimally so in mildly affected patients (497). Dur- of a target that is briefly flashed, indicating a retained ability ing tracking of a target moving in a predictable, sinusoidal to encode the location of objects in extrapersonal space pattern, pursuit gain (eye velocity/target velocity) is de- (478,494). Saccadic latencies during nonpredictive tracking creased, leading to catch-up saccades (512). Because the may be normal or mildly increased (478,485,495,496). Pa- catch-up saccades are hypometric, however, the cumulative tients with Parkinson’s disease often show defects in gener- tracking eye movement is less than that of the target, and ating predictive saccades (495,497). Saccadic velocity may this may be one mechanism for defective smooth tracking be normal or, particularly in advanced cases, mildly reduced (497,513). Despite the impairment of smooth-pursuit gain, (478,498). Patients with mild disease perform normally on the phase relationship between eye and target movement the antisaccade task (499,500), but with advanced disease, during tracking of a periodic target is normal (495); this errors increase, especially when patients are also taking anti- implies a normal predictive smooth tracking strategy (484). cholinergic drugs (501,502) Patients with Parkinson’s dis- Cancellation of the VOR, by fixing upon an object moving ease have also been shown to be impaired in saccade search with the head, is abnormal, however, so that during com- strategies that require use of ‘‘working memory’’ (503,504) bined, active eye-head tracking, gain (gaze velocity/target although this may depend upon the difficulty of the task velocity) is reduced (510). (505). Saccade adaptation may be affected in Parkinson’s Both caloric and low-frequency rotational vestibular re- SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 939 sponses, in darkness, may be hypoactive in patients with to an obsessive fixation of a thought. The eyes typically Parkinson’s disease (514,515). At higher frequencies of head deviate upward and sometimes laterally (Fig. 19.34); they rotation, however, and particularly during visual fixation, rarely deviate downward. During the period of upward de- the gain of the VOR is close to 1.0, which accounts for the viation, the movements of the eyes in the upper field of gaze lack of complaint of oscillopsia in such patients (515). appear nearly normal. Affected patients have great difficulty Patients with the syndrome of amyotrophic lateral scle- looking down, except when they combine a blink and down- rosis plus parkinsonism plus dementia (Lytico-Bodig syn- ward saccade. Thus, the ocular disorder may reflect an im- drome), who live on the islands of the South Pacific Ocean, balance of the vertical gaze-holding mechanism. Anticholin- including Guam, may show more severe deficits than pa- ergic drugs promptly terminate both the thought disorder tients with idiopathic Parkinson’s, including limitation of and the ocular deviation, a finding that suggests that the vertical gaze (432). Patients with diffuse Lewy body disease, disorders of thought and eye movements are linked by a Creutzfeldt-Jakob disease, multisystem atrophy (516), corti- pharmacologic imbalance common to both (521). Ocu- cal-basal ganglion degeneration, Guadeloupean parkinson- logyric crises are distinct from the brief upward ocular devia- ism, and PSP must also be distinguished from those with tions that occur in Tourette’s syndrome (discussed later), in idiopathic Parkinson’s (402,424,426,428,430,433,517,518). Rett’s syndrome (523), and in most patients with tardive Antisaccade testing is abnormal in cortical-basal ganglion (524). In some patients with tardive , degeneration (519) and in PSP (discussed previously). MR however, the upward eye deviations are longer-lasting and imaging may help in the differential diagnosis between Par- also are associated with the characteristic neuropsychologi- kinson’s and similar disorders (408,520). cal features of oculogyric crises (525), thus making the dif- Oculogyric crises, once encountered mainly in patients ferentiation between the two entities difficult. Delayed ocu- with postencephalitic parkinsonism, now occur primarily as logyric crises occurred in one patient after a focal a side effect of drugs, especially neuroleptic agents (521), striatocapsular infarction (526) and in another with bilateral although they may occur in other conditions such as reversi- lentiform lesions (527). Episodic brief spells of tonic up- ble posterior leukoencephalopathy (522). A typical attack gaze occurred after bilateral lentiform lesions in a patient begins with feelings of fear or depression, which give rise (528).

Figure 19.34. in patients with postencephalitic parkinsonism. A, In a young man. Note hyperextension of neck, opening of the mouth, and conjugate deviation of the eyes up and to the right. B, In a middle-aged man. Again, the eyes are conjugately deviated up and to the right. (From Kyrieleis W. Die Augenvera¨nderungen bei entzu¨ndlichen Erkrankungen des Zentralnervensystems. III. Die nichteitrigen entzu¨ndlichen Erkrankungen des Zentralnervensystems. A. Die nichteitrige epidemische Encephalitis [Encephalitis epidemica, lethargica]. In: Schieck F, Bru¨ckner A, eds. Kurzes Handbuch der Ophthalmo- logie, vol 6. Berlin: Julius Springer, 1931Ϻ712–738.) 940 CLINICAL NEURO-OPHTHALMOLOGY

In general, treatment of Parkinson’s disease with L-dopa holding and the VOR are preserved, though there may be a does not seem to improve the ocular motor deficits (479), deficit in VOR adaptive capability (550). Late in the disease, except for saccadic accuracy (i.e., saccades become larger) rotational stimulation causes the eyes to tonically deviate (498,529). Occasionally, reversal of saccadic slowing occurs with few or no quick phases. with treatment (530), and newly diagnosed patients with id- Fixation is abnormal in some patients with Huntington’s iopathic disease may experience improved smooth pursuit disease because of saccadic intrusions (541,551). This defect after institution of dopaminergic therapy (529). In patients of steady fixation is particularly evident when the patient with parkinsonism caused by methyl-4-phenyl-1,2,3,6-tetra- views a textured background (89). The paradoxical findings hydropyridine (MPTP) , saccadic latency is short- of difficulty in initiating voluntary saccades but with an ex- ened and saccadic accuracy improved by dopaminergic cess of extraneous saccades during attempted fixation can agents (531). In addition, reflex blepharospasm is improved be further elucidated using novel test stimuli. Such studies in these patients with treatment. These findings are similar reveal an excessive distractibility in, for example, an antisac- to those in monkeys. When these animals are given MPTP, cade task (542). A second finding is that saccades to visual they develop increased latency and duration of saccades, stimuli are made at normal latency, whereas those made to decreased rate of spontaneous saccades, and inappropriate command are delayed. These findings may be related to the saccades, all of which are reversed by dopaminergic therapy parallel pathways that control the various types of saccadic (532,533). In patients with idiopathic disease who show pro- responses. On the one hand, disease affecting either the fron- nounced drug-related fluctuations, there is disagreement as tal lobes or the , which inhibits the pars reti- to whether smooth pursuit shows increase in gain during culata of the substantia nigra, may lead to difficulties in ‘‘on’’ periods (498,534,535). Although patients who are re- initiating voluntary saccades in tasks that require learned or ceiving L-dopa and who show dyskinetic movements of other predictive behavior (537,538). On the other hand, Hunting- body parts (536) are said to have unsteady gaze, particularly ton’s also affects the pars reticulata of the substantia nigra in darkness, this phenomenon usually is not evident clini- (552). Because the pars reticulata of the substantia nigra cally. inhibits the superior colliculus and therefore suppresses re- The dopaminergic pars compacta of the substantia nigra flexive saccades to visual stimuli, one might expect exces- does not appear to contain neurons related to eye movement, sive distractibility during attempted fixation. The slowing whereas the pars reticulata does (537,538). In monkeys with of saccades may reflect damage to saccadic burst neurons MPTP-induced parkinsonism, the cerebral metabolic rate (553), but at least some pathologic evidence (554) suggests was reduced in the frontal eye field and paralamellar medio- that disturbance of prenuclear inputs, such as the superior dorsal thalamus (539). It is possible that these metabolic colliculus or frontal eye field, is responsible. changes are related to loss of projections from the dopamine- Despite the nearly ubiquitous finding of abnormal eye depleted substantia nigra. movements in patients with Huntington’s, some persons with the disease who are evaluated before they become HUNTINGTON’S DISEASE symptomatic show normal eye movements (89,555). Thus, routine testing of eye movements is not a reliable method Huntington’s disease produces disturbances of voluntary for determining which offspring of affected patients will go gaze, particularly saccades (540–543). It is caused by a ge- on to develop the disease. Reveley et al. (556) reported some netic defect of the IT15 gene (the ‘‘Huntington’’ gene) on improvement in the eye movement abnormalities in patients chromosome 4, producing a CAG triplet repeat (544). Initia- with Huntington’s who were treated with sulpiride. tion of saccades in patients with Huntington’s disease may Neuroacanthocytosis (557) must be distinguished from be difficult. Such patients show prolonged latencies, espe- Huntington’s disease (and PSP), but eye movement abnor- cially when the saccade is to be made on command, in antici- malities do not seem to be an important feature of this disor- pation of a target moving in a predictable fashion, or in a der. Dentatorubropallidoluysian atrophy (558), also called self-paced fashion. An obligatory blink or head turn may be the Haw River syndrome (559), is another CAG triplet repeat used to start the eye moving (545). Saccades may be slow disease (B37, chromosome 12) that must be distinguished in the horizontal or vertical plane; this deficit can often be from Huntington’s, although myoclonus and ataxia are more detected early in the disease if eye movements are measured common in dentatorubropallidoluysian atrophy than in Hun- (89), but it may not be evident clinically until late in the tington’s. course (541). Saccades may be slower in patients who be- come symptomatic at an earlier age, and such persons are OTHER DISEASES OF BASAL GANGLIA more likely to have inherited the disease from their father (546). The slowing of vertical saccades in Huntington’s dis- A number of conditions other than Parkinson’s and Hun- ease probably does not occur in patients with chorea from tington’s that affect the basal ganglia cause abnormal eye nondegenerative conditions or tardive dyskinesia (547). movements. Hepatolenticular degeneration (Wilson’s dis- Longitudinal studies of saccades can be used to quantify the ease) and juvenile dystonic lipidosis (Niemann-Pick type progression of the disease (540,548). 2s) are discussed below. Other conditions include caudate Smooth pursuit may also be impaired with decreased gain hemorrhage, which has been associated with ipsilateral gaze in patients with Huntington’s disease, but it often is rela- preference (560) and, rarely, Sydenham’s chorea. Vermersch tively spared compared with saccades. By contrast, gaze- et al. (561) recorded saccadic eye movements in patients SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 941 with bilateral lesions in the lentiform nucleus and found tasks, though there is considerable disagreement among prominent abnormalities in saccades that required an internal these studies as to the exact deficits (568–573). A confound- representation of the target (remembered saccades, saccades ing factor in these studies is the presence or absence of asso- to sequences, predictive saccades) but normal responses for ciated attention-deficit/hyperactivity disorder (ADHD) visually guided saccades, including antisaccades (which are (572,574,575). nonetheless triggered by a visual target). Lekwuwa et al. Routine ocular motor functions are normal in most pa- (562) suggested that defects in control of predictive pursuit tients with essential blepharospasm (576). In some studies, eye movements are a feature of striatal damage. however, abnormal saccade latencies are seen for both vis- Patients with Tourette’s syndrome may show a variety of ually guided and memory-guided saccades (577,578), sug- ocular abnormalities, including blepharospasm and eye tics gesting a possible basal ganglia localization. Patients with that include involuntary gaze deviations (563,564), but sac- spasmodic torticollis may show abnormalities in vestibular cades, fixation, and pursuit, when tested in conventional lab- function, including the torsional VOR (88,579). Patients oratory testing paradigms, show no abnormalities (565). The with tardive dyskinesia may show increased saccade distrac- gaze deviations must be distinguished from benign eye tibility (580). Severely affected patients with Lesch-Nyhan movement tics, which children often outgrow (566,567). syndrome (hypoxanthine-guanine phosphoribosyltransfer- Quantitative testing of patients with Tourette’s syndrome ase [HPRT] deficiency) show excessive distractibility, diffi- have shown various saccade abnormalities in ‘‘higher- culty initiating saccades, and an oculomotor apraxia-like level’’ tasks such as the antisaccade task and ‘‘go/no-go’’ condition (581).

OCULAR MOTOR SYNDROMES CAUSED BY LESIONS IN THE CEREBRAL HEMISPHERES Extensive reviews of the effects of cerebral hemisphere lesions on eye movements were done by Pierrot-Deseilligny (582) and Leigh and Kennard (583). One particular issue that must be considered in assessing abnormalities of eye movements in patients with lesions in the cerebral hemi- spheres is the sometimes confounding role of defects in di- recting visual attention and the associated neglect (584–587).

ACUTE LESIONS Following an acute lesion of one cerebral hemisphere, the eyes usually deviate conjugately toward the side of the lesion (Prevost’s or Vulpian’s sign) (588) (Fig. 19.35). Gaze devia- tions are more common after large involving pre- dominantly the right postrolandic cortex (589,590) (Fig. 19.36). Visual hemineglect often accompanies such gaze de- viations (590,591). Gaze deviation that occurs after a usually resolves within a few days to a week; if gaze devia- tion persists, there may be a previous lesion in the contra- lateral frontal lobe (592). In general, for comparably sized lesions, ocular motor defects—both pursuit and saccades—are more profound when the lesion is in the non- dominant hemisphere (593). In the acute phase, patients may not voluntarily direct their eyes toward the side of the intact hemisphere, in part because of neglect. Shortly thereafter, however, vestibular stimulation usually produces a full range of horizontal move- ment (with the slow phase), in contrast to most gaze palsies associated with pontine lesions (313). When quick phases of caloric nystagmus directed away from the side of the Figure 19.35. Persistent hemispheral (supranuclear) horizontal gaze palsy lesion are absent, consciousness is usually impaired. Some- following severe frontal head trauma. Five weeks before this photograph was taken, the patient sustained a severe blow to the head, causing a de- times, in addition to ipsiversive deviation of the eyes, there pressed right frontal skull fracture. Operation disclosed extensive contusion is a small-amplitude nystagmus with ipsilateral quick and necrosis of the right frontal lobe (arrows indicate the site of the surgical phases. A similar finding occurs acutely after hemidecortica- incision). The patient had and was hemiplegic on the right tion in the monkey (594). The slow phases of this nystagmus side. His eyes remained in right conjugate gaze at all times, but he could may reflect unopposed pursuit drives directed away from follow a very slowly moving target to the left. 942 CLINICAL NEURO-OPHTHALMOLOGY

Figure 19.36. Neuroimaging of an acute infarct in the territory of the left anterior cerebral and middle cerebral arteries. A, T2-weighted axial magnetic resonance image shows diffuse hyperintensity in the distribution of the left anterior cerebral artery (large arrowheads). Note the small area of hyperintensity in the right frontal region, consistent with an old infarct (small arrowhead). B, T2-weighted magnetic resonance image at a lower plane reveals an area of hyperintensity in the region of the left middle cerebral artery (arrowhead). The patient had a supranuclear conjugate gaze paresis to the left side and right hemineglect. He also had difficulty generating rightward saccades. The ocular motor signs were transient, as expected from a left-sided cerebral lesion. the side of the lesion. Vertical saccades may be abnormal; right hemisphere lesions. In the right hemisphere, the lesions they are dysmetric with an inappropriate horizontal compo- are located predominantly in the subcortical frontoparietal nent toward the side of the lesion (595). Because both hemi- region and the internal capsule. In the left hemisphere, the spheres usually must be activated to elicit a purely vertical lesions are usually larger, covering the entire frontotemporo- saccade, the loss of one hemisphere may cause such oblique parietal area. The larger the lesion, the more persistent the saccades. conjugate deviation. Quantitative recordings in such patients Following intracarotid injection of (Wada test show that both the pursuit and the saccade deficits associated to determine cerebral dominance), a transient gaze prefer- with conjugate deviation of the eyes are predominantly con- ence may occur in association with hemiparesis; such a de- tralateral, but as the conjugate deviation resolves, an ipsi- viation occurs more commonly with right-sided injections, lateral pursuit defect becomes more apparent (599). Cranio- providing further evidence for the dominance of the right topic defects in saccades and especially pursuit may outlast hemisphere in directing attention (596). During the period the conjugate deviation, however, being greater in the field of hemiparesis of the Wada test, contralateral and ipsilateral contralateral to the lesion (600). The initial conjugate devia- saccades are still possible, with relatively minor slowing of tion may reflect the effect on eye movements of cerebral the former. This persistence of voluntary saccades is proba- hemisphere asymmetries in attention mechanisms (596), bly related to the influence of posterior cerebral areas, which whereas the more enduring eye movement defects after the receive blood supply from the vertebrobasilar system and conjugate deviation resolves may reflect more specific de- which project to the superior colliculus, independently of fects in motor control mechanisms mediated by the hemi- the frontal eye fields (597). spheres. Conjugate eye deviation is occasionally ‘‘wrong- As noted above, acute cerebral hemisphere lesions are way’’ (i.e., contralateral to the side of the lesion) (601). The frequently associated with a conjugate deviation of the eyes lesions are almost always hemorrhagic, most commonly in (598) (Fig. 19.35). The deviation is usually ipsilateral to the the thalamus. Affected patients usually have signs of rostral side of the affected hemisphere and is more common with brain stem dysfunction and a shift of midline structures. Epi- SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 943 leptic phenomena, impairment of ipsilateral pursuit path- ways, and more caudal damage to descending pathways near the brain stem are evoked as explanations. In most cases, the last cause seems most plausible. Acute hemisphere le- sions may cause epileptic seizures with contralateral devia- tion of the eyes or nystagmus, often with an associated head turn. The ocular motor manifestations of are dis- cussed below. Unilateral or bilateral ptosis may also occur with acute lesions in the cerebral hemispheres, usually in Figure 19.37. Conjugate lateral deviation of the eyes on forced closure of the eyelids in a patient with a tuberculoma of the left occipitoparietal the nondominant hemisphere (311,602). region. (From Cogan DG. Neurologic significance of lateral conjugate de- PERSISTENT DEFICITS CAUSED BY LARGE viation of the eyes on forced closure of the lids. Arch Ophthalmol 1948; 39;37–42.) UNILATERAL LESIONS Persistent ocular motor deficits caused by lesions such as hemidecortication for intractable seizures are summarized in Table 19.7. Although there may be no resting deviation movements may be too fast (pursuit gain more than 1); for of the eyes, forced eyelid closure may cause a contralateral higher target velocities, pursuit gain toward the intact side spastic conjugate eye movement, the mechanism of which is normal (605,606). This disturbance of smooth pursuit is not understood (Fig. 19.37) (603,604). This tonic deviation probably reflects loss of both posterior (occipital-parietal- (Cogan’s sign) differs from the tonic deviation associated temporal) and frontal influences. with Wallenberg’s syndrome, because in the former, active A convenient way to demonstrate the asymmetry of or attempted eyelid closure is necessary to cause the eyes smooth pursuit that occurs with large hemisphere lesions is to deviate, whereas in the latter, the deviation occurs even with a hand-held optokinetic tape or drum (607). The re- with eyes open in darkness. Cogan’s sign occurs most fre- sponse is decreased when the stripes are moved or the drum quently with parietotemporal lesions. is rotated toward the side of the lesion. At the bedside, this In primary position, a small-amplitude nystagmus may optokinetic response is usually judged by the frequency and be present that is best seen during ophthalmoscopy. It is amplitude of quick phases, but because these quick-phase characterized by slow phases directed toward the side of the variables also depend on slow-phase velocity, a decreased intact hemisphere and may represent an imbalance in response may reflect impaired slow-phase generation, im- smooth-pursuit tone (605). Horizontal pursuit gain (eye ve- paired quick-phase generation, or a combination of the two. locity/target velocity) is low for tracking of targets moving Hemidecortication causes abnormalities of both contralat- toward the side of the lesion for all stimulus velocities. For eral and ipsilateral horizontal saccades (605,608). Saccades targets moving slowly toward the intact hemisphere, the eye are usually slower than normal for refixations into and some- times out of the hemianopic field. Saccadic latency is also prolonged in both directions (605). For small refixations, Table 19.7 contralaterally directed saccades have greater latencies than Persistent Effects of Large Unilateral Lesions of the Cerebral ipsilateral saccades. Prolonged saccadic reaction time may Hemispheres on Ocular Motor Function reflect (a) defects in visual detection because of the hemia- nopia, (b) defects in directing visual attention, and (c) abnor- Fixation: In darkness, eyes usually drift away from the side of the lesion. This may also be evident during fixation (on ophthalmoscopic exam) as mal motor programming. Saccadic accuracy is impaired nystagmus with quick phases toward the side of the lesion. Square-wave asymmetrically: most contralaterally directed saccades do jerks. not put the eye on target (608). Saccades: Slower saccades to both sides, especially contralaterally; latency The horizontal VOR may be mildly asymmetric in hemi- longer for small saccades directed contralateral to the side of the lesion; decorticate patients, with the gain (eye velocity/head veloc- inaccurate (hypometric and hypermetric) saccades into the “blind” ity) being greater for compensatory eye movements directed hemifield. away from the side of the lesion (609,610). More asymmetry Smooth pursuit: Reduced pursuit gain toward the side of the lesion; smooth- appears when visual fixation and vestibular stimulation are pursuit gain away from the side of the lesion may be increased for low- velocity targets. combined (during rotation while fixating a stationary ob- Optokinetic: Reduced gain for stimuli directed toward the side of the lesion; ject), probably reflecting the ipsilateral smooth-pursuit defi- impaired optokinetic after-nystagmus; may be relatively preserved cit. The asymmetry is still present during head rotation if compared with pursuit, with prolonged build-up of slow-phase velocity. the patient imagines a stationary object (610). Vestibular: During sinusoidal rotation, VOR gain in darkness may be slightly asymmetric (greater for eye movements away from the side of the lesion); with attempted fixation of an imagined or real stationary target, FOCAL LESIONS the asymmetry is increased. Forced eyelid closure: Eyes usually deviate conjugately away from the side Ocular motor disturbances that occur from focal lesions of the lesion (“spasticity of conjugate gaze”). of the cerebral hemispheres depend on a variety of factors, including the location and size of the lesion and whether the VOR, vestibulo-ocular reflex. lesion is unilateral or bilateral. 944 CLINICAL NEURO-OPHTHALMOLOGY

Occipital Lobe Lesions paired (627). Such a behavior deficit occurs in patients with lesions affecting the temporoparieto-occipital cortex A unilateral lesion of either causes a contra- (625,628–631). lateral visual field defect and an ocular motor deficit (sac- Lesions of adjacent cortex (including Brodmann areas 19 cadic dysmetria) that reflects the patient’s homonymous and 39), which probably correspond to the medial superior hemianopia. Saccades into the hemianopic visual field are temporal (MST) visual area and underlying white matter, dysmetric, usually hypometric, and similar patterns of sac- lead to a directional defect in smooth pursuit that is charac- cades are seen with acoustic targets, implying some degree terized by impaired tracking (reduced gain) for targets mov- of common motor programming, perhaps influenced by as- ing toward the side of the lesion, irrespective of the visual sociated defects in directing spatial attention (611). Charac- hemifield in which the target lies (628,632,633). Barton et teristic patterns also occur in patients who have hemianopic al. (630) did not find a direct correlation between pursuit and dyslexia (612). Patients with hemianopia usually show a va- motion detection deficits. Either could be present without the riety of compensatory strategies to increase saccadic accu- other, although unavoidable differences in the paradigms for racy (613), unless hemineglect is also present (614). These eye movement and for motion perception precluded a strict strategies include (a) a staircase of search saccades with one-to-one correspondence. Lawden et al. (634) showed that backward, glissadic drifts; (b) a deliberate overshooting sac- pursuit is also influenced by the presence of a stationary- cade to bring the target into the intact hemifield of vision; structure background in patients with cerebral lesions. Le- and (c) with predictable targets, saccades using memory of sions in the inferior parietal cortex (area 40) and in white previous attempts. Such findings have been used to develop matter containing frontoparietal connections were associated simple clinical tests for distinguishing homonymous hemia- with a significant degradation in pursuit capability when pa- nopias with and without neglect (615). Rapid gaze shifts tients had to track a small target moving on a structured achieved by combined movements of eye and head also show background rather than a small target moving in darkness. increased latency of head movements and development of In some patients, pursuit gain away from the side of the compensatory strategies when looking to the hemianopic lesion is also somewhat reduced, especially side (616). Smooth pursuit remains intact with unilateral when the eyes move into the contralateral field of gaze. This lesions of the striate cortex, as long as the moving stimulus phenomenon probably results from contralateral neglect is presented to the intact hemifield (617). Optokinetic nys- (600,635) and may be associated with a paucity of explora- tagmus elicited at the bedside is usually symmetric, unless tory eye movements in the contralateral field (636). In other subcortical pathways involved in smooth tracking are also patients, contralateral pursuit gain is increased (628,637). affected (618; discussed later). Subcortical, thalamic, and brain stem lesions may cause an Bilateral occipital lesions cause . A pa- ipsilateral defect from damage to involvement of the de- tient with bilateral, congenital occipital lesions and little re- scending pathway for smooth pursuit (623,638,639). Full- sidual vision was reported to be able to make voluntary sac- field optokinetic stimulation is also impaired with experi- cades but not smooth pursuit (619). Optokinetic responses mental lesions of MST in monkeys (627). In human patients are present in monkeys following bilateral occipital lobec- with parietal lesions, optokinetic nystagmus (OKN) is often tomy (620), but this may not be the case in humans (621). impaired in response to stimuli moving toward the side of Focal occipital seizures may cause either contralateral or the lesion, although it may be relatively spared compared ipsilateral deviation of the eyes and nystagmus (discussed with foveal tracking (624,640). Optokinetic after-nystagmus later). (OKAN) and circularvection (the sensation of self-rotation) Parietal Lobe Lesions may also be impaired (625). In monkeys, acute lesions of the parietal lobe cause asymmetries of the VOR (641), but Unilateral lesions of the parietal lobes, especially those comparable studies in humans are lacking. involving the inferior parietal lobule and underlying deep Unilateral lesions of the parietal lobe may affect saccadic white matter, cause abnormalities of ocular tracking of mov- initiation, causing an increase in saccadic latencies, either ing targets, including an asymmetry of smooth pursuit and bilaterally (642) or only for saccades to contralateral targets of optokinetic nystagmus as tested at the bedside with hand- (643,644). These changes are independent of any visual field held drums or tapes (607,622–625). Lesions at the temporo- defect. The latency defects are enhanced when the fixation parieto-occipital junction probably affect secondary visual target remains on and a new target appears in the periphery areas that are important for motion processing and for pro- (the ‘‘overlap’’ task) and are diminished when the fixation gramming of smooth-pursuit eye movements (626). One target is extinguished before the presentation of the new such area is likely to be the human homologue of what in target in the periphery (the ‘‘gap’’ task). Other deficits re- monkeys is called the middle temporal (MT) visual area. ported with unilateral parietal lesions include inaccuracy and Lesions of MT in monkeys impair the ability to estimate hypometria of contralaterally directed visually guided sac- the speed of a moving target that is within the affected visual cades (645), mild saccadic slowing (646), and disturbances field, although stationary objects can be seen and accurately of predictive saccadic tracking (647). Memory-guided sac- localized. The ocular motor consequences of this cades are especially impaired, being both delayed and inac- for motion or retinotopic defect are that saccades made to curate (648,649). These changes in saccades are more promi- targets moving in the affected, contralateral hemifield are nent with right hemisphere lesions. The saccade defects inaccurate and that the initiation of smooth pursuit is im- associated with parietal lesions may be caused in part by SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 945 difficulties in shifting attention from one position to another feeling of dizziness is common with a variety of seizure in extrapersonal space (650), but there are no strict correla- types, a true sensation of rotation—tornado epilepsy—is tions between attention and the ocular motor deficits a rare but well-described epileptic phenomenon (664, (651–653). In a double-step paradigm in which subjects had 673–679). to be able to use nonretinal, corollary discharge information about the saccade to the first target to compute the size and Frontal Lobe Lesions direction of the second saccade, patients with parietal lesions were not able to make the appropriate computation about Lesions of the frontal lobe, in both monkeys and humans, the size and direction of the saccade (i.e., their spatial com- may produce an ipsilateral conjugate deviation of the eyes putation was impaired, implying a defect in using extrareti- that resolves with time (313,680). Rarely, contralateral de- nal signals of eye position) (645,654,655). In general, pari- viation occurs with acute frontal lesions (305), which may etal lesions have relatively little effect on generation of more include just the frontal eye fields (681) or frontoparietal le- volitional internally guided saccades (i.e., when the target sions (682). Enduring deficits after frontal lobe lesions in- position is memorized or known) or on reflexive saccade clude abnormalities of saccades and smooth pursuit (651). inhibition (as in the antisaccade task) but seem more in- Three areas within the frontal lobes that play an important volved with triggering the superior colliculus directly to ini- role in the control of eye movements are (a) the frontal eye tiate reflexive saccades to the unexpected appearance of fields; (b) the supplementary eye fields in the supplementary novel targets in the visual environment (656). Patients with motor area; and (c) the prefrontal cortex, especially the dor- parietal lesions also show defects in visual search tasks, solateral portion (683). though they may involve both the ipsilateral and contralat- Unilateral frontal eye field lesions lead to a slight increase eral visual fields (587,649,657,658). Selection of targets of in saccade latency to reflexively triggered saccades with a interest among competing visual stimuli is also affected by predominantly contralateral hypometria (684–687). Laten- experimental parietal lesions (659), and this finding may be cies are greatest in the overlap task (when the initial fixation related to parietal patients who show simultagnosia. target remains on, even after the peripheral target appears), Bilateral parietal lobe lesions may cause acquired ocular suggesting a role for the frontal eye fields in disengagement motor apraxia, particularly if the lesions are large (discussed from central fixation. Saccades may show a prolonged la- later). When smooth pursuit is possible, it is particularly tency to predictable target jumps, particularly in patients limited with higher-acceleration target motion (660). with right-sided frontal lesions (684,688). There is also a bilateral deficit in latency and accuracy for saccades to a Lesions remembered visual target but not for remembered saccades after a vestibular (rotation) input (671). With attempted ver- The effects of temporal lobe lesions on eye movements tical saccades, a horizontal component directed toward the are less well studied. In patients with posterior temporal side of the lesion often causes an oblique movement (595). lesions, fixation-suppression of caloric-induced nystagmus Mild slowing of contralateral saccades occurs in some pa- is said to be impaired when slow phases are directed away tients (684). Deep, unilateral frontal lobe lesions cause in- from the side of the lesion (661). This abnormality may creased latency for contralateral saccades (642). This deficit reflect impairment of visual-motion or smooth-pursuit path- is probably caused by damage of efferent and afferent con- ways rather than any effect on vestibular nystagmus per se. nections of the frontal eye fields. Patients with homonymous hemianopia and lesions affecting Patients with unilateral frontal lesions show defects in the temporal lobes may lack the sensation of self-rotation the antisaccade task. This requires the subject to suppress a (circularvection) that normally occurs during full-field op- reflexive glance toward a peripheral visual target and, in- tokinetic stimulation, compared with patients who have a stead, look toward its mirror location in the contralateral homonymous hemianopia from occipital lesions and who do visual field (689). Patients with unilateral frontal lobe lesions experience circularvection (662). These findings support the that involve the dorsolateral portion of the frontal cortex localization of the vestibular cortex to the superior temporal have difficulty in suppressing the reflexive glance, especially gyrus and, perhaps, the adjacent parietal cortex (663–670). if the visual stimulus appears in the visual hemifield contra- Patients with parietoinsular lesions may have tilts of the sub- lateral to the side of the lesion (690). Patients with lesions jective visual vertical, usually contraversive (107). This is in the frontal eye fields alone do not have this problem, not associated with a skew deviation, although occasionally though they do have difficulty in generating the antisaccade there is some monocular torsion. Patients with lesions in the when the target lies in the visual hemifield ipsilateral to the same area may also have a defect in generating memory- lesion (689,691). After lesions in the dorsolateral portion of guided saccades after a vestibular (rotational) stimulus (671). the frontal cortex, there is also a deficit in both vestibular Finally, patients with lesions in the medial temporal (rotation) and visually guided memory saccades (692) as lobe—the hippocampus—show marked impairment of gen- well as deficits in generating predictive saccades (693) and erating sequences of saccades, whereas their spatial memory in visual search (587,694). Taken together, the results of is intact (672). lesion studies suggest that the dorsolateral portion of the Seizures emanating in the temporal lobes may cause a frontal cortex has an overall role in decisional processes variety of vestibular sensations, although lesions in the fron- related to suppressing unwanted saccades and facilitating tal lobes may produce similar phenomenon. Although a mild saccades to known upcoming target locations. Lesion studies 946 CLINICAL NEURO-OPHTHALMOLOGY also suggest that the anterior may also play drome, the lesions are usually bilateral and located in the a role in suppressing unwanted saccades (695). Lesions af- parietal or occipital lobes. The anatomic basis for this dis- fecting the supplementary eye fields, especially in the left turbance is not known, although it has been attributed to hemisphere, impair the ability of patients to make a remem- defects in the inhibitory control of the superior colliculus by bered sequence of saccades (696,697). Likewise, patients the substantia nigra, pars reticulata (712). Spasm of fixation with such lesions cannot make a memory-guided saccade to refers to a rare defect in generating voluntary eye movements a visual target, but they can do so after a body rotation when a fixation target is continuously present; it is alleviated (692,698). Patients with lesions of the supplementary eye when the central fixation target is removed (712). fields may also show abnormalities in changing the direction of saccades from a previously learned sequence (699). OCULAR MOTOR APRAXIA Patients with unilateral frontal lobe lesions also show pur- Ocular motor apraxia is characterized by an impaired abil- suit deficits (562,593,625,687,700,701). The frontal and ity to generate saccades on command (Fig. 19.38). Congeni- supplementary eye fields and perhaps the dorsolateral por- tal ocular motor apraxia was first described by Cogan in tion of the prefrontal cortex play a role in this abnormality. 1965. An abnormality in eye movements may be recognized Defects are in both initiation and maintenance (more so at shortly after birth, when the child does not appear to fixate higher target speeds and frequencies). If lesions are in the upon objects normally and may appear blind (713). Some supplementary eye field, defects are ipsilateral; if lesions are children with congenital ocular motor apraxia are said to in the frontal eye field, defects are bilateral but usually have a transient head and limb tremor in the first few days greater for ipsilateral tracking. Patients with lesions of the of life. Between ages 4 and 6 months, characteristic thrusting supplementary eye field may have delayed reversal with pe- horizontal head movements develop, sometimes with promi- riodic constant-velocity stimuli, implying impaired anticipa- nent blinking or even rubbing of the eyelids, when the child tion of the target trajectory. Saccades to moving targets are attempts to change fixation. In children with poor head con- also inaccurate in some of these patients. Patients with le- trol, development of head thrusting may be delayed or ab- sions affecting the frontal eye fields may also show cranio- sent. Almost all patients also show a defect in generating topic as well as directional (ipsilateral) defects in pursuit quick phases of nystagmus, which can usually be appreciated (600). Tracking in the field contralateral to the lesion is at the bedside by manual spinning of the patient, either when worse than in the ipsilateral field. Visual exploration deficits holding the child out at arm’s length or by rotating the child may also play a role in some of the eye movement deficits on a swivel chair (if necessary sitting in an adult’s lap). seen with frontal lesions (702). Despite difficulties in shifting horizontal gaze, vertical vol- Acute bilateral frontal or frontoparietal lesions may pro- untary eye movements are normal. duce a striking disturbance of ocular motility that is called Measurements of eye and head movements have further acquired ocular motor apraxia. It is characterized by loss characterized congenital ocular motor apraxia (714–716). of voluntary control of eye movements, both saccades and With the head immobilized, patients show both impaired pursuit, with preservation of reflex movements, including initiation (increased latency) and decreased amplitude (hy- the VOR and quick phases of nystagmus. There is also rela- pometria) of voluntary saccades in response to either a sim- tive preservation of saccades made to visual targets, com- ple verbal command to look left or right or in tracking a step pared with internally guided saccades made on command, displacement of a target. Saccades are also delayed during and preservation of saccades made with blinking or with attempted refixations between auditory targets in complete head movements (703,704). Voluntary movements of the darkness. Thus, abnormal visual-fixation reflexes cannot ex- eyes are limited in the horizontal and usually also in the vertical plane. The defect of voluntary eye movements prob- ably reflects disruption of descending pathways from both the frontal eye fields and the parietal cortex so that the supe- rior colliculus and brain stem reticular formation are bereft of their supranuclear inputs. The behavior deficit is similar to that produced by combined, experimental lesions of the frontal eye fields and superior colliculus (705). When a similar disorder of ocular motility, called ‘‘psychic paralysis of gaze,’’ is associated with optic ataxia and disturbance of visual attention (simultanagnosia), the eponym Ba´lint’s syndrome is used (703–706) (see Chapter 13). The main abnormalities are a defect in the visual guid- ance of saccades (increased latency and decreased accuracy) Figure 19.38. Congenital ocular motor apraxia. The patient was looking and inaccurate searching saccades. Voluntary saccades may to the right (left) when he was told to glance at the camera. His head rapidly moved to the left, while the eyes remained deviated toward the right. As be made with greater ease than those in response to visual a result, the patient’s head had to turn further to the left in order to permit targets, the converse of what one usually finds with more fixation ahead (center). Once the patient was able to fixate on the camera, frontal lesions (707–709). Spontaneous blinking may be lost his head turned slowly back to the right until it was straight (right). (From (710). In one patient, the visual scene faded during fixation Urrets-Zavalia A, Remonda C. Congenital ocular motor apraxia. Ophthal- and intentional blinking restored it (711). In Ba´lint’s syn- mologica 1957;134Ϻ157–167.) SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 947 plain the difficulty initiating saccades. Saccadic velocities quick phases (both horizontal and vertical) are typically af- are normal, however, and saccades or quick phases of nys- fected, and because saccades may be slow. In the early stages tagmus of large amplitude can occasionally be generated. of these diseases, however, the ocular motor apraxia may These findings indicate that in these patients, the premotor be indistinguishable from congenital ocular motor apraxia. brain stem burst neurons that generate saccadic eye move- Thus, patients with ataxia-telangiectasia (Louis-Bar syn- ments are intact. In younger patients, however, the timing drome, 11q22-23) and its variants (ataxia-oculomotor and amplitude (but not velocity) of quick phases of vestibular apraxia syndrome of Aicardi, types 1 and 2), Gaucher’s dis- and optokinetic nystagmus may be impaired (i.e., the eyes ease (types 2 and 3,1q21-31), Niemann-Pick disease type intermittently deviate tonically in the direction of the slow 2s, Pelizaeus-Merzbacher disease, Cockayne’s syndrome, phase because of a defect in the initiation of quick phases). Huntington’s disease, hepatolenticular degeneration, vita- Sometimes the saccade defects and head thrusts are asym- min E deficiency, some of the peroxisome disorders, Whip- metric (717). Pursuit eye movements may also be of low ple’s disease, and many other storage diseases and amino- gain, but the corrective saccades are usually promptly gener- acidurias may appear to have ocular motor apraxia ated. The defects in congenital ocular motor apraxia are usu- (723,724,727–735) (Fig. 19.32). ally restricted to the horizontal plane, an important differen- A few features that can be used to distinguish some of tial diagnostic point because most acquired cases of ocular these disorders from congenital ocular motor apraxia include motor apraxia also cause defects in the vertical plane. the following: The head thrusts made by patients with congenital ocular motor apraxia probably reflect one of several adaptive strate- 1. In Pelizaeus-Merzbacher disease, pendular, elliptical gies to facilitate changes in gaze (714,718). Younger patients nystagmus, upbeat nystagmus, saccade dysmetria, and appear to use their intact VOR, which drives their eyes into other cerebellar signs (especially truncal titubation) are an extreme contraversive position in the orbit. As the head often present. continues to move past the target, the eyes are dragged along 2. Patients with Niemann-Pick disease characteristically in space until they become aligned with the target. The head show a vertical and especially downgaze disturbance then rotates backward and the eyes maintain fixation as they along with dysarthria, ataxia, cognitive dysfunction, and are brought back to the primary position in the orbit by the emotional changes. Such patients have so-called foam VOR. In contrast, older patients appear to use the head cells in the bone marrow. Rarely, a horizontal gaze dis- movement alone to trigger the generation of a saccadic eye turbance dominates the picture. The typical vertical pat- movement that cannot normally be made with the head still. tern of gaze disturbance in Niemann-Pick must be dis- This strategy may reflect the use of a phylogenetically old tinguished from the vertical gaze palsies associated with linkage between head and saccadic eye movements that oc- , in which athetosis and hearing loss are com- curs reflexively in afoveate animals such as the rabbit when monly associated. they desire to redirect their center of visual attention. 3. In Gaucher’s disease, vertical gaze may be affected, but The cause of congenital ocular motor apraxia is unknown. horizontal gaze disturbances are more common. Cogan (719) suggested a delay in the normal development 4. Tay-Sachs disease impairs vertical and, subsequently, of the mechanisms by which we assume voluntary control horizontal eye movements. over eye movements. Delayed psychomotor development 5. Late-onset hexosaminidase deficiency also preferen- (especially in learning to read and in speech), infantile hypo- tially affects vertical gaze. tonia, strabismus, incoordination, torsional nystagmus, and 6. In Whipple’s disease, supranuclear and especially verti- clumsiness occur in some patients (720). Associated anoma- cal gaze palsy is frequent and may be associated with lies, especially agenesis of the corpus callosum and cerebel- the convergence oscillations of ‘‘oculomasticatory myo- lar dysplasia and hypoplasia (e.g., as part of Joubert’s syn- rhythmia.’’ drome), are found in a number of patients with congenital 7. In Joubert’s syndrome, patients commonly have skew ocular motor apraxia (721–724). Most likely, however, such deviations that change with lateral gaze and also sponta- anomalies are markers of abnormal development rather than neously alternate as to which eye is higher, along with being directly responsible for the eye movement disorder. seesaw and pendular nystagmus. Patients with Joubert’s Congenital ocular motor apraxia is occasionally familial and syndrome may also have pigmentary degeneration of has been reported in monozygotic twins (725,726). Patients the as well as breathing disturbances early in life. usually improve with age, with the head movements becom- 8. In vitamin E deficiency (e.g., abetalipoproteinemia), a ing less prominent as the patients are better able to direct slow but full range of abduction and limited range but their eyes voluntarily. normal speed of adduction are characteristic. A variety of disorders that directly involve the brain stem 9. Patients with ataxia-telangiectasia have other cerebellar mechanisms for generating saccades, including structural or eye signs and elevated levels of alpha-fetoprotein. degenerative processes within the pontine and mesence- 10. In childhood Huntington’s disease, slowing of saccades phalic reticular formations, are characterized by the develop- is prominent, and rigidity may be more apparent than ment of a strategy of head thrusting or blinking to shift gaze chorea. that superficially resembles congenital ocular motor apraxia. These disorders usually can be differentiated from congeni- The role of blinking in facilitating eye movements in pa- tal ocular motor apraxia because all types of saccades and tients with congenital ocular motor apraxia may be related 948 CLINICAL NEURO-OPHTHALMOLOGY to a natural synkinesis of blinking and saccade-generating OCULAR MOTOR MANIFESTATIONS OF SEIZURES or in some cases to lesions in the posterior fossa. Zee et al. (308) studied two patients who could make saccades of nor- Eye and head movements are common manifestations of mal velocity and amplitude only in association with a simul- epileptic seizures, and the wide representation of oculomotor taneous blink. In one patient, the initiation of saccades was and vestibular control within the leads to a also facilitated by blinks, much like patients with typical number of possible ways that seizures can affect eye move- congenital ocular motor apraxia; however, both patients had ments themselves and visual or vestibular perceptions (670). other signs of cerebellar or brain stem dysfunction, suggest- A variety of eye movements can occur with seizures, in- ing a posterior fossa localization for blink facilitation of sac- cluding horizontal or vertical gaze deviation, skew deviation, cadic velocity and amplitude. and spontaneous, retractory, periodic alternating or monocu- lar nystagmus (122,766–776). Epileptic convergence nys- tagmus may occur with periodic lateralizing epileptiform ABNORMAL EYE MOVEMENTS AND DEMENTIA discharges (777) and with burst-suppression patterns Patients with various dementing processes have abnormal (778,779). The seizure focus may arise from any lobe, al- eye movements, reflecting either disturbances in cerebral though the lesions usually are more posterior. Epileptic nys- cortical structures or in other subcortical structures that may tagmus can occur with typical absence seizures (780) and with infantile spasms (781). also be affected by that particular disease. Excessive errors Patients with epileptic foci affecting the temporoparieto- on the antisaccade test (736–739), particularly when associ- occipital cortex may show either ipsiversive or contraversive ated with a ‘‘visual grasp reflex’’ (741), are a useful indicator eye deviation and nystagmus (766,782–784). Seizures are of an organic process when pseudodementia is a diagnostic usually accompanied by relatively fast cortical activity. Ip- consideration in a patient with a possible cognitive decline. siversive gaze deviation may be caused by a smooth-pursuit Patients with Alzheimer’s disease have excessive numbers movement, with the quick phases of nystagmus being trig- of square-wave jerks and defects in saccade latency and, gered by the eccentric eye position and ipsilateral drift of occasionally, accuracy and velocity (737,739,740). Alzhei- the eyes. Contraversive deviation may be initiated by sac- mer’s patients show longer mean fixation durations and a cades and the subsequent nystagmus from an impaired gaze- reduced number of exploring saccades when viewing simple holding mechanism (deficient neural integrator). Furman et but not complex scenes, perhaps reflecting a motivation defi- al. (678) described a patient with a temporoparietal seizure cit. Impaired predictive tracking and excessive number of focus who showed no gaze deviation before the onset of anticipatory saccades are a feature of Alzheimer’s disease nystagmus. Her attacks were accompanied by vertigo, sug- (737), as is difficulty generating saccades during reading gesting involvement of the ‘‘vestibular cortex’’ of the supe- (738). Eye movements may be used to monitor higher-level rior temporal sulcus. processes, such as ‘‘curiosity,’’ by looking at exploratory Overall, contraversive deviation of the eyes is more com- behavior (visual search and scan paths) while a subject views mon than ipsiversive deviation during seizures. In cases with novel visual stimuli (587). This capability too appears to posterior foci (temporal, parietal, or occipital lobes), experi- be diminished in Alzheimer’s patients (742). Impairment of mental studies suggest that eye movements may be mediated spatially directed attention may also be reflected in eye by projections via either the superior colliculus or the frontal movement abnormalities (743), and a Ba´lint-like syndrome eye fields (785). Thus, saccades may be generated by more may develop (744). Pursuit abnormalities also occur in Alz- than one of the descending parallel pathways. Frontal lobe heimer’s patients (745,746). foci are also reported to cause contraversive deviations un- Patients with Creutzfeldt-Jakob disease may show limita- less they are bilateral, in which case vertical deviations also tion of vertical gaze and slow vertical saccades (747,748) may occur (786). These results are consistent with stimula- as well as two rare forms of nystagmus: periodic alternating tion studies in monkeys, in which unilateral stimulation of nystagmus (748) and centripetal nystagmus (749). Over- the frontal eye fields typically causes oblique saccades with doses of lithium or bismuth may lead to syndromes that a contralateral horizontal component; the direction of the mimic Creutzfeldt-Jakob disease (750,751). Cerebellar eye vertical saccade depends upon a cortical map (787). Purely signs are typically found in another prion disorder: Gerst- vertical movements require bilateral stimulation of the fron- mann-Stru¨ssler-Scheinker disease (752,753). tal eye fields. Because there are also neurons in the frontal Patients infected with the human immunodeficiency virus eye fields that contribute to smooth pursuit, it is theoretically (HIV) show a number of ocular motor abnormalities that possible that frontal lobe foci could lead to an ipsiversive usually reflect the effects of an opportunistic infection or deviation. coexistent neoplasia (754–758). However, HIV encephalop- Head turning is a common accompaniment of epileptic athy itself can cause disturbed ocular motility (759), includ- gaze deviation. In patients who are conscious during the ing errors on the antisaccade task, increased fixation instabil- seizure, a frontal focus is likely, and the initial direction of ity, increased latencies of horizontal and, especially, vertical head turning is usually, but not invariably, contralateral to saccades (760), acquired ocular motor apraxia (761), cere- the seizure focus (788–790). A contralateral focus is also bellar and brain stem signs, including gaze-evoked and dis- likely in a patient who shows marked, sustained, and unnatu- sociated nystagmus (762,763), slow saccades (764), and de- ral lateral positioning of the head and eyes (789,790). In creased but especially asymmetric pursuit gain (760,765). patients who are unconscious during the seizure, the focus SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 949 may arise from any lobe, and head turning may be toward or of the lids (eyes-open coma) may also occur in unconscious away from the side of the lesion (788,791,792). As discussed patients and may be related to pontomesencephalic dysfunc- above, seizures emanating in the superior temporal lobes tion (803). may cause a variety of vestibular sensations (670), and oc- Deviations of the visual axes in coma may be caused by cipital lobe seizures may produce oscillopsia (793). Rarely, skew deviation, by a phoria that is normally compensated for seizures are precipitated by movements of the eyes, such as by fusional mechanisms, or by paralysis of the oculomotor, convergence (794) or sustained lateral deviation (795). trochlear, or abducens nerves. Restrictive ophthalmopathy, particularly blowout fracture of the orbit, may be an addi- EYE MOVEMENTS IN STUPOR AND COMA tional mechanism in patients who have suffered facial trauma. The diagnosis of the cause of the deviation depends The ocular motor examination is especially useful for upon determining whether the range of movement of the evaluating the unconscious patient, because both arousal and eyes induced by head rotation or caloric stimulation (dis- eye movements are controlled by neurons in the brain stem cussed later) is reduced in a pattern corresponding to specific reticular formation. Comatose patients do not make eye muscle weakness. In addition, assessment of the pupils and movements that depend upon cortical visual processing. Vol- other brain stem reflexes may help. Complete oculomotor untary saccades and smooth pursuit are in abeyance, and nerve palsy causes pupillary dilation, ipsilateral ptosis, and quick phases of nystagmus also may be absent. The ocular deviation of the eye down and out. Pupillary involvement motor examination of the unconscious patient, therefore, is an early sign of uncal herniation (797), and disturbances consists of observing the resting position of the eyes, looking of eye movements usually follow (804). Vertical misalign- for any spontaneous movements, and reflexively inducing ment of the eyes is usually caused by skew deviation or eye movements (796–798). trochlear nerve palsy, the latter being particularly common Gaze Deviations following head trauma. Bilateral abducens palsy occurs when increased intracranial pressure compromises the Conjugate, horizontal deviation of the eyes is common in nerves as they bend over the petroclinoid ligament. Occa- coma. When this is caused by lesions above the brain stem sionally, skew deviation and INO occur in patients with met- ocular motor decussation between the midbrain and pons, abolic encephalopathy or drug intoxication. the eyes are usually directed toward the side of the lesion and away from the hemiparesis. A vestibular stimulus, however, Spontaneous Eye Movements usually drives the eyes across the midline. If the conjugate deviation is caused by a lesion below the ocular motor decus- Spontaneous eye movements that occur in unconscious sation, the eyes will be directed away from the side of the patients may help establish the etiology of the coma. Slow lesion and toward the hemiparesis. The latter is typically conjugate or disconjugate roving eye movements are similar seen with pontine lesions, but also in some patients with to the eye movements of light sleep but slower than the rapid thalamic (18) and rarely hemispheric disease above the thala- eye movements (REM) of paradoxic or REM sleep. If a mus (so-called wrong-way deviations) (305,682). complete range can be shown, their presence indicates that Intermittent deviations of the eyes and head are usually brain stem gaze mechanisms are intact (796). caused by seizure activity. At the onset of each attack, gaze Other types of spontaneous eye movements consist of var- is usually deviated contralateral to the side of the seizure ious forms of vertical to-and-fro movements, often called focus and may be followed by nystagmus with contralat- ‘‘bobbing.’’ Typical ocular bobbing consists of intermittent, erally directed quick phases. Toward the end of the seizure, usually conjugate, rapid downward movement of the eyes gaze drifts to an ipsilateral (paretic) position. followed by a slower return to the primary position Tonic downward deviation of the eyes, often accompanied (805,806). Reflex horizontal eye movements are usually ab- by convergence, occurs in patients with thalamic hemor- sent. Ocular bobbing is a classic sign of intrinsic pontine rhage (453,456) and with lesions affecting the dorsal mid- lesions, usually hemorrhage, but it also occurs in patients brain. It may be induced by unilateral caloric stimulation, with cerebellar lesions that compress the pons and in meta- after the initial horizontal deviation subsides, in patients with bolic or toxic encephalopathy. Inverse bobbing also occurs coma induced by drugs (799). Forced downward in this setting. This eye movement abnormality, which is also deviation of the eyes can also be seen in patients with nonor- called ocular dipping, is characterized by a slow downward ganic (feigned) coma or seizures (800). movement, followed by a rapid return to midposition. Re- Tonic upward deviation of the eyes is uncommon in coma, verse bobbing consists of rapid deviation of the eyes upward but it may be seen after an hypoxic-ischemic insult, even and a slow return to the horizontal, whereas converse bob- when no pathologic lesions are found in the midbrain (801). bing (also called reverse dipping) is characterized by a slow Patients who survive after manifesting this ocular motor dis- upward drift of the eyes that is followed by a rapid return turbance typically develop downbeating nystagmus, the up- to primary position. These variants of ocular bobbing are ward drift of which is thought to be caused by loss of inhibi- less reliable for localization than is straightforward ocular tion on the upward vertical VOR (802). Upward deviation bobbing. Nevertheless, the occurrence in some patients of of the eyes also occurs as a component of oculogyric crisis, several different types of ocular bobbing during their illness which usually occurs as a side effect of certain drugs, espe- suggests a common underlying pathophysiology (807–810). cially neuroleptic agents (521). Tonic uninhibited elevation Because the pathways that mediate upward and downward 950 CLINICAL NEURO-OPHTHALMOLOGY eye movements differ anatomically and probably pharmaco- jects, this is unlikely to be the case in unconscious patients. logically, it seems likely that these movements represent a Therefore, eye rotations induced by head rotation in uncon- varying imbalance of mechanisms for vertical gaze. Rarely, scious individuals principally result from the effects of the large-amplitude pendular vertical oscillations occur in the semicircular canals and their central connections (i.e., the acute phase of a brain stem stroke (804). VOR). The head should not be rotated in unconscious pa- Repetitive vertical eye movements, including variants of tients unless there is certainty that no neck injury or abnor- ocular bobbing, may contain convergent-divergent compo- mality is present. Conventionally, high-frequency (1–2 Hz) nents. Such movements are usually caused by disease affect- quasi-sinusoidal rotations or position-step stimuli are ap- ing the dorsal midbrain (109,808,812). plied; the latter consists of a sudden head turn to a new Monocular bobbing movements may occur as a synkinesis position. Both horizontal and vertical rotations should be with jaw movement. Such movements are similar to those performed. If small-amplitude head rotations are performed, seen in the congenital condition called the Marcus Gunn the adequacy of the VOR can be estimated by observing the jaw-winking phenomenon and primarily involve the neural of one eye with an ophthalmoscope (344). pathways to the inferior rectus muscle (813). Caloric irrigation of the external auditory meatus causes Ping-pong gaze consists of slow, horizontal, conjugate convection currents of the vestibular endolymph that dis- deviations of the eyes that alternate every few seconds. In place the cupula of a semicircular canal; thus, this procedure more alert patients, ping-pong gaze has been shown to be a also tests the VOR. The canal stimulated depends on the series of small saccades that take the eyes back and forth orientation of the head (e.g., with the head elevated 30Њ from from one extreme orbital position to the other (814,815). the supine position, the horizontal canals are principally Although ping-pong gaze can occur in patients with posterior stimulated). Before caloric stimulation, the physician should fossa hemorrhage (816), it is usually a sign of bilateral in- always check that the tympanic membrane is intact. Usually farction of the cerebral hemispheres. In one reported case, only about 5 cc of ice water need be introduced into the no cerebral hemispheric lesions were noted, but bilateral external auditory meatus, but large quantities (100 cc or lesions of the cerebral peduncles were present (817). Some- more) may be necessary to induce a response in some coma- times oscillations with a periodicity similar to that of ping- tose patients. pong gaze can be induced transiently by a rapid head rotation Caloric stimulation with ice water is sometimes a more in patients with bilateral hemisphere disease (818). effective stimulus than head rotation, in part because of the Rapid, small-amplitude, vertical eye movements may be sustained nature of the stimulus and in part because of the the only manifestation of epileptic seizures in patients with arousing effect of the cold water. Combined cold caloric coexistent brain stem injury (819). Rapid, monocular eye stimulation and head rotation may be the most effective stim- ulus in the deeply unconscious patient (796). This usually movements with horizontal, vertical, or torsional compo- produces tonic deviation of the eyes toward the irrigated ear nents that occur in patients with coma may also indicate in patients with intact vestibular function. brain stem dysfunction. In testing reflex eye movements in unresponsive patients, Identification of patients who are conscious but quadri- it is important to note (a) the magnitude of the response; (b) plegic, the locked-in syndrome or de-efferented state, de- whether the ocular deviation is conjugate; (c) the dynamic pends on identifying preserved voluntary vertical eye move- response to position-step head rotations; and (d) the occur- ments (797,820). The syndrome is typically caused by rence of any quick phases of nystagmus, particularly during pontine infarction and is characterized in part by a variable caloric stimulation. Impaired abduction suggests abducens loss of voluntary and reflex horizontal movements, such that nerve palsy. Impaired adduction usually indicates either INO eyelid or vertical eye movements may be the only means or , although occasionally impaired of communication during the acute illness. The locked-in adduction to vestibular stimulation is observed in patients syndrome also occurs with midbrain lesions, in which case with metabolic coma or drug intoxication. Vertical responses ptosis and ophthalmoplegia may be present (821). may be impaired with disease of the midbrain (822) or bilat- Reflex Eye Movements eral lesions of the MLF. Pontine lesions may abolish the reflex eye movements in the horizontal plane but spare the Reflex eye movements may be elicited in unconscious vertical responses. When reflex eye movements are present patients either by head rotation (the doll’s head or oculoce- in an unresponsive patient, the brain stem is likely to be phalic maneuver) or by caloric stimulation (798). Head rota- structurally intact. When reflex eye movements are abnormal tion, with the patient supine, stimulates the labyrinthine or absent, the cause may be structural disease, profound met- semicircular canals, the otoliths, and neck muscle proprio- abolic coma, or drug intoxication (823). Intoxication with a ceptors. However, unless there has been prior loss of vestibu- variety of may lead to partial or total loss of lar function (e.g., from aminoglycoside toxicity), the contri- reflexive eye movements. Metabolic disturbances can cause bution made by neck muscle proprioceptors to generating restriction of reflex eye movements, but usually only with reflex eye movements (the cervico-ocular reflex) is insignifi- profound coma (824). When used in combination with other cant (818). Furthermore, the otolith contribution is probably clinical signs, reflex eye movements are useful in evaluating small compared with that of the labyrinthine semicircular the prognosis of comatose patients (825,826). canals. Finally, although visually mediated eye movements, If reflex eye movements are intact in an unconscious pa- such as fixation and smooth pursuit, can influence the eye tient, the eyes are carried into a corner of the orbit when the movements produced by head rotation in normal, awake sub- head is rapidly rotated horizontally to a new position (posi- SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 951 tion step stimulus). If the head is held stationary in its new Singh and Strobos (827) induced retraction nystagmus with position, the eyes may slowly drift back to the midline. This caloric stimulation. Patients who survive coma but who are implies that the gaze-holding mechanism (neural integrator) left in a persistent vegetative state, with severe damage of is not functioning normally. Patients with more rapid centri- the cerebral hemispheres but preservation of the brain stem petal drift may have more severe brain injury (818). (828), regain nystagmus with caloric or rotational stimula- Quick phases of nystagmus are usually absent in acutely tion (818). Caloric nystagmus was present in patients with unconscious patients. Their presence, without a tonic devia- neocortical death and an isoelectric electroencephalogram tion of the eyes, should raise the possibility of nonorganic (829). (i.e., feigned) coma. In patients who are stuporous but un- Normal subjects have been studied during syncope (830). cooperative, caloric nystagmus may be a useful way of in- Subjects developed downbeat nystagmus and tonic upward ducing eye movements that cannot be initiated voluntarily. deviation of the eyes. There was also an increased amplitude For example, in a patient with a pineal tumor but no overt of the VOR in most of the subjects. These findings are most abnormalities during testing of voluntary eye movements, compatible with cerebellar hypoperfusion.

OCULAR MOTOR DYSFUNCTION AND MULTIPLE SCLEROSIS MS causes a variety of ocular motor deficits; bilateral (610,843,844), and impaired cancellation of the horizontal INO, cerebellar eye signs (including gaze-evoked nystag- VOR may be present (610). Abnormalities of vertical gaze- mus), and acquired pendular nystagmus are the most com- holding, smooth pursuit, and eye-head tracking occur in pa- mon (224,250,831–835) (see Chapter 60). Pendular or ellip- tients with an INO (359). Other abnormalities in patients tical nystagmus is a frequent disabling manifestation of the with MS include horizontal and vertical gaze palsies (845), disease (836,837). Measurement of eye movements may gaze-evoked blepharoclonus (846), vertical nystagmus, sac- help establish the diagnosis of MS during early stages of cadic pulsion (847), various vestibular abnormalities (832), the disease. Saccadic abnormalities may be more readily and superior oblique myokymia. detected when targets are presented randomly so that neither Diagnosis of early MS depends on the demonstration of their time of onset nor their location can be predicted. Large lesions disseminated throughout the nervous system. Subtle saccades (20Њ or more) are more likely to show changes deficits of ocular motility can provide a sensitive method in velocity than are small saccades (838). Because normal for identifying subclinical lesions, but there is a need for subjects show differences in the peak velocity of abducting caution because these tests are not specific for MS (848). and adducting saccades, it is worth developing a normative The clinician must compare the results of ocular motor stud- database for each of these. With these provisos, most patients ies with other clinical or laboratory findings before making with MS show saccadic abnormalities that include prolonged a diagnosis (849). latency, inaccuracy, and decreased velocity. In particular, A number of drugs, including gabapentin and minor slowing of adducting saccades may indicate an INO (an agent with NMDA blocking, AMPA receptor modula- that may not be evident with clinical testing (839). tion, and dopaminergic action) may ameliorate the visually Some patients with MS show saccadic oscillations disabling pendular nystagmus commonly observed in MS (243,840–842). Smooth-pursuit gain may be decreased (12,850).

OCULAR MOTOR MANIFESTATIONS OF SOME METABOLIC DISORDERS Some babies who ultimately develop normally show tran- idase deficiency also preferentially affects vertical gaze sient ocular motor disturbances, including upward or down- (855). Variants of Niemann-Pick disease that begin after the ward deviation of the eyes (but with a full range of reflex first year of life (e.g., the sea-blue histiocyte syndrome or vertical movement), intermittent opsoclonus, and skew de- juvenile dystonic lipidosis) are characterized by deficits of viation (86,320,851,852). The last may be associated with voluntary vertical eye movements, particularly saccades and the eventual development of horizontal strabismus. In addi- smooth pursuit; vertical vestibular and horizontal eye move- tion, premature babies may show reduced excursion of the ments are relatively preserved (856–859) (Fig. 19.32). adducting eye with caloric stimulation, suggesting an INO. Gaucher’s disease is associated with a prominent deficit of In such infants, full deviation of both eyes usually occurs horizontal gaze, and slow saccades may be a prominent find- with rotational stimuli (853), although quick phases of nys- ing in adults with this condition (860). tagmus may be absent (854). The time constant of the VOR Wernicke’s encephalopathy is characterized by the triad (as reflected in the duration of nystagmus to a sustained of ophthalmoplegia, mental confusion, and gait ataxia. It is constant velocity stimulus) in newborns is low (typically 6 caused by thiamine deficiency and is most commonly en- seconds) and does not attain adult values until the infant is countered in alcoholics. The ocular motor findings include about 2 months old (854). Thus, normal infants may have weakness of abduction, gaze-evoked nystagmus and primary abnormal eye movements. position vertical nystagmus, impaired vestibular responses The lipid storage diseases are often characterized by gaze to caloric and rotational stimulation, INO, the one-and-a- palsies. Tay-Sachs disease impairs vertical and, subse- half syndrome, and horizontal and vertical gaze palsies that quently, horizontal eye movements. Adult-onset hexosamin- may progress to total ophthalmoplegia (861–866). The oph- 952 CLINICAL NEURO-OPHTHALMOLOGY thalmoplegia is bilateral but may be asymmetric. Experi- (113) reported seesaw nystagmus and the OTR in a patient mental thiamine deficiency in monkeys causes an orderly with Leigh’s disease. progression of ophthalmoplegia associated with well-cir- Pelizaeus-Merzbacher disease is an X-linked recessive cumscribed histopathologic changes (867). These changes leukodystrophy (877) with severe cerebellar signs, including consist of neuronal loss and gliosis in the oculomotor, troch- saccadic dysmetria. Some patients also show difficulty initi- lear, abducens, and vestibular nuclei. In humans, demyelin- ating saccades and pendular nystagmus (730,878). ation, vascular changes, and hemorrhage also may occur. In Deficiency of vitamin E may cause a progressive neuro- addition to the sites listed above, lesions are found in the logic condition characterized by areflexia, , paraventricular regions of the thalamus, hypothalamus, peri- and loss of joint position sense. Ocular motor involvement aqueductal gray matter, superior vermis of the cerebellum, includes progressive gaze restriction, sometimes with stra- and dorsal motor nucleus of the vagus. Thus, gaze-evoked bismus. Vitamin E deficiency is more common in children, nystagmus and the impaired caloric responses can be attrib- in whom it may be caused by abetalipoproteinemia (Bassen- uted to vestibular nucleus involvement. The abduction weak- Kornzweig disease). It is also reported in adults who have ness and INO may reflect damage to the abducens nerve and bowel or liver diseases that interfere with fat absorption or the MLF, respectively, whereas paralysis of horizontal gaze as an inherited ataxia on chromosome 8q13, the site of the may be caused by damage to the PPRF or abducens nucleus. ␣-tocopherol transfer protein gene (879). Vitamin E defi- The one-and-a-half syndrome indicates dysfunction of the ciency is characterized by a dissociated ophthalmoplegia and MLF and either the abducens nucleus or the PPRF, and total by nystagmus in which adduction is fast but with a limited ophthalmoplegia may be caused by damage to all of the range and abduction is slow but with a full range (880). A ocular motor nerve nuclei. Most likely, affected areas of the posterior INO of Lutz (discussed previously) occurs in some brain contain neurons that use high amounts of glucose and, patients (254,255). The combination of ocular motor find- therefore, are particularly dependent upon thiamine, an im- ings in patients with vitamin E deficiency probably reflects portant coenzyme in glucose metabolism (868). Administra- a mixture of central and peripheral pathology. tion of thiamine usually causes rapid improvement of the Wilson’s disease, hepatolenticular degeneration, is an in- ocular motor signs, although complete recovery may take herited disorder of copper metabolism that is transmitted in several weeks. Coexistent magnesium deficiency should also an autosomal-recessive fashion. The defect is in a copper- be treated. In patients with Wernicke’s disease who go on transporting adenosine triphosphatase (ATPase) gene at to develop Korsakoff’s syndrome, which is primarily charac- q14.3 on chromosome 13. CT scanning in patients with this terized by a severe and enduring memory loss, ocular motor condition shows hypodense areas, and PET scanning indi- abnormalities may persist (869,870). The ocular motor ab- cates a decreased rate of glucose metabolism in the globus normalities include slow and inaccurate saccades, impaired pallidus and putamen (881). The classic clinical picture is smooth pursuit, and gaze-evoked nystagmus. MR imaging a prominent dysarthria, abnormal movements, psychiatric shows a distinctive pattern of abnormality in Wernicke’s symptoms, and liver disease. Ocular motor disorders in hepa- disease, with T2 intensity in structures around the third and tolenticular degeneration include a distractibility of gaze, fourth ventricles and the aqueduct (871). with inability to voluntarily fix upon an object unless other Leigh’s syndrome is a fatal subacute necrotizing encepha- competing visual stimuli are removed (e.g., fixation of a lopathy of infancy or childhood characterized by psychomo- solitary light in an otherwise dark room) (882). Slow sac- tor retardation, seizures, and brain stem abnormalities, in- cades were reported in one patient with this condition (883), cluding eye movements. It is an inherited disorder caused and apraxia of eyelid opening can also occur (884). either by abnormalities of mitochondrial deoxyribonucleic Amyotrophic lateral sclerosis is associated with various acid (DNA), in which case it is maternally inherited, or by eye-movement disorders, including nystagmus (885), sac- a chromosomal abnormality, in which case it is transmitted cade disturbances that suggest a frontal lobe disturbance as an autosomal-recessive disorder. Some cases show an un- (886–888), and impaired pursuit (889). Occasionally slow derlying defect of pyruvate dehydrogenase (872); others re- saccades are a feature (886). However, the existence of mul- flect a mitochondrial cytopathy (873,874). Both the distur- tisystem diseases in which motor neuron degeneration is just bances of ocular motility and the pathologic findings one neurologic feature makes specific clinical diagnoses dif- resemble those caused by experimental thiamine deficiency ficult (890,891). Ocular motor abnormalities were reported or Wernicke’s encephalopathy (875,876). Halmagyi et al. in a single patient with Kugelberg-Welander disease (892).

EFFECTS OF DRUGS ON EYE MOVEMENTS Many substances affect eye movements (Table 19.8). In Patients with drug-induced abnormalities of eye move- some cases, the drug induces abnormalities of eye move- ments most often complain of diplopia, caused by ocular ments at therapeutic concentrations (e.g., anticonvulsants) misalignment, or oscillopsia, caused by spontaneous nystag- (893). In other cases, abnormalities of eye movements de- mus or an inappropriate VOR (894). Many drugs have their velop only when concentrations of the drug in the central effect on central vestibular and cerebellar connections and nervous system are inappropriately elevated. In still other cause ataxia and gaze-evoked nystagmus (895). cases, the eye-movement abnormalities are caused by sub- Although all classes of eye movements may be affected by stances not meant for internal use. therapeutic doses of various drugs, smooth pursuit, eccentric SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 953

Table 19.8 ties, including fixation instability and downbeat nystagmus Effect of Drugs on Eye Movements (908). In one patient who showed marked impairment of all types of horizontal eye movements and downbeat nystagmus Benzodiazepines: Reduced velocity and increased duration of saccades, prior to death, neuronal loss was mainly confined to the NPH impaired smooth pursuit, decreased gain and increased time constant of and adjacent MVN(909). Thus, the pathophysiology was VOR, divergence paralysis Tricyclic antidepressants: Internuclear ophthalmoplegia, partial or total similar to that produced by experimental lesions of these gaze paresis, opsoclonus nuclei in monkeys (910); that is, the neural integrator (gaze- Phenytoin: Impaired smooth pursuit and VOR suppression, gaze-evoked holding network) was disrupted. The propensity of lithium nystagmus, downbeat nystagmus, periodic alternating nystagmus, total to damage this area of the medulla is related to its proximity gaze paresis, convergence spasm to the plexus of the fourth ventricle and thus to : Decreased velocity of saccades, impaired smooth pursuit, elevated local concentrations of lithium (909). Cerebellar gaze-evoked nystagmus, oculogyric crisis, downbeat nystagmus, total damage may also occur after lithium intoxication (911). gaze paresis Amiodarone can cause prominent cerebellar eye signs, in- Phenobarbital and other barbiturates: Reduced peak saccadic velocity, cluding downbeat nystagmus (912). gaze-evoked nystagmus, impaired smooth pursuit, impaired vergence, In addition to drugs, certain can cause abnormal eye decreased VOR gain, perverted caloric responses, vertical nystagmus, partial or total gaze paresis movements. Some, such as chlordecone (913) and thallium Phenothiazines: Oculogyric crisis, internuclear ophthalmoplegia (914), cause saccadic oscillations. Intoxication with hydro- Lithium carbonate: Saccadic dysmetria, impaired smooth pursuit, gaze- carbons can cause a vestibulopathy (915,916), and exposure evoked nystagmus, downbeat nystagmus, oculogyric crisis, internuclear to trichloroethylene and other solvents may affect pursuit, ophthalmoplegia, total gaze paresis, opsoclonus suppression of the VOR, and saccades (917,918). Prolonged : Reduced saccade latency, increased accommodative exposure to toluene, especially in glue-sniffing addiction, convergence/ ratio may lead to a variety of ocular motor disturbances, including Ethyl : Reduced peak velocity, increased latency, and hypometria of pendular (919) and downbeat (920) nystagmus, saccadic os- saccades, impaired smooth pursuit and VOR suppression, gaze-evoked cillations (921–923), and INO (924). has a num- nystagmus, positionally induced nystagmus, normal static VOR ber of ocular motor effects. It causes upbeat nystagmus compensation, reversed compensation of vestibular lesions (925,926), impaired pursuit (926), decreased saccade latency : Upbeat nystagmus, in darkness: Square-wave jerks, impaired horizontal and vertical pursuit (927), and increased square-wave jerks during pursuit (928), Methadone: Saccadic hypometria, impaired smooth pursuit although performance is normal on the antisaccade test Baclofen: Reduced VOR time constant, total gaze paresis (898,929,930). can affect eye movements, with op- ␤-Blockers: Diplopia, internuclear ophthalmoplegia soclonus being the most dramatic abnormality (931). Chloral hydrate: Impaired smooth pursuit Ototoxicity, especially that associated with administration Nitrous oxide: Reduced saccadic peak velocity, impaired smooth pursuit of aminoglycosides, is an important cause of VOR loss (934). Intravenous gentamicin is most often responsible. Its VOR, vestibulo-ocular reflex. toxicity may be insidious, occurring without hearing symp- toms and even with normal blood levels and relatively short periods of administration (935). Some patients who develop gaze holding, and convergence are particularly susceptible. ototoxicity may be genetically predisposed to its toxic side For example, diazepam, methadone, phenytoin, barbiturates, effects (936,937). Topical (intratympanic) gentamicin is chloral hydrate, and alcohol all impair smooth-pursuit track- used to purposefully ablate labyrinthine function as part of ing. Some drugs have specific effects on ocular motility and the treatment of intractable Me´nie`re’s syndrome but occa- thus provide insights into both the function of the ocular sionally leads to unwanted labyrinthine loss when used to motor system and the mode of drug action. For example, treat external ear infections (938). Cisplatin is probably not diazepam reduces saccadic peak velocity but does not impair as vestibulotoxic as originally thought (939–941). accuracy; methadone shows the converse effect. Diazepam reduces the gain of the VOR; in experimental animals the REFERENCES time constant is prolonged, but in humans it is reduced (896). 1. Leigh RJ, Zee DS. The Neurology of Eye Movements. Philadelphia: FA Davis, 1999. In contrast, baclofen reduces the time constant of the VOR 2. Arnold DB, Robinson DA, Leigh RJ. Nystagmus induced by pharmacological (897). Nicotine causes upbeat nystagmus (898–900). Alco- inactivation of the ocular motor integrator in monkey. Vision Res hol affects eye movements by virtue of its effects on the 1999;39:4286–4295. 3. Tilikete C, Hermier M, Pelisson D, Vighetto A. Saccadic lateropulsion and density of endolymph and the cupula in the labyrinth as well upbeat nystagmus: disorders of caudal medulla. Ann Neurol 2002;52:658–662. as on more central structures (901–906). 4. Ranalli PJ, Sharpe JA. Upbeat nystagmus and the ventral tegmental pathway of At toxic levels, neuroactive drugs can impair all eye the upward vestibulo-ocular reflex. Neurology 1988;38:1329–1330. 5. Guillain G. The syndrome of synchronous and rhythmic palato-pharyngo- movements, particularly when consciousness is also im- laryngo-oculo-diaphragmatic myoclonus. Proc R Soc Med (UK) 1938;31: paired. Phenytoin may cause a complete ophthalmoplegia in 1031–1038. 6. Deuschl G, Toro C, Valls-Sole´ J, et al. Symptomatic and essential palatal tremor. awake patients, and therapeutic levels may cause ophthal- 1. Clinical, physiological, and MRI analysis. Brain 1994;117:767–774. moplegia in patients in stupor (823). Phenytoin and diaze- 7. Schwankhaus JD, Parisi JE, Gulledge WR, et al. Hereditary adult-onset Alexan- pam can lead to opsoclonus (907). The tricyclic antidepres- der’s disease with palatal myoclonus, spastic paraparesis, and cerebellar ataxia. Neurology 1995;45:2266–2271. sants may cause complete ophthalmoplegia or an INO in 8. Deuschl G, Wilms H. Palatal tremor: the clinical spectrum and physiology of stuporous patients. Lithium causes a variety of abnormali- a rhythmic movement disorder. Adv Neurol 2002;89:115–130. 954 CLINICAL NEURO-OPHTHALMOLOGY

9. Gresty MA, Ell JJ, Findley LJ. Acquired pendular nystagmus: its characteristics, 43. Malis DD, Guyot JP. Room tilt illusion as a manifestation of peripheral vestibular localising value and pathophysiology. J Neurol Neurosurg Psychiatry 1982;45: disorders. Ann Otol Rhinol Laryngol 2003;112:600–605. 431–439. 44. Kommerell G, Hoyt WF. Lateropulsion of saccadic eye movements. Electro- 10. Nakada T, Kwee IL. Oculopalatal myoclonus. Brain 1986;109:431–441. oculographic studies in a patient with Wallenberg’s syndrome. Arch Neurol 11. Jacobs L, Bender MB. Palato-ocular synchrony during eyelid closure. Arch Neu- 1973;28:313–318. rol 1976;33:289–291. 45. Hornsten G. Wallenberg’s syndrome. II. Oculomotor and oculostatic distur- 12. Stahl JS, Rottach K, Averbuch-Heller L, et al. A pilot study of gabapentin as bances. Acta Neurol Scand 1974;50:447–468. treatment for acquired nystagmus. Neuroophthalmology 1996;16:107–113. 46. Baloh RW, Yee RD, Honrubia V. Eye movements in patients with Wallenberg’s 13. Kulkarni PK, Muthane UB, Taly AB, et al. Palatal tremor, progressive multiple syndrome. Ann NY Acad Sci 1981;374:600–613. cranial nerve palsies, and cerebellar ataxia: a case report and review of literature 47. Waespe W, Baumgartner R. Enduring dysmetria and impaired gain adaptivity of palatal tremors in neurodegenerative disease. Mov Disord 1999;14:689–693. of saccadic eye movements in Wallenberg’s lateral medullary syndrome. Brain 14. Koeppen AH, Barron KD, Dentinger MP. Olivary hypertrophy: histochemical 1992;115:1125–1146. demonstration of hydrolytic enzymes. Neurology 1980;30:471–480. 48. Bu¨ttner U, Straube A. The effect of cerebellar midline lesions on eye movements. 15. Guillain G, Mollaret P. Deux cas myoclonies synchrones et rhythme´es ve´lo- Neuroophthalmology 1995;15:75–82. pharyngo-laryngooculo-diaphragmatiques: le proble`me anatomique et physio- 49. Solomon D, Galetta SL, Liu GT. Possible mechanisms for horizontal gaze devia- pathologique de ce syndrome. Rev Neurol (Paris) 1931;2:545–566. tion and lateropulsion in the lateral medullary syndrome. J Neuroophthalmol 16. Alpini D, Caputo D, Pugnetti L, et al. Vertigo and multiple sclerosis: aspects 1995;15:26–30. of differential diagnosis. Neurol Sci 2001;22:584–587. 50. Kirkham TH, Guitton D, Gans M. Task-dependent variations of ocular lateropul- 17. Frohman EM, Kramer PD, Dewey RB, et al. Benign paroxysmal positioning sion in Wallenberg’s syndrome. Can J Neurol Sci 1981;8:21–22. vertigo in multiple sclerosis: diagnosis, pathophysiology and therapeutic tech- 51. Ohtsuka K, Sawa M, Matsuda S, et al. Nonvisual eye position control in a patient niques. Mult Scler 2003;9:250–255. with ocular lateropulsion. Ophthalmologica 1988;197:85–89. 18. Fisher CM. Some neuro-ophthalmological observations. J Neurol Neurosurg 52. Morrow MJ, Sharpe JA. Torsional nystagmus in the lateral medullary syndrome. Psychiatry 1967;30:383–392. Ann Neurol 1988;24:390–398. 19. Gomez CR, Cruz-Flores S, Malkoff MD, et al. Isolated vertigo as a manifestation 53. Helmchen C, Glasauer S, Bu¨ttner U. Pathological torsional eye deviation during of vertebrobasilar ischemia. Neurology 1996;47:94–97. voluntary saccades: a violation of Listing’s law. J Neurol Neurosurg Psychiatry 20. Baloh RW. Vertebrobasilar insufficiency and stroke. Otolaryngol Head Neck 1997;62:253–260. Surg 1995;112:114–117. 54. Daroff RB, Hoyt WF, Sanders MD, et al. Gaze-evoked eyelid and ocular nystag- 21. Tho¨mke F, Hopf HC. Pontine lesions mimicking acute peripheral vestibulopathy. mus inhibited by the near reflex: unusual ocular motor phenomena in a lateral J Neurol Neurosurg Psychiatry 1999;66:340–349. medullary syndrome. J Neurol Neurosurg Psychiatry 1968;31:362–367. 22. Meienberg O, Rover J, Kommerell G. Prenuclear paresis of homolateral inferior 55. Dieterich M, Brandt T. Wallenberg’s syndrome: lateropulsion, cyclorotation, and rectus and contralateral superior oblique eye muscles. Arch Neurol 1978;35: subjective visual vertical in thirty-six patients. Ann Neurol 1992;31:399–408. 231–233. 56. Brandt T, Dieterich M. Pathological eye-head coordination in roll: tonic ocular 23. Lawden MC, Bronstein AM, Kennard C. Repetitive paroxysmal nystagmus and tilt reaction in mesencephalic and medullary lesions. Brain 1987;110:649–666. vertigo. Neurology 1995;45:276–280. 57. Estanol B, Lopez-Rios G. Neuro-otology of the lateral medullary infarct syn- 24. Bu¨ttner U, Ott M, Helmchen C, et al. Bilateral loss of eighth nerve function as drome. Arch Neurol 1982;39:176–179. the only clinical sign of vertebrobasilar dolichoectasia. J Vestib Res 1995;5: 58. Waespe W, Wichmann W. Oculomotor disturbances during visual-vestibular 47–51. interaction in Wallenberg’s lateral medullary syndrome verified by magnetic 25. Passero S, Nuti D. Auditory and vestibular findings in patients with vertebrobasi- resonance imaging. Brain 1990;113:821–846. lar dolichoectasia. Acta Neurol Scand 1996;93:50–55. 59. Waespe W. Saccadic gain adaptivity in the two eyes in Wallenberg’s lateral 26. Brackmann DE, Kesser BW, Day JD. Microvascular decompression of the ves- medullary syndrome. Neuroophthalmology 1995;15:193–201. tibulocochlear nerve for disabling positional vertigo: the House Ear Clinic expe- 60. Robinson FR, Straube A, Fuch AF. Role of the caudal fastigial nucleus in saccade rience. Otol Neurotol 2001;22:882–887. generation. II. Effects of muscimol inactivation. J Neurophysiol 1993;70: 27. Brandt T, Dieterich M. Vestibular paroxysmia (disabling positional vertigo). 1741–1758. Neuroophthalmology 1994;14:359–369. 61. Grad A, Baloh RW. Vertigo of vascular origin. Clinical and electronystagmo- 28. Brandt T, Dieterich M. Central vestibular syndromes in roll, pitch and yaw. graphic features in 84 cases. Arch Neurol 1989;46:281–284. Neuroophthalmology 1995;15:291–303. 62. Oas J, Baloh RW. Vertigo and the anterior inferior cerebellar artery syndrome. 29. Bu¨ttner U, Helmchen C, Bu¨ttner-Ennever JA. The localizing value of nystagmus Neurology 1992;42:2274–2279. in brainstem disorders. Neuroophthalmology 1995;15:283–290. 63. Lee H, Sohn SI, Jung DK, et al. Sudden deafness and anterior inferior cerebellar 30. Dieterich M. The topographic diagnosis of acquired nystagmus in brainstem artery infarction. Stroke 2002;33:2807–2812. disorders. Strabismus 2002;10:137–145. 64. Keane JR. Alternating skew deviation: 47 patients. Neurology 1985;35:725–728. 31. Fisher CM, Karnes WE, Kubik CS. Lateral medullary infarction: the pattern of 65. Rousseaux M, Petit H, Dubois F, et al. ‘‘Laterally alternating’’ skew deviation. vascular occlusion. J Neuropathol Exp Neurol 1961;20:323–378. Mechanism and significance. Neuroophthalmology 1985;5:277–280. 32. Fisher CM, Tapia J. Lateral medullary infarction extending to the lower pons. 66. Moster ML, Schatz NJ, Savino PJ, et al. Alternating skew on lateral gaze (bilat- J Neurol Neurosurg Psychiatry 1987;50:620–624. eral abducting hypertropia). Ann Neurol 1988;23:190–192. 33. Menendez-Gonzalez M, Garcia C, Suarez E, et al. [Wallenberg’s syndrome 67. Zee DS. Considerations on the mechanisms of alternating skew deviation in secondary to dissection of the vertebral artery caused by chiropractic manipula- patients with cerebellar lesions. J Vest Res 1996;6:1–7. tion]. Rev Neurol 2003;37:837–839. 68. Donahue SP, Lavin PJ, Hamed LM. Tonic ocular tilt reaction simulating a supe- 34. Smith DB, Demasters BKK. Demyelinative disease presenting as Wallenberg’s rior oblique palsy: diagnostic confusion with the 3-step test. Arch Ophthalmol syndrome. Report of a patient. Stroke 1981;12:877–878. 1999;117:347–352. 35. Hornsten G. Wallenberg’s syndrome. I. General symptomatology, with special 69. Donahue SP, Lavin PJ, Mohney B, Hamed L. Skew deviation and inferior reference to visual disturbances and imbalance. Acta Neurol Scand 1974;50: oblique palsy. Am J Ophthalmol 2001;132:751–756. 434–446. 70. Rabinovitch HE, Sharpe JA, Sylvester TO. The ocular tilt reaction. A paroxysmal 36. Tiliket C, Ventre J, Vighetto A, et al. Room tilt illusion: a central otolith dysfunc- dyskinesia associated with elliptical nystagmus. Arch Ophthalmol 1977;95: tion. Arch Neurol 1996;53:1259–1264. 1395–1398. 37. Slavin ML, LoPinto RJ. Isolated environmental tilt associated with lateral medul- 71. Hedges TR III, Hoyt WF. Ocular tilt reaction due to an upper brainstem lesion: lary compression by dolichoectasia of the vertebral artery. J Clin Neuroophthal- paroxysmal skew deviation, torsion, and oscillation of the eyes with head tilt. mol 1987;7:29–33. Ann Neurol 1982;11:537–540. 38. Mehler MF. Complete visual inversion in vertebrobasilar ischaemic disease. J 72. Dieterich M, Brandt T. Ocular torsion and tilt of subjective visual vertical are Neurol Neurosurg Psychiatry 1988;51:1236–1237. sensitive brainstem signs. Ann Neurol 1993;33:292–299. 39. Charles N, Froment C, Rode G, et al. Vertigo and upside down vision due to 73. Brandt T, Dieterich M. Vestibular syndromes in the roll plane: topographic an infarct in the territory of the medial branch of the posterior inferior cerebellar diagnosis from brain stem to cortex. Ann Neurol 1994;36:337–347. artery caused by dissection of a vertebral artery. J Neurol Neurosurg Psychiatry 74. Galetta SL, Liu GT, Raps EC, et al. Cyclodeviation in skew deviation. Am J 1992;55:188–189. Ophthalmol 1994;118:509–514. 40. Lo¨pez L, Ochoa S, Mesropian H, et al. Acute transient upside-down inversion 75. Brodsky MC. Three dimensions of skew deviation. Br J Ophthalmol 2003;87: of vision with brainstem-cerebellar infarction. Neuroophthalmology 1995;15: 1440–1441. 277–280. 76. Brandt T, Dieterich M. Two types of ocular tilt reactions: the ‘‘ascending’’ 41. Aldridge AJ, Kline LB, Girkin CA. Environmental tilt illusion as the only symp- pontomedullary VOR-OTR and the ‘‘descending’’ mesencephalic integrator- tom of a thalamic astrocytoma. J Neuroophthalmol 2003;23:145–147. OTR. Neuroophthalmology 1998;19:83–92. 42. Solms M, Kaplan-Solms K, Saling M, et al. Inverted vision after frontal lobe 77. Cogan DG. Neurology of the Ocular Muscles, ed 2. Springfield, IL, Charles C disease. Cortex 1988;24:499–509. Thomas, 1956. SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 955

78. Keane JR. Ocular skew deviation. Analysis of 100 cases. Arch Neurol 1975; 113. Halmagyi GM, Pamphlett R, Curthoys IS. Seesaw nystagmus and ocular tilt 32:185–190. reaction due to adult Leigh’s disease. Neuroophthalmology 1992;12:1–9. 79. Brandt T, Dieterich M. Skew deviation with ocular torsion: a vestibular sign of 114. Corbett JJ, Schatz NJ, Shults WT, et al. Slowly alternating skew deviation: topographic diagnostic value. Ann Neurol 1993;33:528–534. description of a pretectal syndrome in three patients. Ann Neurol 1981;10: 80. Riordan-Eva P, Harcourt JP, Faldon M, et al. Skew deviation following vestibu- 540–546. lar nerve surgery. Ann Neurol 1997;41:94–99. 115. Mitchell JM, Smith CL, Quencer RB. Periodic alternating skew deviation. J Clin 81. Arbusow V, Dieterich M, Strupp M, et al. Herpes zoster neuritis involving Neuroophthalmol 1981;1:5–8. superior and inferior parts of the vestibular nerve causes ocular tilt reaction. 116. Greenberg HS, DeWitt LD. Periodic nonalternating ocular skew deviation ac- Neuroophthalmology 1998;19:17–22. companied by head tilt and pathologic lid retraction. J Clin Neuroophthalmol 82. Verhulst E, Van Lammeren M, Dralands L. Diplopia from skew deviation in 1983;3:181–184. Ramsey-Hunt syndrome. A case report. Bull Soc Belge Ophtalmol 2000;27–32. 117. Lewis JM, Kline LB. Periodic alternating nystagmus associated with periodic 83. Tilikete C, Vighetto A. Internuclear ophthalmoplegia with skew deviation. Two alternating skew deviation. J Clin Neuroophthalmol 1983;13:115–117. cases with an isolated circumscribed lesion of the medial longitudinal fasciculus. 118. Westheimer G, Blair SM. The ocular tilt reaction. A brainstem oculomotor rou- Eur Neurol 2000;44:258–259. tine. Invest Ophthalmol 1975;14:833–839. 84. Merikangas JR. Skew deviation in pseudotumor cerebri. Ann Neurol 1978;4: 119. Westheimer G, Blair SM. Synkinese der Augenund Kopfbewegungen bei Hirns- 583. tammreizungen am Wachen Macaws-Affen. Exp Brain Res 1975;24:89–95. 85. Frohman LP, Kupersmith MJ. Reversible vertical ocular deviations associated 120. Lueck CJ, Hamlyn P, Crawford TJ, et al. A case of ocular tilt reaction and with raised intracranial pressure. J Clin Neuroophthalmol 1985;5:158–163. torsional nystagmus due to direct stimulation of the midbrain in man. Brain 86. Hoyt CS, Mousel DK, Weber AA Transient supranuclear disturbances of gaze 1991;114:2069–2079. in healthy neonates. Am J Ophthalmol 1980;89:708–713. 121. Straube A, Bu¨ttner U. Pathophysiology of saccadic contrapulsion in unilateral 87. Glasauer S, Dieterich M, Brandt T. Simulation of pathological ocular counter- rostral cerebellar lesions. Neuroophthalmology 1994;14:3–7. roll and skew-torsion by a 3-D mathematical model. Neuroreport 1999;10: 122. Galimberti CA, Versino M, Sartori I, et al. Epileptic skew deviation. Neurology 1843–1848. 1988;50:1469–1472. 88. Averbuch-Heller L, Rottach KG, Zivotofsky AZ, et al. Torsional eye movements 123. Holmes G. The symptoms of acute cerebellar injuries due to gunshot injuries. in patients with skew deviation and spasmodic torticollis: responses to static Brain 1917;40:461–535. and dynamic head roll. Neurology 1997;48:506–514. 124. Daroff RB. Eye signs in humans with cerebellar dysfunction. In: Lennerstrand 89. Collewijn H, Van der Steen J, Ferman L, et al. Human ocular counterroll: assess- G, Zee DS, Keller EL, eds. Functional Basis of Ocular Motility Disorders. Ox- ment of static and dynamic properties from electromagnetic scleral coil record- ford, UK, Pergamon, 1982:463–465. ings. Exp Brain Res 1985;59:185–196. 125. Zee DS, Walker MF. Cerebellar control of eye movements. In: Chalupa LM, 90. Pansell T, Schworm HD, Ygge J. Torsional and vertical eye movements during Werner JS, eds. The Visual Neurosciences. Cambridge, MA, MIT Press, 2003: head tilt dynamic characteristics. Invest Ophthalmol Vis Sci 2003;44: 1485–1498. 2986–2990. 126. Takagi M, Tamargo R, Zee DS. Effects of lesions of the cerebellar oculomotor 91. Kushner BJ. Ocular torsion: rotations around the ‘‘WHY’’ axis. J AAPOS 2004; vermis on eye movements in primate: binocular control. Prog Brain Res 2003; 8:1–12. 142:19–33. 92. Jauregui-Renaud K, Faldon M, Clarke AH, et al. Otolith and semicircular canal 127. Takagi M, Zee DS, Tamargo R. Effects of lesions of the oculomotor vermis on contributions to the human binocular response to roll oscillation. Acta Otolaryn- eye movements in primate: saccades. J Neurophysiol 1998;80:1911–1930. gol 1998;118:170–176. 128. Takagi M, Zee DS, Tamargo R. Effects of lesions of the oculomotor cerebellar 93. Halmagyi GM, Gresty MA, Gibson WPR. Ocular tilt reaction with peripheral vermis on eye movements in primate: smooth pursuit. J Neurophysiol 2000;83: vestibular lesion. Ann Neurol 1979;6:80–83. 247–262. 94. Deecke L, Mergner T, Plester D. Tullio phenomenon with torsion of the eyes 129. Gaymard B, Rivaud S, Amarenco P, et al. Influence of visual information on and subjective tilt of the visual surround. Ann NY Acad Sci 1981;374:650–655. cerebellar saccadic dysmetria. Ann Neurol 1994;35:108–112. 95. Dieterich M, Brandt T, Fries W. Otolithic function in man. Results from a case 130. Kanayama R, Bronstein AM, Shallo-Hoffmann J, et al. Visually and memory of otolithic Tullio phenomenon. Brain 1989;112:1377–1392. guided saccades in a case of cerebellar saccadic dysmetria. J Neurol Neurosurg 96. Minor LB, Solomon D, Zinreich JS, Zee DS. Sound- and/or pressure-induced Psychiatry 1994;57:1081–1084. vertigo due to bone dehiscence of the superior semicircular canal. Arch Otolaryn- 131. Bronstein AM, Shallo-Hoffmann J, Kanayama R, et al. Corrective saccades in gol Head Neck Surg 1998;124:249–258. cerebellar dysmetria. Ann Neurol 1995;37:413–414. 97. Minor LB. Labyrinthine fistulae: pathobiology and management. Curr Opin Oto- 132. Bo¨tzel K, Rottach K, Bu¨ttner U. Normal and pathological saccadic dysmetria. laryngol Head Neck Surg 2003;11:340–346. Brain 993;116:337–353. 98. Baloh RW. Superior semicircular canal dehiscence syndrome: leaks and squeaks 133. Straube A, Deubel H, Spuler A, et al. Differential effect of bilateral deep cerebel- can make you dizzy. Neurology 2004;62:684–685. lar nuclei lesion on externally and internally triggered saccades in humans. Neu- 99. Curthoys IS. Eye movements produced by utricular and saccular stimulation. roophthalmology 1995;15:67–74. Aviat Space Environ Med 1987;58(Suppl):A192–A197. 134. Nawrot M, Rizzo M. Motion perception deficits from midline cerebellar lesions 100. Suzuki JI, Tokumasu K, Goto K. Eye movements from single utricular nerve in human. Vision Res 1995;35:723–731. stimulation. Acta Otolaryngol (Stockh) 1969;68:350–362. 135. Kurzan R, Straube A, Bu¨ttner U. The effect of muscimol micro-injections into 101. Gacek RR. Neuroanatomical correlates of vestibular function. Ann Otol Rhinol the fastigial nucleus on the optokinetic response and the vestibulo-ocular reflex Laryngol 1980;89:2–5. in the alert monkey. Exp Brain Res 1993;94:252–260. 102. Uchino Y. Otolith and semicircular canal inputs to single vestibular neurons in 136. Bu¨ttner U, Straube A, Spuler A. Saccadic dysmetia and ‘‘intact’’ smooth-pursuit cats. Biol Sci Space 2001;15:375–381. eye movements after bilateral deep cerebellar nuclei lesion. J Neurol Neurosurg 103. Brandt T, Dieterich M. Different types of skew deviation. J Neurol Neurosurg Psychiatry 1994;57:832–834. Psychiatry 1991;54:549–550. 137. Westheimer G, Blair SM. Functional organization of primate oculomotor system 104. Versino M, Hurko O, Zee DS. Disorders of binocular control of eye movements revealed by cerebellectomy. Exp Brain Res 1974;21:463–472. in patients with cerebellar dysfunction. Brain 1996;119:1933–1950. 138. Rambold H, Churchland A, Selig Y, et al. Partial ablations of the flocculus and 105. Mossman S, Halmagyi GM. Partial ocular tilt reaction due to unilateral cerebellar ventral paraflocculus in monkeys cause linked deficits in smooth pursuit eye lesion. Neurology 1997;49:491–493. movements and adaptive modification of the VOR. J Neurophysiol 2002;87: 106. Radtke A, Bronstein AM, Gresty MA, et al. Paroxysmal alternating skew devia- 912–924. tion and nystagmus after partial destruction of the uvula. J Neurol Neurosurg 139. Bu¨ttner U, Grundei T. Gaze-evoked nystagmus and smooth-pursuit deficits: their Psychiatry 2001;70:790–793. relationship studied in 52 patients. J Neurol 1995;242:384–389. 107. Brandt T, Dieterich M. Vestibular syndromes in the roll plane: topographic 140. Grant MP, Leigh RJ, Seidman SH, et al. Comparison of predictable smooth diagnosis from brainstem to cortex. Ann Neurol 1994;36:337–347. ocular and combined eye-head tracking behaviour in patients with lesions affect- 108. Nashold BS Jr, Gills JP Jr. Ocular signs from brain stimulation and lesions. ing the brainstem and cerebellum. Brain 1992;115:1323–1342. Arch Ophthalmol 1967;77:609–618. 141. Waterston JA, Barnes GR, Grealy MA. A quantitative study of eye and head 109. Keane JR. Pretectal pseudobobbing. Five patients with ‘‘V’’-pattern conver- movements during smooth pursuit in patients with cerebellar disease. Brain 1992; gence nystagmus. Arch Neurol 1985;42:592–594. 115:1343–1358. 110. Halmagyi GM, Brandt T, Dieterich M, et al. Tonic contraversive ocular tilt 142. Lekwuwa GU, Barnes GR, Grealy MA. Effects of prediction on smooth-pursuit reaction due to unilateral meso-diencephalic lesion. Neurology 1990;40: eye velocity gain in cerebellar patients and controls. In: Findlay JM, Walker 1503–1509. R, Kentridge RW, eds. Eye Movement Research: Mechanisms, Processes, and 111. Dieterich M, Brandt T. Ocular torsion and perceived vertical in oculomotor, Applications. Amsterdam, Elsevier, 1995:119–129. trochlear and abducens palsies. Brain 1993;116:1095–1104. 143. Moschner C, Zangemeister WH, Demer JL. Anticipatory smooth eye movements 112. Dichgans M, Dieterich M. Third nerve palsy with contralateral ocular torsion of high velocity triggered by large target steps: normal performance and effect and binocular tilt of visual vertical, indicating a midbrain lesion. Neuroophthal- of cerebellar degeneration. Vision Res 1996;36:1341–1348. mology 1995;15:315–320. 144. Cohen B, John P, Yakushin SB, et al. The nodulus and uvula: source of cerebellar 956 CLINICAL NEURO-OPHTHALMOLOGY

control of spatial orientation of the angular vestibulo-ocular reflex. Ann NY nystagmus: Another variant of benign positional nystagmus? Neurology 1995; Acad Sci 2002;978:28–45. 45:1297–1301. 145. Hain TC, Zee DS, Maria BL. Tilt suppression of vestibulo-ocular refiex in 180. Strupp M, Brandt T, Steddin S. Horizontal canal benign paroxysmal positioning patients with cerebellar lesions. Acta Otolaryngol (Stockh) 1988;105:13–20. vertigo: reversible ipsilateral caloric hypoexcitability caused by canalolithiasis. 146. Wiest G, Deecke L, Trattnig S, Mueller C. Abolished tilt suppression of the Neurology 1995;45:2072–2076. vestibulo-ocular reflex caused by a selective uvulo-nodular lesion. Neurology 181. Nuti D, Vannucchi P, Pagnini P. Benign paroxysmal positional vertigo of the 1999;52:417–419. horizontal canal: a form of canalolithiasis with variable clinical features. J Vest 147. Walker MF, Steffen H, Zee DS Three-axis approaches to ocular motor control: Res 1996;6:173–184. a role for the cerebellum. In: Harris LR, Jenkin M, eds. Levels of Perception. 182. Leigh RJ, Mapstone T, Weymann C. Eye movements in children with the Dandy- New York, Springer-Verlag, 2002:399–413. Walker syndrome. Neuroophthalmology 1992;12:285–288. 148. Zee DS. Cerebellar control of eye movements. In: Honrubia V, Brazier MAB, 183. Calogero JA. Vermian agenesis and unsegmented midbrain tectum. Case report. eds. Nystagmus and Vertigo: Clinical Approaches to the Patient with Dizziness. J Neurosurg 1977;47:605–608. New York, Academic Press, 1982:241–249. 184. Sarnat HB, Alcala H. Human : a syndrome of diverse 149. FitzGibbon E, Calvert P, Zee DS, et al. Torsional nystagmus during vertical causes. Arch Neurol 1980;37:300–305. smooth pursuit. J Neuroophthal 1996;16:79–90. 185. DeBassio WA, Kemper TL, Knoefel JE. Coffin-Siris syndrome. Neuropatho- 150. Angelaki DE, Dickman JD. Premotor neurons encode torsional eye velocity logic findings. Arch Neurol 1985;42:350–353. during smooth-pursuit eye movements. J Neurosci 2003;23:2971–2979. 186. Lambert SR, Kriss A, Gresty M, et al. Joubert syndrome. Arch Ophthalmol 151. Zee DS, Yee RD, Cogan DG, et al. Ocular motor abnormalities in hereditary 1989;107:709–713. cerebellar ataxia. Brain 1976;99:207–234. 187. Boltshauser E. Joubert syndrome: more than lower cerebellar vermis hypoplasia, 152. Hotson JR. Cerebellar control of fixation eye movements. Neurology 1982;32: less than a complex brain malformation. Am J Med Genet 2002;109:332. 31–36. 188. Lagier-Tourenne C, Boltshauser E, Breivik N, et al. Homozygosity mapping of 153. Walker MF, Zee DS. Directional abnormalities of vestibular and optokinetic a third Joubert syndrome locus to 6q23. J Med Genet 2004;41:273–277. responses in cerebellar disease. Ann NY Acad Sci 1999;871:205–220. 189. Margolis RL. The spinocerebellar ataxias: order emerges from chaos. Curr Neu- 154. Zee DS, Walker MF, Ramat S. The cerebellar contribution to eye movements rol Neurosci Rep 2002;2:447–456. based upon lesions: binocular, three-axis control and the translational vestibulo- 190. Di Donato S, Gellera C, Mariotti C. The complex clinical and genetic classifica- ocular reflex. Ann NY Acad Sci 2002. tion of inherited ataxias. II. Autosomal recessive ataxias. Neurol Sci 2001;22: 155. Yee RD, Baloh RW, Honrubia V, et al. Slow build-up of optokinetic nystagmus 219–228. associated with downbeat nystagmus. Invest Ophthalmol Vis Sci 1979;18: 191. Di Donato S. The complex clinical and genetic classification of inherited ataxias. 622–629. I. Dominant ataxias. Ital J Neurol Sci 1998;19:335–343. 156. Leech J, Gresty M, Hess K, et al. Gaze failure, drifting eye movements, and 192. Rivaud-Pechoux S, Gaymard B, Canel G, et al. Eye movement abnormalities centripetal nystagmus in cerebellar disease. Br J Ophthalmol 1977;61:774–781. correlate with genotype in autosomal dominant cerebellar ataxia type 1. Ann 157. Daroff RB, Troost BT. Upbeat nystagmus. JAMA 973;225:312. Neurol 1998;43:297–302. 158. Bertholon P, Bronstein AM, Davies RA, et al. Positional down-beating nystag- 193. Buttner N, Geschwind D, Jen JC, et al. Oculomotor phenotypes in autosomal mus in 50 patients: cerebellar disorders and possible anterior semicircular canali- dominant ataxias. Arch Neurol 1998;55:1353–1357. thiasis. J Neurol Neurosurg Psychiatry 2002;72:366–372. 194. Durig JS, Jen JC, Demer JL. Ocular motility in genetically defined autosomal 159. Bronstein AM, Hood JD. Cervical nystagmus due to loss of cerebellar inhibition dominant cerebellar ataxia. Am J Ophthalmol 2002;133:718–721. on the cervicoocular reflex: a case report. J Neurol Neurosurg Psychiatry 1985; 195. Schols l, Bauer P, Schmidt T, et al. Autosomal dominant cerebellar ataxias: 48:128–131. clinical features, genetics, and pathogensis. Lancet Neurol 2004;3:291–304. 160. Wiest G, Tian JR, Baloh RW, et al. Otolith function in cerebellar ataxia due to 196. Swartz BE, Li S, Bespalova I, et al. Pathogenesis of clinical signs in recessive mutations in the calcium channel gene CACNA1A. Brain 2001;124:2407–2416. ataxia with saccadic intrusions. Ann Neurol 2003;54:824–828. 161. Wiest G, Tian JR, Baloh RW, et al. Initiation of the linear vestibulo-ocular reflex 197. Fetter M, Klockgether T, Schulz JB, et al. Oculomotor abnormalities and MRI in cerebellar dysfunction. Ann NY Acad Sci 2001;942:505–507. findings in idiopathic cerebellar ataxia. J Neurol 1994;241:234–241. 162. Ito M. Mechanisms of motor learning in the cerebellum. Brain Res 2000;886: 198. Bataller L, Dalmau J. Paraneoplastic neurologic syndromes. Neurol Clin 2003; 237–245. 21:221–247. 163. Yagi T, Shimizu M, Sekine S, et al. New neurotological test for detecting cerebel- 199. Darnell RB, Posner JB. Paraneoplastic syndromes involving the nervous system. lar dysfunction. Vestibulo-ocular reflex changes with horizontal vision-reversal NEngl J Med 2003;349:1543–1554. prisms. Ann Otorhinolaryngol 1981;90:276–280. 200. Jen J, Kim GW, Baloh RW. Clinical spectrum of episodic ataxia type 2. Neurol- 164. Milder DG, Reinecke RD. Phoria adaptation to prisms. A cerebellar–dependent ogy 2004;62:17–22. response. Arch Neurol 1983;40:339–342. 201. Jen J. Familial episodic ataxias and related ion channel disorders. Curr Treat 165. Hain TC, Luebke A. Phoria adaptation in patients with cerebellar lesions. Invest Options Neurol 2000;2:429–431. Ophthalmol Vis Sci 1990;31:1394–1397. 202. Subramony SH, Schott K, Raike RS, et al. Novel CACNA1A mutation causes 166. Wania JH, Walsh FB. Absence of ocular signs with cerebellar ablation in an febrile episodic ataxia with interictal cerebellar deficits. Ann Neurol 2003;54: infant. Arch Ophthalmol 1959;61:655–656. 725–731. 167. Eckmiller R, Westheimer G. Compensation of oculomotor deficits in monkeys 203. Kullmann DM. The neuronal channelopathies. Brain 2002;125:1177–1195. with neonatal cerebellar ablations. Exp Brain Res 1983;49:315–326. 204. Pietrobon D. Calcium channels and channelopathies of the central nervous sys- 168. Cheng JS, Nash J, Meyer GA. Chiari type I malformation revisited: diagnosis tem. Mol Neurobiol 2002;25:31–50. and treatment. Neurologist 2002;8:357–362. 205. Lee H, Yi HA, Cho YW, Sohn CH, et al. Nodulus infarction mimicking acute 169. Schijman E. History, anatomic forms, and pathogenesis of Chiari I malforma- peripheral vestibulopathy. Neurology 2003;60:1700–1702. tions. Childs Nerv Syst 2004;2–13. 206. Amarenco P, Hauw J-J. Cerebellar infarction in the territory of the anterior and 170. Zee DS, Yamazaki A, Butler PH, et al. Effects of ablation of flocculus and inferior cerebellar artery. A clinicopathological study of 20 cases. Brain 1990; paraflocculus on eye movements in primate. J Neurophysiol 1981;46:878–899. 113:139–155. 171. Spooner JW, Baloh RW. Arnold-Chiari malformation. Improvement in eye 207. Kase CS, White JL, Joslyn JN, et al. Cerebellar infarction in the superior cerebel- movements after surgical treatment. Brain 1981;104:51–60. lar artery distribution. Neurology 1985;35:705–711. 172. Spillane JD, Pallis C, Jones AM. Developmental abnormalities in the region of 208. Ranalli PJ, Sharpe JA. Contrapulsion of saccades and ipsilateral ataxia: a unilat- the foramen magnum. Brain 1957;80:11–48. eral disorder of the rostral cerebellum. Ann Neurol 1986;20:311–316. 173. Gregorius FK, Crandall PH, Baloh RW. Positional vertigo with cerebellar astro- 209. Benjamin EE, Zimmerman CF, Troost BT. Lateropulsion and upbeat nystagmus cytoma. Surg Neurol 1976;6:283–286. are manifestations of central vestibular dysfunction. Arch Neurol 1986;43: 174. Watson P, Barber HO, Deck J, et al. Positional vertigo and nystagmus of central 962–964. origin. Can J Neurol Sci 1981;8:133–137. 210. Uno A, Mukuno K, Sekiya H, et al. Lateropulsion in Wallenberg’s syndrome and 175. Kattah JC, Kolsky MP, Luessenhop AJ. Positional vertigo and the cerebellar contrapulsion in the proximal type of the superior cerebellar artery syndrome. vermis. Neurology 1984;34:527–529. Neuroophthalmology 1989;9:75–80. 176. Baloh RW, Honrubia V, Jacobson K. Benign positional vertigo: clinical and 211. Waespe W, Mu¨ller-Meisser E. Directional reversal of saccadic dysmetria and oculographic features in 240 cases. Neurology 1987;37:371–378. gain adaptivity in a patient with a superior cerebellar artery infarction. Neurooph- 177. Sakata E, Ohtsu K, Shimura H, et al. Positional nystagmus of benign paroxysmal thalmology 1996;16:65–74. type (BPPN) due to cerebellar vermis lesions. Pseudo–BPPN. Auris Masus Lar- 212. Vahedi K, Rivaud S, Amarenco P, et al. Horizontal eye movement disorders ynx (Tokyo) 1987;14:17–21. after posterior vermis infarctions. J Neurol Neurosurg Psychiatry 1995;58: 178. Brandt T, Steddin S. Current view of the mechanism of benign paroxysmal 91–94. positioning vertigo: cupulolithiasis or canalolithiasis. J Vestibular Res 1993;3: 213. Ekbom K, Horsten G, Johansson T. Posterior cranial fossa tumors. Headaches, 373–382. oculostatic disorders and scintillation camera findings. Headache 1974;14: 179. Baloh RW, Yue Q, Jacobson BA, et al. Persistent direction-changing positional 119–132. SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 957

214. Yamazaki A, Zee DS. Rebound nystagmus: EOG analysis of a case with a 247. Guitton D, Kirkham T. Vertical oculomotor deficits following MLF lesions in floccular tumor. Br J Ophthalmol 1979;63:782–786. man and monkey are similar. Soc Neurosci Abst 1978;4:163. 215. Humphriss RL, Baguley DM, Moffat DA. Head-shaking nystagmus in patients 248. Gordon RM, Bender MB. Visual phenomena in lesions of the medial longitudinal with a vestibular schwannoma. Clin Otolaryngol 2003;28:514–519. fasciculus. Arch Neurol 1966;15:238–240. 216. Ott KH, Kase CS, Ojemann RG, et al. Cerebellar hemorrhage: diagnosis and 249. Gresty MA, Hess K, Leech J. Disorders of the vestibulo-ocular reflex producing treatment. Arch Neurol 1974;31:160–167. oscillopsia and mechanisms compensating for loss of labyrinthine function. 217. Brennen RW, Bergland RM. Acute cerebellar hemorrhage. Analysis of clinical Brain 1977;100:693–716. findings and outcome in 12 cases. Neurology 1977;27:527–532. 250. Frohman EM, Solomon D, Zee DS. Nuclear, supranuclear and internuclear eye 218. Dunne JW, Chakera T, Kermode S. Cerebellar haemorrhage: diagnosis and treat- movement abnormalities in multiple sclerosis. Int J MS 1996;2:79–89. ment: a study of 75 consecutive cases. Q J Med 1987;64:739–754. 250a. Kim JS. Internuclear ophthalmoplegia as an isolated or predominant symptom 219. Cogan DG. Internuclear ophthalmoplegia: typical and atypical. Arch Ophthalmol of brainstem infarction. Neurology 2004;62:1491–1496. 1970;84:583–589. 251. Mutschler V, Eber AM, Rumbach L, et al. Internuclear ophthalmoplegia in 14 220. Zee DS. Internuclear ophthalmoplegia: clinical and pathophysiological consider- patients. Clinical and topographic correlation using magnetic resonance imaging. ations. In: Bu¨ttner U, Brandt T, eds. Oculomotor Disorders in the Brain Stem. Neuroophthalmology 1990;10:319–325. London, WB Saunders, 1992:455–470. 252. Schmidt F, Kastrup A, Nagele T, et al. Isolated ischemic internuclear ophthal- 221. Smith JL, David NJ. Internuclear ophthalmoplegia: two new clinical signs. Neu- moplegia: toward the resolution limits of DW-MRI. Eur J Neurol 2004;11: rology 1964;14:307–309. 67–68. 222. Ventre J, Vighetto A, Bailly G, et al. Saccade metrics in multiple sclerosis: 253. Lutz A. Ueber die Bahnen der Blickwendung und deren Dissoziierung. Klin versional velocity disconjugacy as the best clue? J Neurol Sci 1991;102: Monatsble Augenheilkd 1923;70:213–235. 144–149. 254. Kommerell G. Internuclear ophthalmoplegia of abduction. Isolated impairment 223. Flipse JP, Straathof CSM, Van der Steen J, et al. Binocular saccadic acceleration of phasic ocular motor activity in supranuclear lesions. Arch Ophthalmol 1975; in multiple sclerosis. Neuroophthalmology 1996;16:43–46. 93:531–534. 224. Frohman EM, Frohman TC, O’Suilleabhain P, et al. Quantitative oculographic 255. Cremer PD, Halmagyi GM. Eye movement disorders in a patient with liver characterisation of internuclear in multiple sclerosis: the ver- disease. Neuroophthalmology 1996;16(Suppl):276. sional dysconjugacy index Z score. J Neurol Neurosurg Psychiatry 2002;73: 256. Oliveri RL, Bono F, Quattrone A. Pontine lesion of the abducens fasciculus 51–55. producing so-called posterior internuclear ophthalmoplegia. Eur Neurol 1997; 225. Kupfer C, Cogan DG. Unilateral internuclear ophthalmoplegia: a clinicopatho- 37:67–69. logical case report. Arch Ophthalmol 1966;75:484–489. 257. Bogousslavsky J, Regli F, Ostinelli B, et al. Paresis of lateral gaze alternating 226. Feldon SE, Hoyt WE, Stark L. Disordered inhibition in internuclear ophthal- with so-called posterior internuclear ophthalmoplegia. A partial paramedian pon- moplegia. Analysis of eye movement recordings with computer simulations. tine reticular formation-abducens nucleus syndrome. J Neurol 1985;232:38–42. Brain 1980;103:113–137. 258. Tho¨mke F, Hopf HC, Kra¨mer G. Internuclear ophthalmoplegia of abduction: clinical and electrophysiological data on the existence of an abduction paresis 227. Tho¨mke F, Hopf HC. Abduction paresis with rostral pontine and/or mesence- phalic lesions: pseudoabducens palsy and its relation to the so-called posterior of prenuclear origin. J Neurol Neurosurg Psychiatry 1992;55:105–111. ¨ internuclear ophthalmoplegia of Lutz. BMC Neurol 2001;1:4. 259. C¸ elebisoy N, Akyu¨rekli O. One-and-a-half syndrome, type II: a case with rostral 228. Kommerell G. Unilateral internuclear ophthalmoplegia. The lack of inhibitory brain stem infarction. Neuroophthalmology 1996;16:373–377. 260. Wiest G, Wanschitz J, Baumgartner C, et al. So-called posterior internuclear involvement in medial rectus muscle activity. Invest Ophthalmol Vis Sci 1981; ophthalmoplegia due to a pontine glioma: a clinicopathological study. J Neurol 21:592–599. 1999;246:412–415. 229. Bronstein AM, Rudge P, Gresty MA, et al. Abnormalities of horizontal gaze. 261. Carpenter MB, McMasters RE, Hanna GR. Disturbances of conjugate horizontal Clinical, oculographic and magnetic resonance imaging findings. II. Gaze palsy eye movements in the monkey. I. Physiological effects and anatomical degenera- and internuclear ophthalmoplegia. J Neurol Neurosurg Psychiatry 1990;53: tion resulting from lesions of the abducens nucleus and nerve. Arch Neurol 200–207. 1963;8:231–247. 230. Thomke F, Hopf HC, Breen LA. Slowed abduction saccades in bilateral in- ¨ 262. Hirose G, Furui K, Yoshioka A, et al. Unilateral conjugate gaze palsy due to a ternuclear ophthalmoplegia. Neuroophthalmology 1992;12:241–246. lesion of the abducens nucleus. J Clin Neuroophthalmol 1993;13:54–58. 231. Crane TB, Yee RD, Baloh RW, et al. Analysis of characteristic eye movement 263. Mu¨ri RM, Chermann JF, Cohen L, et al. Ocular motor consequences of damage abnormalities in internuclear ophthalmoplegia. Arch Ophthalmol 1983;101: to the abducens nucleus area in humans. J Neuroophthalmol 1996;16:191–195. 206–210. 264. Miller NR, Biousse V, Hwang T, et al. Isolated acquired unilateral horizontal 232. Pola J, Robinson DA. An explanation of eye movements seen in internuclear gaze paresis from a putative lesion of the abducens nucleus. J Neuroophthalmol ophthalmoplegia. Arch Neurol 1976;33:447–452. 2002;22:204–207. 233. Pierrot-Deseilligny C, Chain F. L’ophtalmople´gie internucle´aire. Rev Neurol 265. Amaya LG, Walker J, Taylor D. Mo¨bius syndrome: a study and report of 18 (Paris) 1979;135:485–513. cases. Binoc Vis Q 1990;5:119–132. 234. Baloh RW, Yee RD, Honrubia V. Internuclear ophthalmoplegia. I. Saccades 266. Bronstein AM, Morris J, Du Boulay G, et al. Abnormalities of horizontal gaze. and dissociated nystagmus. Arch Neurol 1978;35:484–489. Clinical, culographic and magnetic resonance imaging findings. I. Abducens 235. Zee DS, Hain TC, Carl JR. Abduction nystagmus in internuclear ophthal- palsy. J Neurol Neurosurg Psychiatry 1990;53:194–199. moplegia. Ann Neurol 1987;21:383–388. 267. Goebel HH, Komatsuzaki A, Bender MB, et al. Lesions of the pontine tegmen- 236. Tho¨mke F. Some observations on abduction nystagmus in internuclear ophthal- tum and conjugate gaze paralysis. Arch Neurol 1971;24:431–440. moplegia. Neuroophthalmology 1996;16:27–38. 268. Pierrot-Deseilligny C, Chain F, Lhermitte F. Syndrome de la formation re´ticu- 237. Gamlin PDR, Gnadt JW, Mays LE. Lidocaine-induced unilateral internuclear laire pontique. Pre´cisions physiopathologiques sur les anomalies des mouve- ophthalmoplegia: effects on convergence and conjugate eye movements. J Neu- ments oculaires volontaires. Rev Neurol (Paris) 1982;138:517–532. rophysiol 1989;62:82–95. 269. Johnston JL, Sharpe JA, Ranalli PJ, et al. Oblique misdirection and slowing of 238. Bu¨ttner-Ennever JA, Horn AKE, Schmidtke K. Cell groups of the medial longitu- vertical saccades after unilateral lesions of the pontine tegmentum. Neurology dinal fasciculus and paramedian tracts. Rev Neurol (Paris) 1989;145:533–539. 1993;43:2238–2244. 239. Nozaki S, Mukuno K, Ishikawa S. Internuclear ophthalmoplegia associated with 270. Johnston JL, Sharpe JA, Morrow MJ. Paresis of contralateral smooth pursuit ipsilateral downbeat nystagmus and contralateral incyclorotatory nystagmus. and normal vestibular smooth eye movements after unilateral brainstem lesions. Ophthalmologica 1983;187:210–216. Ann Neurol 1992;31:495–502. 240. Cremer PD, Migliaccio AA, Halmagyi GM, Curthoys IS. Vestibulo-ocular reflex 271. Thier P, Bachor A, Faiss J, et al. Selective impairment of smooth-pursuit eye pathways in internuclear ophthalmoplegia. Ann Neurol 1999;45:529–533. movements due to an ischemic lesion of the basal pons. Ann Neurol 1991;29: 241. Lin HY, Tsai RK. Internuclear ophthalmoplegia associated with transient tor- 443–448. sional nystagmus. J Formos Med Assoc 2002;101:446–448. 272. Gaymard B, Pierrot-Deseilligny C, Rivaud S, et al. Smooth-pursuit eye move- 242. Srivastava AK, Tripathi M, Gaikwad SB, et al. Internuclear ophthalmoplegia ment deficits after pontine nuclei lesions in humans. J Neurol Neurosurg Psychia- and torsional nystagmus: an MRI correlate. Neurol India 2003;51:271–272. try 1993;56:799–807. 243. Herishanu YO, Sharpe JA. Saccadic intrusions in internuclear ophthalmoplegia. 273. Deleu D, Michotte A, Ebinger G. Impairment of smooth pursuit in pontine Ann Neurol 1983;14:67–72. lesions: functional topography based on MRI and neuropathologic findings. Acta 244. Evinger LC, Fuchs AF, Baker R. Bilateral lesions of the medial longitudinal Neurol Belg 1997;97:28–35. fasciculus in monkeys: effects on the horizontal and vertical components of 274. Furman JMR, Hurtt MR, Hirsch WL. Asymmetrical ocular pursuit with posterior voluntary and vestibular induced eye movements. Exp Brain Res 1977;28:1–20. fossa tumors. Ann Neurol 1991;30:208–211. 245. Leigh RJ, Newman SA, King WM. Vertical gaze disorders. In: Lennerstrand 275. Pierrot-Deseilligny C, Goasguen J. Isolated abducens nucleus damage due to G, Zee DS, Keller EL, eds. Functional Basis of Ocular Motility Disorders. Ox- histiocytosis X. Brain 1984;107:1019–1032. ford, UK, Pergamon, 982:257–266. 276. Kommerell G, Henn V, Bach M, et al. Unilateral lesion of the paramedian 246. Kirkham TH, Katsarkas A. An electrooculographic study of internuclear ophthal- pontine reticular formation. Loss of rapid eye movements with preservation of moplegla. Ann Neurol 1977;2:385–392. vestibulo-ocular reflex and pursuit. Neuroophthalmology 1987;7:93–98. 958 CLINICAL NEURO-OPHTHALMOLOGY

277. Monteiro MLR, Coppeto JR. Cryptic disseminated tuberculosis presenting as 310. Zee DS, Robinson DA. A hypothetical explanation of saccadic oscillations. Ann gaze palsy. J Clin Neuroophthalmol 1985;5:27–29. Neurol 1979;5:405–414. 278. Hanson MR, Hamid MA, Tomsak RL, et al. Selective saccadic palsy caused by 311. Averbuch-Heller L, Kori AA, Rottach K, et al. Dysfunction of pontine omni- pontine lesions: clinical, physiological, and pathological correlations. Ann Neu- pause neurons causes impaired fixation: macrosaccadic oscillations with a unilat- rol 1986;20:209–217. eral pontine lesion. Neuroophthalmology 1996;16:99–106. 279. Nishida T, Tychsen L, Corbett JJ. Resolution of saccadic palsy after treatment 312. Bhidayasiri R, Plant GT, Leigh RJ. A hypothetical scheme for the brainstem of brain-stem metastasis. Arch Neurol 1986;43:1196–1197. control of vertical gaze. Neurology 2000;54:1985–1993. 280. Henn V, Lang W, Hepp K, et al. Experimental gaze palsies in monkeys and 313. Daroff RR, Hoyt WE. Supranuclear disorders of ocular control systems in man: their relation to human pathology. Brain 1984;107:619–636. clinical, anatomical and physiological correlations, 1969. In: Bach-y-Rita P, 281. Beigi B, O’Keeffe M, Logan P, et al. Convergence substitution for paralysed Collins CC, Hyde JE, eds. The Control of Eye Movements. New York, Academic horizontal gaze. Br J Ophthalmol 1995;79:229–232. Press, 1971:175–235. 282. Bogousslavsky J, Regli F. Convergence and divergence synkinesis. A recovery 314. Keane JR. The pretectal syndrome: 206 patients. Neurology 1990;40:684–690. pattern in benign pontine hematoma. Neuroophthalmology 1984;4:219–225. 315. Pasik T, Pasik P, Bender MB. The superior colliculi and eye movements: an 283. Brusa G, Meneghini S, Piccardo A, et al. Regressive pattern of horizontal gaze experimental study in the monkey. Arch Neurol 1966;15:420–436. palsy. Case report. Neuroophthalmology 1987;7:301–306. 316. Bu¨ttner U, Bu¨ttner-Ennever JA, Rambold H, Helmchen C. The contribution of 284. Christoff N. A clinicopathologic study of vertical eye movements. Arch Neurol midbrain circuits in the control of gaze. Ann NY Acad Sci 2002;956:99–110. 1974;31:1–8. 317. Moschovakis AK, Scudder CA, Highstein SM. Structure of the primate oculomo- 285. Dominguez RO, Bronstein AM. Complete gaze palsy in pontine haemorrhage. tor burst generator. I. Medium-lead burst neurons with upward on-directions. J J Neurol Neurosurg Psychiatry 1988;51:150–151. Neurophysiol 1991;65:203–217. 286. Toyoda K, Hasegawa Y, Yonehara T, et al. Bilateral medial medullary infarction 318. Tamura EE, Hoyt CS. Oculomotor consequences of intraventricular hemorrhages with oculomotor disorders. Stroke 1992;23:1657–1659. in premature infants. Arch Ophthalmol 1987;105:533–535. 287. Hennerici M, Fromm C. Isolated complete gaze palsy: an unusual ocular move- 319. Barontini F, Simonetti C, Ferranini F, et al. Persistent upward eye deviation. ment deficit probably due to a bilateral parapontine reticular formation (PPRF) Report of two cases. Neuroophthalmology 1983;3:217–224. lesion. Neuroophthalmology 1981;1:165–173. 320. Ouvrier RA, Billson F. Benign paroxysmal tonic upgaze of childhood. J Child 288. Slavin ML. A clinicoradiographic correlation of bilateral horizontal gaze palsy Neurol 1988;3:177–180. with slowed vertical saccades with midline dorsal pontine lesion on magnetic 321. Apak RA, Topcu M. A case of paroxysmal tonic upgaze of childhood with resonance imaging. Am J Ophthalmol 1986;101:118–120. ataxia. Eur J Paediatr Neurol 1999;3:129–131. 289. Izaki A, Shimo-oku M, Suzuki A. Two cases of bilateral medial longitudinal 322. Hayman M, Harvey AS, Hopkins IJ, et al. Paroxysmal tonic upgaze: a reappraisal fasciculus syndrome with vertical saccades of slow velocity. Neuroophthalmol- of outcome. Ann Neurol 1998;43:514–520. ogy 1996;16(suppl):289. 323. Lispi ML, Vigevano F. Benign paroxysmal tonic upgaze of childhood with 290. Kaneko CRS. Effect of ibotenic acid lesions of the omnipause neurons on sac- ataxia. Epileptic Disord 2001;3:203–206. cadic eye movements in rhesus macaques. J Neurophysiol 1996;75:2229–2242. 324. Rouveyrol F, Stephan JL. Benign paroxysmal tonic upgaze of infancy: 2 addi- 291. Schon F, Hodgson TL, Mort D, Kennard C. Ocular flutter associated with a tional cases. Arch Pediatr 2003;10:527–529. localized lesion in the paramedian pontine reticular formation. Ann Neurol 2001; 325. Ruggieri VL, Yepez II, Fejerman N. [Benign paroxysmal tonic upward gaze 50:413–416. syndrome]. Rev Neurol 1998;27:88–91. 292. Wall M, Wray SH. The one-and-a-half syndrome: a unilateral disorder of the 326. Spalice A, Parisi P, Iannetti P. Paroxysmal tonic upgaze: physiopathological pontine tegmentum: a study of 20 cases and review of the literature. Neurology considerations in three additional cases. J Child Neurol 2000;15:15–18. 1983;33:971–980. 327. Verrotti A, Trotta D, Blasetti A, et al. Paroxysmal tonic upgaze of childhood: 293. Deleu D, Solheid C, Michotte A, et al. Dissociated ipsilateral horizontal gaze effect of age-of-onset on prognosis. Acta Paediatr 2001;90:1343–1345. palsy in one-and-a-half syndrome: a clinicopathologic study. Neurology 1988; 328. Pullicino P, Lincoff N, Truax BT. Abnormal vergence with upper brainstem 38:1278–1280. infarcts: pseudoabducens palsy. Neurology 2000;55:352–358. 294. Sharpe JA, Rosenberg MA, Hoyt WF, et al. Paralytic pontine exotropia: a sign of 329. Ochs AL, Stark L, Hoyt WF, et al. Opposed adducting saccades in convergence- acute unilateral pontine gaze palsy and internuclear ophthalmoplegia. Neurology retraction nystagmus. A patient with sylvian aqueduct syndrome. Brain 1979; 1974;24:1076–1081. 102:497–508. 295. Johkura K, Komiyama A, Kuroiwa Y. Eye deviation in patients with one-and- 330. Schnyder H, Bassetti C. Bilateral convergence nystagmus in unilateral dorsal a-half syndrome. Eur Neurol 2000;44:210–215. midbrain stroke due to occlusion of the superior cerebellar artery. Neuroophthal- 296. Johnston JL, Sharpe JA. Sparing of the vestibulo-ocular reflex with lesions of mology 1996;16:59–63. the paramedian pontine reticular formation. Neurology 1989;39:876. 331. Pasik T, Pasik P, Bender MB. The pretectal syndrome in monkeys. II. Sponta- 297. Heywood S, Ratcliff G. Long-term oculomotor consequences of unilateral colli- neous and induced nystagmus, and ‘‘lightning’’ eye movements. Brain 1969; culectomy in man. In: Lennerstrand G, Bach-y-Rita P, eds. Basic Mechanisms 92:871–884. of Ocular Motility and Their Clinical Implications. Oxford, Pergamon, 1975: 332. Rambold H, Kompf D, Helmchen C. Convergence retraction nystagmus: a disor- 561–564. der of vergence? Ann Neurol 2001;50:677–681. 298. de Seze J, Lucas C, Leclerc X, et al. One-and-a-half syndrome in pontine infarcts: 333. Galetta SL, Gray LG, Raps EC, et al. Pretectal eyelid retraction and lag. Ann MRI correlates. Neuroradiology 1999;41:666–669. Neurol 1993;33:554–557. 299. Bandini F, Faga D, Simonetti S. Ocular myasthenia mimicking a one-and-a-half 334. Saeki N, Yamaura A, Sunami K. Bilateral ptosis with sparing because of syndrome. J Neuroophthalmol 2001;21:210–211. a discrete midbrain lesion: magnetic resonance imaging evidence of topographic 300. Kunavarapu C, Kesavan RB, Pevil-Ulysee M, Mohan SS. Systemic lupus erythe- arrangement within the oculomotor nerve. J Neuroophthalmol 2000;20:130–134. matosus presenting as ‘‘one-and-a-half syndrome.’’ J Rheumatol 2001;28: 335. Wray SH. The neuro-ophthalmic and neurologic manifestations of pinealomas. 874–875. In: Schmider HH, ed. Pineal Tumors. New York, Masson, 1977:21–59. 301. Larner AJ, Rudge P. One-and-a-half syndrome resulting from spontaneous verte- 336. Baloh RW, Furman JM, Yee RD. Dorsal midbrain syndrome: clinical and oculo- bral artery dissection. Int J Clin Pract 2002;56:480–481. graphic findings. Neurology 1985;35:54–60. 302. Minagar A, Schatz NJ, Glaser JS. Case report: one-and-a-half-syndrome and 337. Swash M. Periaqueductal dysfunction (the Sylvian aqueduct syndrome): a sign tuberculosis of the pons in a patient with AIDS. AIDS Patient Care STDS 2000; of hydrocephalus? J Neurol Neurosurg Psychiatry 1974;37:21–26. 14:461–464. 338. Corbett JJ. Neuro-ophthalmologic complications of hydrocephalus and shunting 303. Crisostomo EA. One-and-a-half syndrome in a patient with metastatic breast procedures. Semin Neurol 1986;6:111–123. disease. J Clin Neuroophthalmol 1985;5:270–272. 339. Bleasel AF, Ell JJ, Johnston I. Pretectal syndrome and ventricular shunt dysfunc- 304. Jackel RA, Gittinger JW Jr, Smith TW, et al. Metastatic adenocarcinoma present- tion. Neuroophthalmology 1992;12:193–196. ing as a one-and-a-half syndrome. J Clin Neuroophthalmol 1986;6:116–119. 340. Rismondo V, Borchert M. Position-dependent Parinaud’s syndrome. Am J Oph- 305. Sharpe JA, Bondar RL, Fletcher WA. Contralateral gaze deviation after frontal thalmol 1992;114:107–108. lobe haemorrhage. J Neurol Neurosurg Psychiatry 1985;48:86–88. 341. Ko¨mpf D, Pasik T, Pasik P, et al. Downward gaze in monkeys. Stimulation and 306. Smith MS, Buchsbaum HW, Masland WS. . Occurrence lesion studies. Brain 1979;102:527–558. after trauma with computerized tomographic correlation. Arch Neurol 1980;37: 342. Henn V, Hepp K, Vilis T. Rapid eye movement generation in the primate. 251. Physiology, pathophysiology, and clinical implications. Rev Neurol (Paris) 307. Carter JE, Rauch RA. One-and-a-half syndrome, type II. Arch Neurol 1994;51: 1989;145:540–545. 87–89. 343. Suzuki Y, Buttner-Ennever JA, Straumann D, et al. Deficits in torsional and 308. Zee DS, Chu FC, Leigh RJ, et al. Blink-saccade synkinesis. Neurology 1983; vertical rapid eye movements and shift of Listing’s plane after uni- and bilateral 33:1233–l236. lesions of the rostral interstitial nucleus of the medial longitudinal fasciculus. 309. Bu¨ttner-Ennever JA, Wadia NH, Sakai H, et al. Neuroanatomy of oculomotor Exp Brain Res 1995;106:215–232. structures in olivo-pontocerebellar atrophy (OPCA) patients with slow saccades. 344. Zee DS. The organization of the brainstem ocular motor subnuclei. Ann Neurol J Neurol 1985;232(suppl):285. 1978;4:384–385. SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 959

345. Cogan DG, Chu FC, Bachman DM, et al. The DAF syndrome. Neuroophthalmol- 376. Jampel RS, Fells P. Monocular elevation paresis caused by a central nervous ogy 1981;2:7–16. system lesion. Arch Ophthalmol 1968;80:45–57. 346. Percheron C. Les arte`res du thalamus humain. II. Arteres et territoires thalamique 377. Lessell S, Wolf PA, Chronley D. Prolonged vertical nystagmus after pentobarbi- paramedians de l’arte`re basilair communicante. Rev Neurol (Paris) 1976;132: tal sodium administration. Am J Ophthalmol 1975;80:151–152. 309–324. 378. Ford CS, Schwartze GM, Weaver RG, et al. Monocular elevation paresis caused 347. Castaigne P, Lhermitte F, Buge A, et al. Paramedian thalamic and midbrain by an ipsilateral lesion. Neurology 1984;34:1264–1267. infarcts: clinical and neuropathological study. Ann Neurol 1981;10:127–148. 379. Bell JA, Fielder AR, Viney S. Congenital double elevator palsy in identical 348. Bu¨ttner-Ennever JA, Bu¨ttner U, Cohen B, et al. Vertical gaze paralysis and the twins. J Clin Neuroophthalmol 1990;10:32–34. rostral interstitial nucleus of the medial longitudinal fasciculus. Brain 1982;105: 380. Ziffer AJ, Rosenbaum AL, Demer JL, et al. Congenital double elevator palsy: 125–149. vertical saccadic velocity utilizing the scleral search coil technique. J Pediatr 349. Pierrot-Deseilligny C, Chain F, Gray F, et al. Parinaud’s syndrome: electro- Ophthalmol Strabis 1992;29:142–149. oculographic and anatomical analyses of six vascular cases with deductions 381. Wiest G, Baumgartner C, Schnider P, et al. Monocular elevation paresis and about vertical gaze organization in the premotor structures. Brain 1982;105: contralateral downgaze paresis from unilateral mesodiencephalic infarction. J 667–696. Neurol Neurosurg Psychiatry 1996;60:579–581. 350. Dehaene I, Casselman JW, Van Zandijcke M. Selective paralysis of downward 382. Gauntt CD, Kashii S, Nagata I. Monocular elevation paresis caused by an oculo- saccades. Cerebrovasc Dis 1994;4:377–378. motor fascicular impairment. J Neuroophthalmol 1995;15:11–14. 351. Deleu D, Buisseret T, Ebinger G. Vertical one-and-a-half syndrome: supranu- 383. Cadera W, Bloom JN, Karlik S, Viirre E. A magnetic resonance imaging study clear downgaze paralysis with monocular elevation palsy. Arch Neurol 1989; of double elevator palsy. Can J Ophthalmol 1997;32:250–253. 46:1361–1363. 384. Steele JC, Richardson JC, Olszewski J. Progressive supranuclear palsy: a hetero- 352. Bogousslavsky J, Regli F. Upgaze palsy and monocular paresis of downward geneous degeneration involving the brain stem, basal ganglia and cerebellum gaze from ipsilateral thalamo-mesencephalic infarction: a vertical ‘‘one-and-a- with vertical gaze and pseudobulbar palsy, nuchal dystonia and dementia. Arch half syndrome.’’ J Neurol 1984;231:43–45. Neurol 1964;10:333–359. 353. Hommel M, Bogousslavsky J. The spectrum of vertical gaze palsy following 385. Burn DJ, Lees AJ. Progressive supranuclear palsy: where are we now? Lancet unilateral brainstem stroke. Neurology 1991;41:1229–1234. 2002;1:359–369. 354. Sekine S, Utsumi H, Kaku H, et al. [A case presenting vertical one-and-a-half 386. Nath U, Ben Shlomo Y, Thomson RG, et al. Clinical features and natural history syndrome and seesaw nystagmus due to thalamomesencephalic infarction]. No of progressive supranuclear palsy: a clinical cohort study. Neurology 2003;60: To Shinkei 2003;55:699–703. 910–916. 355. Ranalli PJ, Sharpe JA, Fletcher WA. Palsy of upward and downward saccadic, 387. Goetz CG, Leurgans S, Lang AE, Litvan I. Progression of gait, speech and pursuit, and vestibular movements with a unilateral midbrain lesion: pathophysi- swallowing deficits in progressive supranuclear palsy. Neurology 2003;60: ologic correlations. Neurology 1988;38:114–122. 917–922. 356. Green JP, Newman NJ, Winterkorn JS. Paralysis of downgaze in two patients 388. Litvan I. Diagnosis and management of progressive supranuclear palsy. Semin with clinical-radiologic correlation. Arch Ophthalmol 1993;111:219–222. Neurol 2001;21:41–48. 357. Riordan-Eva P, Faldon M, Bu¨ttner-Ennever JA, et al. Abnormalities of torsional 389. Morris HR, Gibb G, Katzenschlager R, et al. Pathological, clinical and genetic fast eye movements in unilateral rostral midbrain disease. Neurology 1996;47: heterogeneity in progressive supranuclear palsy. Brain 2002;125:969–975. 201–207. 390. Golbe LI. Progressive supranuclear palsy in the molecular age. Lancet 2000; 358. Helmchen C, Rambold H, Kempermann U, et al. Localizing value of torsional 356:870–871. nystagmus in small midbrain lesions. Neurology 2002;59:1956–1964. 391. Riley DE, Fogt N, Leigh RJ. The syndrome of ‘‘pure akinesia’’ and its relation- 359. Ranalli PJ, Sharpe JA. Vertical vestibulo-ocular reflex, smooth-pursuit and eye- ship to progressive supranuclear palsy. Neurology 1994;44:1025–1029. head tracking dysfunction in internuclear ophthalmoplegia. Brain 1988;111: 392. de Ye´benes JG, Sarasa JL, Daniel SE, et al. Familial progressive supranuclear 1299–1317. palsy: description of a pedigree and review of the literature. Brain 1995;118: 360. Kokkoroyannis T, Scudder CA, Balaban CD, et al. Anatomy and physiology of 1095–1103. the primate interstitial nucleus of Cajal: Efferent projections. J Neurophysiol 393. Tetrud JW, Golbe LI, Forno LS, et al. Autopsy-proven progressive supranuclear 1996;75:725–739. palsy in two siblings. Neurology 1996;46:931–934. 361. Helmchen C, Rambold H, Fuhry L, Buttner U. Deficits in vertical and torsional 394. Seemungal BM, Faldon M, Revesz T, et al. Influence of target size on vertical eye movements after uni- and bilateral muscimol inactivation of the interstitial gaze palsy in a pathologically proven case of progressive supranuclear palsy. nucleus of Cajal of the alert monkey. Exp Brain Res 1998;119:436–452. Mov Disord 2003;18:818–822. 362. Helmchen C, Rambold H, Fuhry L, Buttner U. Deficits in vertical and torsional 395. Hamilton SR. Neuro-ophthalmology of movement disorders. Curr Opin Ophthal- eye movements after uni- and bilateral muscimol inactivation of the interstitial mol 2000;11:403–407. nucleus of Cajal of the alert monkey. Exp Brain Res 1998;119:436–452. 396. Dix MR, Harrison MJG, Lewis PD. Progressive supranuclear palsy (the Steele- 363. Fukushima K. The interstitial nucleus of Cajal and its role in the control of Richardson-Olszewski syndrome). A report of 9 cases with particular reference movements of head and eyes. Prog Neurobiol 1987;29:107–192. to the mechanism of the oculomotor disorder. J Neurol Sci 1971;13:237–256. 364. Halmagyi GM, Aw ST, Dehaene I, et al. Jerk-waveform seesaw nystagmus due 397. Troost BT, Daroff RB. The ocular motor defects in progressive supranuclear to unilateral meso-diencephalic lesion. Brain 1994;117:775–788. palsy. Ann Neurol 1977;2:397–403. 365. Rambold H, Helmchen C, Buttner U. Unilateral muscimol inactivations of the 398. Rascol O, Hain TC, Brefel C, et al. Antivertigo medications and drug-induced interstitial nucleus Cajal in the alert rhesus monkey do not elicit seesaw nystag- vertigo. Drugs 1995;50:777–791. mus. Neurosci Lett 1999;272:75–78. 399. Mastaglia FL, Grainger KMR. Internuclear ophthalmoplegia in progressive su- 366. Rambold H, Helmchen C, Buttner U. Vestibular influence on the binocular pranuclear palsy. J Neurol Sci 1975;25:303–308. control of vertical-torsional nystagmus after lesions in the interstitial nucleus of 400. Kitthaweesin K, Riley DE, Leigh RJ. Vergence disorders in progressive supranu- Cajal. Neuroreport 2000;11:779–784. clear palsy. Ann NY Acad Sci 2002;956:504–507. 367. Sharpe JA, Ranalli PJ. Vertical vestibulo-ocular reflex control after supranuclear 401. Pierrot-Deseilligny C, Rivaud S, Pillon B, et al. Lateral visually guided saccades midbrain damage. Acta Otolaryngol Suppl 1991;481:194–198. in progressive supranuclear palsy. Brain 1989;112:471–487. 368. Jacobs L, Heffner RR Jr, Newman RP. Selective paralysis of downward gaze 402. Pierrot-Deseilligny C, Rivaud-Pechoux S. [Contribution of oculomotor examina- caused by bilateral lesions of the mesencephalic periaqueductal gray matter. tion for the etiological diagnosis of parkinsonian syndromes]. Rev Neurol (Paris) Neurology 1985;35:516–521. 2003;159:3S75–3S81. 369. Thames PB, Trobe JD, Ballinger WE. Gaze paralysis caused by lesion of the 403. Rivaud-Pechoux S, Vermersch AI, Gaymard B, et al. Improvement of memory- periaqueductal gray matter. Arch Neurol 1984;41:437–440. guided saccades in parkinsonian patients by high-frequency subthalamic nucleus 370. Waitzman DM, Silakov VL, DePalma-Bowles S, Ayers AS. Effects of reversible stimulation. J Neurol Neurosurg Psychiatry 2000;68:381–384. inactivation of the primate mesencephalic reticular formation. II. Hypometric 404. Goffinet AM, De Volder AG, Gillian C, et al. Positron tomography demonstrates vertical saccades. J Neurophysiol 2000;83:2285–2299. frontal lobe hypometabolism in progressive supranuclear palsy. Ann Neurol 371. Waitzman DM, Silakov VL, DePalma-Bowles S, Ayers AS. Effects of reversible 1989;25:131–139. inactivation of the primate mesencephalic reticular formation. I. Hypermetric 405. Rafal RD, Posner MI, Friedman JH, et al. Orienting of visual attention in progres- goal-directed saccades. J Neurophysiol 2000;83:2260–2284. sive supranuclear palsy. Brain 1988;111:267–280. 372. Zackon DH, Sharpe JA. Midbrain paresis of horizontal gaze. Ann Neurol 1984; 406. Fisk JD, Goodale MA, Burkhart G, et al. Progressive supranuclear palsy: the 16:495–504. relationship between ocular motor dysfunction and psychological test perfor- 373. Reagan TJ, Trautmann JC. Combined nuclear and supranuclear defects in ocular mance. Neurology 1982;32:698–705. motility. A clinicopathologic study. Arch Neurol 1978;35:133–137. 407. Verny M, Duyckaerts C, Agid Y, et al. The significance of cortical pathology 374. Ropper AH, Miller DC. Acute traumatic midbrain hemorrhage. Ann Neurol in progressive supranuclear palsy: clinico-pathological data in 10 cases. Brain 1985;18:80–86. 1996;119:1123–1136. 375. Worthington JM, Halmagyi GM. Bilateral total ophthalmoplegia due to midbrain 408. Yekhlef F, Ballan G, Macia F, et al. Routine MRI for the differential diagnosis of hematoma. Neurology 1996;47:1176–1177. Parkinson’s disease, MSA, PSP, and CBD. J Neural Transm 2003;110:151–169. 960 CLINICAL NEURO-OPHTHALMOLOGY

409. Kato N, Arai K, Hattori T. Study of the rostral midbrain atrophy in progressive 441. Knox DL, Green Wm R, Troncoso JC, et al. Cerebral ocular Whipple’s disease: supranuclear palsy. Neurol Sci 2003;210:57–60. a 62-year odyssey from death to diagnosis. Neurology 1995;45:617–625. 410. Kida M, Koo H, Grossniklaus HE, et al. Neuropathologic findings in progressive 442. Misbah SA, Aslam A, Costello C. Whipple’s disease. Lancet 2004;363:654–656. supranuclear palsy. J Clin Neuroophthalmol 1988;8:161–170. 443. Averbuch-Heller L, Paulson GW, Daroff RB, Leigh RJ. Whipple’s disease mim- 411. Juncos JL, Hirsch EC, Malessa S, et al. Mesencephalic cholinergic nuclei in icking progressive supranuclear palsy: the diagnostic value of eye movement progressive supranuclear palsy. Neurology 1991;41:25–30. recording. J Neurol Neurosurg Psychiatry 1999;66:532–535. 412. Fukushima-Kudo J, Fukushima K, Tashiro K. Rigidity and dorsiflexion of the 444. Finelli PF, McEntee WJ, Lessell S, et al. Whipple’s disease with predominantly neck in progressive supranuclear palsy and the interstitial nucleus of Cajal. J neuroophthalmic manifestations. Ann Neurol 1977;1:247–252. Neurol Neurosurg Psychiatry 1987;50:1197–1203. 445. Knox DL, Bayless TM, Pittman FE. Neurologic disease in patients with treated 413. Revesz T, Sangha H, Daniel SE. The nucleus raphe interpositus in the Steele- Whipple’s disease. 1976;55:467–476. Richardson-Olszewski syndrome (progressive supranuclear palsy). Brain 1996; 446. Schwartz MA, Selhorst JB, Ochs AL, et al. Oculomasticatory myorhythmia: a 119:1137–1143. unique movement disorder occurring in Whipple’s disease. Ann Neurol 1986; 414. Averbuch-Heller L, Gordon C, Zivotofsky A, et al. Small vertical saccades have 20:677–683. normal speeds in progressive supranuclear palsy (PSP). Ann NY Acad Sci 2002; 447. Grotta JC, Pettigrew LC, Schmidt WA, et al. Oculomasticatory myorhythmia. 956:434–437 Ann Neurol 1987;22:395–396. 415. Bhidayasiri R, Riley DE, Somers JT, et al. Pathophysiology slow vertical sac- 448. Simpson DA, Wishnow R, Gargulinski RB, et al. Oculofacial-skeletal myorhyth- cades in progressive supranuclear palsy. Neurology 2001;57:2070–2077. mia in Whipple’s disease: additional case and review of 416. Halliday GM, Hardman CD, Cordato NJ, et al. A role for the substantia nigra the literature. Mov Disord 1995;10:195–200. pars reticulata in the gaze palsy of progressive supranuclear palsy. Brain 2000; 449. Quinn N. Rhythmic tremor of the palate and other cranial limb muscles, with 123:724–732. cerebellar ataxia: consider Whipple’s disease. Mov Disord 2001;16:787. 417. Malessa S, Gaymard B, Rivaud S, et al. Role of pontine nuclei damage in 450. Rajput AJ, Mchattie JD. Ophthalmoplegia and leg myorhythmia in Whipple’s smooth-pursuit impairment of progressive supranuclear palsy: a clinical patho- disease. Mov Disord 1997;12:111–114. logic study. Neurology 1994;44:716–721. 451. Albano JE, Mishkin M, Westbrook LE, et al. Visuomotor deficits following 418. Dubinsky RM, Jankovic J. Progressive supranuclear palsy and a multi-infarct ablation of monkey superior colliculus. J Neurophysiol 1982;48:338–350. state. Neurology 1987;37:570–576. 452. Pierrot-Deseilligny C, Rosa A, Masmoudi K, et al. Saccade deficits after a unilat- 419. Moses Ill H, Zee D. Multi-infarct PSP. Neurology 1987;37:1819. eral lesion affecting the superior colliculus. J Neurol Neurosurg Psychiatry 1991; 420. Josephs KA, Ishizawa T, Tsuboi Y, et al. A clinicopathological study of vascular 54:1106–1109. progressive supranuclear palsy: a multi-infarct disorder presenting as progressive 453. Fisher A, Knezevic W. Ocular and ocular motor aspects of primary thalamic supranuclear palsy. Arch Neurol 2002;59:1597–1601. haemorrhage. Clin Exp Neurol 1985;21:129–139. 421. Curran T, Lang AE. Parkinsonian syndromes associated with hydrocephalus: 454. Keane JR. Contralateral gaze deviation with supratentorial hemorrhage: three case reports, a review of the literature, and pathophysiological hypotheses. Mov pathologically verified cases. Arch Neurol 1975;32:119–122. Disord 1994;9:508–520. 455. Brigell M, Babikian V, Goodwin JA. Hypometric saccades and low-gain pursuit 422. Averbuch-Heller L, Paulson GW, Daroff RB, Leigh RJ. Whipple’s disease mim- resulting from a thalamic hemorrhage. Ann Neurol 1984;15:374–378. icking progressive supranuclear palsy: the diagnostic value of eye movement 456. Fisher CM. The pathologic and clinical aspects of thalamic hemorrhage. Trans recording. J Neurol Neurosurg Psychiatry 1999;66:532–535. Am Neurol Assoc 1959;84:56–59. 423. Mokri B, Ahlskog JE, Fulgham JR, Matsumoto JY. Syndrome resembling PSP 457. Choi KD, Jung DS, Kim JS. Specificity of ‘‘peering at the tip of the nose’’ for after surgical repair of ascending aorta dissection or aneurysm. Neurology 2004; a diagnosis of thalamic hemorrhage. Arch Neurol 2004;61:417–422. 62:971–973. 458. Gomez CR, Gomez SM, Selhorst JB. Acute thalamic esotropia. Neurology 1988; 424. Kuniyoshi S, Riley DE, Zee DS, et al. Distinguishing progressive supranuclear 38:1759–1762. palsy from other forms of Parkinson’s disease: evaluation of new signs. Ann 459. Hertle RW, Bienfang DC. Oculographic analysis of acute esotropia secondary NY Acad Sci 2002;956:484–486. to a thalamic hemorrhage. J Clin Neuroophthalmol 1990;10:21–26. 425. Riley DE, Lang AE, Lewis A, et al. Cortical-basal ganglionic degeneration. 460. Lindner K, Hitzenberger P, Drlicek M, et al. Dissociated unilateral convergence Neurology 1990;40:1203–1212. paralysis in a patient with thalamotectal haemorrhage. J Neurol Neurosurg Psy- 426. Vidailhet M, Rivaud S, Gouider-Khouja N, et al. Eye movements in parkinsonian chiatry 1992;55:731–733. syndromes. Ann Neurol 1994;35:420–426. 461. Siatkowski RM, Schatz NJ, Sellitti TP, et al. Do thalamic lesions really cause 427. Bergeron C, Pollanen MS, Weyer L, et al. Unusual clinical presentations of vertical gaze palsies? J Clin Neuroophthalmol 1993;13:190–193. cortical-basal ganglionic degeneration. Ann Neurol 1996;40:893–900. 428. Rottach K, Riley DE, DiScenna AO, et al. Dynamic properties of horizontal 462. Clark JM, Albert GW. Vertical gaze palsies from medial thalamic infarctions and vertical eye movements in parkinsonian syndromes. Ann Neurol 1996;39: without midbrain involvement. Stroke 1995;26:1467–1470. 368–377. 463. Biller J, Sand JJ, Corbett JJ, et al. Syndrome of the paramedian thalamic arteries: 429. Litvan I, Agid Y, Goetz C, et al. Accuracy of the clinical diagnosis of corticobasal clinical and neuroimaging correlation. J Clin Neuroophthalmol 1985;5:217–223. degeneration: a clinicopathologic study. Neurology 1997;48:119–125. 464. Lepore FE, Gulli V, Miller DC. Neuro-ophthalmological findings with neuro- 430. Caparros-Lefebvre D, Sergeant N, Lees A, et al. Guadeloupean parkinsonism: pathological correlation in bilateral thalamic-mesencephalic infarction. J Clin a cluster of progressive supranuclear palsy-like tauopathy. Brain 2002;125: Neuroophthalmol 1985;5:224–228. 801–811. 465. Weisbrod M, Ko¨lmel HW, Ha¨ttig H, et al. The significance of vertical gaze 431. Lepore FE, Steele JC, Cox TA, et al. Supranuclear disturbances of ocular motility palsy in the paramedian thalamic artery syndrome. Neuroophthalmology 1992; in Lytico-Bodig. Neurology 1988;38:1849–1853. 12:85–90. 432. Steele JC, Caparros-Lefebvre D, Lees AJ, Sacks OW. Progressive supranuclear 466. Beversdorf DQ, Jenkyn LR, Petrowski III JT, et al. Vertical gaze paralysis and palsy and its relation to pacific foci of the parkinsonism-dementia complex and intermittent unresponsiveness in a patient with a thalamomesencephalic stroke. Guadeloupean parkinsonism. Parkinsonism Relat Disord 2002;9:39–54. J Neuroophthalmol 1995;15:230–235. 433. Brett FM, Henson C, Staunton H. Familial diffuse Lewy body disease, eye 467. Albano JE, Wurtz RH. Deficits in eye position following ablation of monkey movement abnormalities, and distribution of pathology. Arch Neurol 2002;59: superior colliculus, pretectum, and posterior-medial thalamus. J Neurophysiol 464–467. 1982;48:318–337. 434. Louis ED, Klataka LA, Liu Y, et al. Comparison of extrapyramidal features in 468. Gaymard B, Rivaud S, Pierrot-Deseilligny C. Impairment of extraretinal eye 31 pathologically confirmed cases of diffuse Lewy body disease and 34 patholog- position signals after central thalamic lesions. Exp Brain Res 1994;102:1–9. ically confirmed cases of Parkinson’s disease. Neruology 1997;48:376–380. 469. Sommer MA, Wurtz RH. What the brain stem tells the frontal cortex. I. Oculo- 435. Saver JL, Liu GT, Charness ME. Idiopathic striopallidodentate calcification with motor signals sent from superior colliculus to frontal eye field via mediodorsal prominent supranuclear abnormality of eye movement. J Neuroophthalmol 1994; thalamus. J Neurophysiol 2004;91:1381–1402. 14:29–33. 470. Sommer MA, Wurtz RH. What the brain stem tells the frontal cortex. II. Role 436. Wszolek ZK, Pfeiffer RF, Bhatt MH, et al. Rapidly progressive autosomal domi- of the SC-MD-FEF pathway in corollary discharge. J Neurophysiol 2004;91: nant parkinsonism and dementia with pallido-ponto-nigral degeneration. Ann 1403–1423. Neurol 1992;32:312–320. 471. Gaymard B, Rivaud S, Pierrot-Deseilligny C. Impairment of extraretinal eye 437. Rafal RD, Grimm RJ. Progressive supranuclear palsy: functional analysis of position signals after central thalamic lesions. Exp Brain Res 1994;102:1–9. the response to methysergide and antiparkinsonian agents. Neurology 1981;31: 472. Gaymard B, Rivaud-Pechoux S, Yelnik J, et al. Involvement of the cerebellar 1507–1518. thalamus in human saccade adaptation. Eur J Neurosci 2001;14:554–560. 438. Newman GC. Treatment of progressive supranuclear palsy with tricyclic antide- 473. Zihl J, Von Cramon D. The contribution of the ‘‘second’’ to pressants. Neurology 1985;35:1189–1193. directed visual attention in man. Brain 1979;102:835–856. 439. Jackson JA, Jankovic J, Ford J. Progressive supranuclear palsy: clinical features 474. Ogren MP, Mateer CA, Wyler AR. Alterations in visually related eye movements and response to treatment in 16 patients. Ann Neurol 1983;13:273–278. following left pulvinar damage in man. Neuropsychologia 1984;22:187–196. 440. Golbe LI. Progressive supranuclear palsy. Curr Treat Options Neurol 2001;3: 475. Vilkki J. Visual hemi-inattention after ventrolateral thalamotomy. Neuropsycho- 473–477. logia 1984;22:399–408. SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 961

476. Bender DB, Baizer JS. Saccadic eye movements following kainic acid lesions 510. White OB, Saint-Cyr JA, Tomlinson RD, et al. Ocular motor deficits in Parkin- of the pulvinar in monkeys. Exp Brain Res 1990;79:467–478. son’s disease. III. Coordination of eye and head movements. Brain 1988;111: 477. Kennard C, Lueck CJ. Oculomotor abnormalities in diseases of the basal ganglia. 115–129. Rev Neurol 1989;145:587–595. 511. Baloh RW, Jacobson K, Honrubia V. Idiopathic bilateral vestibulopathy. Neurol- 478. White OB, Saint-Cyr JA, Tomlinson RD, et al. Ocular motor deficits in Parkin- ogy 1989;39:272–275. son’s disease. II. Control of the saccadic and smooth-pursuit systems. Brain 512. Lekwuwa GU, Barnes GR, Collins CJ, Limousin P. Progressive bradykinesia 1983;106:571–587. and of ocular pursuit in Parkinson’s disease. J Neurol Neurosurg 479. Corin MS, Elizan TS, Bender MB. Oculomotor function in patients with Parkin- Psychiatry 1999;66:746–753. son’s disease. J Neurol Sci 1972;15:251–265. 513. Melvill Jones G, DeJong JD. Visual tracking of sinusoidal target movement in 480. Chamberlain W. Restriction in upward gaze with advancing age. Am J Ophthal- Parkinson’s disease. In D.R.B. Aviation Medical Research Unit (Department of mol 1971;71:341–346. Physiology), Vol 5. McGill University, Montreal, 1974–1976:271–288. 481. Repka MX, Claro MC, Loupe DN, et al. Ocular motility in Parkinson’s disease. 514. Reichert WH, Doolittle J, McDowell FH. Vestibular dysfunction in Parkinson Pediatr Ophthalmol Strabis 1996;33:144–147. disease. Neurology 1982;32:1133–1138. 481a. Biousse V, Skibell BC, Watts RL, et al. Ophthalmologic feature of Parkinson’s 515. White OB, Saint-Cyr JA, Sharpe JA. Ocular motor deficits in Parkinson’s dis- disease. Neurology 2004;62:177–180. ease. I. The horizontal vestibulo-ocular reflex and its regulation. Brain 1983; 482. Crawford T, Goodrich S, Henderson L, et al. Predictive responses in Parkinson’s 106:555–570. disease: manual keypresses and saccadic eye movements to regular stimulus 516. Merchut MP, Brigell M. Olivopontocerebellar atrophy presenting with hemipar- events. J Neurol Neurosurg Psychiatry 1989;52:1033–1042. kinsonian ocular motor signs. J Clin Neuroophthalmol 1990;10:210–214. 483. DeJong JD, Melvill Jones G. Akinesia, hypokinesia, and bradykinesia in the 517. Rinne JO, Lee MS, Thompson PD, et al. Corticobasal degeneration: a clinical oculomotor system of patients with Parkinson’s disease. Exp Neurol 1971;32: study of 36 cases. Brain 1994;117:1183–1196. 58–68. 518. Garbutt S, Harwood MR, Kumar AN, et al. Evaluating small eye movements 484. Melvill Jones C, DeJong JD. Dynamic characteristics of saccadic eye movements in patients with saccadic palsies. Ann NY Acad Sci 2003;1004:337–346. in Parkinson’s disease. Exp Neurol 1971;31:17–31. 519. Jacobs DH, Adair JC, Heilman KM. Visual grasp in corticobasal degeneration. 485. Teravainen H, Calne DB. Studies of parkinsonian movement: I. Programming Ann Neurol 1994;36:679. and execution of eye movements. Acta Neurol Scand 1980;62:137–148. 520. Eckert T, Sailer M, Kaufmann J, Schrader C, et al. Differentiation of idiopathic 486. Muller C, Wenger S, Fertl L, et al. Initiation of visual-guided random saccades Parkinson’s disease, multiple system atrophy, progressive supranuclear palsy, and remembered saccades in parkinsonian patients with severe motor-fluctua- and healthy controls using magnetization transfer imaging. Neuroimage 2004; tions. J Neural Transm 1994;7:101–108. 21:229–235. 487. Hikosaka O, Segawa M, Imai H. Voluntary saccadic eye movements. Applica- 521. Leigh RJ, Foley JM, Remler BF, et al. Oculogyric crisis: a syndrome of thought tion to analyze basal ganglion functions. In: Highlights in Neuro-Ophthalmol- disorder and ocular deviation. Ann Neurol 1987;22:13–17. ogy. Proceedings of Sixth Meeting of the International Neuroophthalmology 522. Antunes NL, Small TN, George D, Boulad F, Lis E. Posterior leukoencephalopa- Society (INOS). Amsterdam, Aeolus Press, 1987:133–138. thy syndrome may not be reversible. Pediatr Neurol 1999;20:241–243. 488. Lueck CJ, Tanyeri S, Crawford TJ, et al. Saccadic eye movements in Parkinson’s 523. FitzGerald P, Jankovic J. Tardive oculogyric crises. Neurology 1989;39: disease. I. Delayed saccades. Quart J Exp Psychol 1992;45A:193–210. 1434–1437. 489. Lueck CJ, Crawford TJ, Henderson L, et al. Saccadic eye movements in Parkin- 524. FitzGerald PM, Jankovic J, Glaze DG, et al. Extrapyramidal involvement in son’s disease. II. Remembered saccades—towards a unified hypothesis? Q J Rett’s syndrome. Neurology 1990;40:293–295. Exp Psychol 1992;45A:211–233. 525. Sachdev P. Tardive and chronically recurrent oculogyric crises. Mov Disord 490. Kimmig H, Haussmann K, Mergner T, Lucking CH. What is pathological with 1993;8:93–97. gaze shift fragmentation in Parkinson’s disease? J Neurol 2002;249:683–692. 526. Liu GT, Carrazana EJ, Macklis JD, et al. Delayed oculogyric crises associated 491. Shaunak S, O’Sullivan E, Blunt S, et al. Remembered saccades with variable with striatocapsular infarction. J Clin Neuroophthalmol 1991;11:198–201. delay in Parkinson’s disease. Mov Disord 1999;14:80–86. 527. Shimpo, Fuse S, Yoshizawa A. Retrocollis and oculogyric crisis in association 492. Nakamura T, Bronstein AM, Lueck CJ, et al. Vestibular, cervical and visual with bilateral putaminal hemorrhages. Clin Neurol 1933;33:40–44. remembered saccades in Parkinson’s disease. Brain 1994;117:1423–1432. 528. Kim JS, Kim HK, Im JH, et al. Oculogyric crisis and abnormal magnetic reso- 493. Vermersch AI, Rivaud S, Vidailhet M, et al. Sequences of memory-guided sac- nance imaging signals in bilateral lentiform nuclei. Mov Disorders 1996;11: cades in Parkinson’sdisease. Ann Neurol 1994;35:487–490. 756–758. 494. Crawford TJ, Henderson L, Kennard C. Abnormalities of nonvisually guided 529. Gibson JM, Pimlott R, Kennard C. Ocular motor and manual tracking in Parkin- eye movements in Parkinson’s disease. Brain 1989;112:1573–1586. 495. Bronstein AM, Kennard C. Predictive ocular motor control in Parkinson’s dis- son’s disease and the effect of treatment. J Neurol Neurosurg Psychiatry 1987; ease. Brain 1985;108:925–940. 50:853–860. 496. Tanyeri S, Lueck CJ, Crawford TJ, et al. Vertical and horizontal saccadic eye 530. Highstein SM, Cohen B, Mones R. Changes in saccadic eye movements of movements in Parkinson’s disease. Neuroophthalmology 1989;9:165–177. patients with Parkinson’s disease before and after L-dopa. Trans Am Neurol 497. Waterston JA, Barnes GR, Grealy MA, et al. Abnormalities of smooth eye and Assoc 1969;94:277–279. head movement control in Parkinson’s disease. Ann Neurol 1996;39:749–760. 531. Hotson JR, Langston EB, Langston JW. Saccade responses to dopamine in 498. Rascol O, Clanet M, Montastruc JL, et al. Abnormal ocular movements in Parkin- human MPTP-induced parkinsonism. Ann Neurol 1986;20:456–463. son’s disease. Evidence for involvement of dopaminergic systems. Brain 1989; 532. Brooks BA, Fuchs AR, Finocchio D. Saccadic eye movement deficits in the 112:1193–1214. MPTP monkey model of Parkinson’s disease. Brain Res 1986;383:404–407. 499. Lueck CJ, Tanyeri S, Crawford TJ, et al. Antisaccades and remembered saccades 533. Schultz W, Romo R, Scarnati E, et al. Saccadic reaction times, eye–arm coordi- in Parkinson’s disease. J Neurol Neurosurg Psychiatry 1990;53:284–288. nation and spontaneous eye movements in normal and MPTP–treated monkeys. 500. Fukushima J, Fukushima K, Miyasaka K, et al. Voluntary control of saccadic Exp Brain Res 1989;78:253–267. eye movement in patients with frontal cortical lesions and parkinsonian patients 534. Gibson JM, Kennard C. Quantitative study of ‘‘on-off’’ fluctuations in the ocular in comparison with that in schizophrenics. Biol Psychiatry 1994;36:21–30. motor system in Parkinson’s disease. Adv Neurol 1986;45:329–333. 501. Briand KA, Strallow D, Hening W, et al. Control of voluntary and reflexive 535. Sharpe JA, Fletcher WA, Lang AE, et al. Smooth pursuit during dose-related saccades in Parkinson’s disease. Exp Brain Res 1999;129:38–48. on-off fluctuations in Parkinson’s disease. Neurology 1987;37:1389–1392. 502. Vidailhet M, Rivaud S, Gouider-Khouja N, et al. Saccades and antisaccades in 536. Shimizu N, Cohen B, Bala SP, et al. Ocular dyskinesias in patients with Parkin- parkinsonian syndromes. Adv Neurol 1999;80:377–382. son’s disease treated with levodopa. Ann Neurol 1977;1:167–171. 503. Hodgson TL, Tiesman B, Owen AM, Kennard C. Abnormal gaze strategies 537. Hikosaka O. Role of basal ganglia in saccades. Rev Neurol (Paris) 1989;145: during problem solving in Parkinson’s disease. Neuropsychologia 2002;40: 580–586. 411–422. 538. Hikosaka O, Wurtz RH. The basal ganglia. In: Wurtz RH, Goldberg ME, eds. 504. Ketcham CJ, Hodgson TL, Kennard C, Stelmach GE. Memory-motor transfor- The Neurobiology of Saccadic Eye Movements. Reviews of Oculomotor Re- mations are impaired in Parkinson’s disease. Exp Brain Res 2003;149:30–39. search. Vol 3. Amsterdam, Elsevier, 1989:257–281. 505. Shaunak S, O’Sullivan E, Blunt S, et al. Remembered saccades with variable 539. Ho PC, Feman SS. Internuclear ophthalmoplegia in Fabry’s disease. Ann Oph- delay in Parkinson’s disease. Mov Disord 1999;14:80–86. thalmol 1981;13:949–951. 506. MacAskill MR, Anderson TJ, Jones RD. Saccadic adaptation in neurological 540. Starr A. A disorder of rapid eye movements in Huntington’s chorea. Brain 1967; disorders. Prog Brain Res 2002;140:417–431. 90:545–567. 507. Averbuch-Heller L, Stahl JS, Hlavin ML, Leigh RJ. Square-wave jerks induced 541. Leigh RJ, Newman SA, Folstein SE, et al. Abnormal ocular motor control in by pallidotomy in parkinsonian patients. Neurology 1999;52:185–188. Huntington’s disease. Neurology 1983;33:1268–1275. 508. Blekher T, Siemers E, Abel LA, Yee RD. Eye movements in Parkinson’s disease: 542. Lasker AG, Zee DS, Hain TC, et al. Saccades in Huntington’s disease: initiation before and after pallidotomy. Invest Ophthalmol Vis Sci 2000;41:2177–2183. defects and distractability. Neurology 1987;37:364–370. 509. Kennard C, Zangemeister WH, Mellors S, et al. Eye-head coordination in Parkin- 543. Winograd-Gurvich CT, Georgiou-Karistianis N, et al. Hypometric primary sac- son’s disease. In: Lennerstrand G, Zee DS, Keller EL, et al. Functional Basis cades and increased variability in visually-guided saccades in Huntington’s dis- of Ocular Motility Disorders. Oxford, UK, Pergamon, 1982:517–520. ease. Neuropsychologia 2003;41:1683–1692. 962 CLINICAL NEURO-OPHTHALMOLOGY

544. Schapira AH. Mitochondrial function in Huntington’s disease: clues for patho- blepharospasm and involuntary levator palpebrae inhibition. Brain 1994;117: genesis and prospects for treatment. Ann Neurol 1997;41:141–142. 1457–1474. 545. Zangemeister WN, Mueller Jensen A. The co-ordination of gaze movements in 578. Bollen E, Van Exel E, van der Velde EA, et al. Saccadic eye movements in Huntington’s disease. Neuroophthalmology 1985;5:193–206. idiopathic blepharospasm. Mov Disord 1996;11:678–682. 546. Lasker AG, Zee DS, Hain TC, et al. Saccades in Huntington’s disease: slowing 579. Stell R, Bronstein AM, Marsden CD. Vestibulo-ocular abnormalites in spas- and dysmetria. Neurology 1988;38:427–431. modic torticollis before and after botulinum injections. J Neurol Neurosurg 547. Hotson JR, Louis AA, Langston EB, et al. Vertical saccades in Huntington’s Psychiatry 1989;52:57–62. disease and nondegenerative choreoathetoid disorders. Neuroophthalmology 580. Thaker GK, Nguyen JA, Tamminga CA. Increased saccadic distractibility in 1984;4:207–217. tardive dyskinesia: functional evidence for subcortical GABA dysfunction. Biol 548. Rubin AJ, King WM, Reinbold KA, et al. Quantitative longitudinal assessment Psychiatry 1989;25:49–59. of saccades in Huntington’s disease. J Clin Neuroophthalmol 1993;13:59–66. 581. Jinnah HA, Lewis RF, Visser JE, et al. Ocular motor dysfunction in Lesch- 549. Garcia Ruiz PJ, Cenjor C, Ulmer E, et al. [Speed of ocular saccades in Huntington Nyhan disease. Pediatr Neurol 2001;24:200–204. disease. Prospective study]. Neurologia 2001;16:70–73. 582. Pierrot-Deseilligny C. Eye movement control by the cerebral cortex. Curr Opin 550. Fielding J, Georgiou-Karistianis N, Bradshaw J, et al. Impaired modulation of Neurol. 2004;17:17–25. the vestibulo-ocular reflex in Huntington’s disease. Mov Disord 2004;19:68–75. 583. Leigh RJ, Kennard C. Using saccades as a research tool in the clinical neurosci- 551. Avanzini G, Girotti F, Caraceni T, et al. Oculomotor disorders in Huntington’s ences. Brain 2004;127:460–477. chorea. J Neurol Neurosurg Psychiatry 1979;42:581–589. 584. Mort DJ, Malhotra P, Mannan SK, Rorden C, et al. The anatomy of visual 552. Oyanagi K, Takeda S, Takahashi H, et al. A quantitative investigation of the neglect. Brain 2003;126:1986–1997. substantia nigra in Huntington’s disease. Ann Neurol 1989;26:13–19. 585. Gitelman DR. Attention and its disorders. Br Med Bull 2003;65:21–34. 553. Koeppen AH. The nucleus pontis centralis caudalis in Huntington’s disease. J 586. Mapstone M, Weintraub S, Nowinski C, et al. Cerebral hemispheric specializa- Neurol Sci 1989;91:129–141. tion for spatial attention: spatial distribution of search-related eye fixations in 554. Leigh RJ, Parhad IM, Clark AW, et al. Brainstem findings in Huntington’s the absence of neglect. Neuropsychologia 2003;41:1396–1409. disease. Possible mechanisms for slow vertical saccades. J Neurol Sci 1985;71: 587. Mort DJ, Kennard C. Visual search and its disorders. Curr Opin Neurol 2003; 247–256. 16:51–57. 555. Rothlind JC, Brandt J, Zee D, et al. Verbal memory and oculomotor control are 588. Goodwin JA, Kansu T. Vulpian’s sign: conjugate eye deviation in acute cerebral unimpaired in asymptomatic adults with the genetic marker for Huntington’s hemisphere lesions. Neurology 1986;36:711–712 . disease. Arch Neurol 1993;50:799–802. 589. Tijssen CC. Horizontal conjugate eye deviation. A clinical and electrophysiologi- 556. Reveley MA, Dursun SM, Andrews H. Improvement of abnormal saccadic eye cal study. Doctoral thesis, Nijmegen University, The Netherlands, 1988. movements in Huntington’s disease by sulpiride: a case study. J Psychopharma- 590. Ko¨mpf D, Gmeiner HJ. Gaze palsy and visual hemineglect in acute hemisphere col 1994;8:262–265. lesions. Neuroophthalmology 1989;9:49–53. 557. Rinne JO, Daniel SE, Scaravilli F, et al. The neuropathological features of neuro- 591. Johnston CW. Eye movements in visual hemi-neglect. In: Johnston CW, Piroz- acanthocytosis. Mov Disord 1994;9:297–304. zolo FJ, eds. Neuropsychology of Eye Movements. Hillsdale, NJ, Lawrence 558. Nielsen JE, Sorensen S, Hasholt L, et al. Dentatorubral-pallidoluysian atropy. Earlbaum, 1988:235–263. Clinical features of a five-generation Danish family. Mov Disord 1996;11: 592. Steiner I, Melamed E. Conjugate eye deviation after acute hemispheric stroke: 533–541. delayed recovery after previous contralateral frontal lobe damage. Ann Neurol 559. Burke JR, Wingfield MS, Lewis KE, et al. The Haw River syndrome: dentatoru- 1984;16:509–511. bropallidoluysian atrophy (DRPLA) in an African-American family. Nature 593. Lekwuwa GU, Barnes GR. Cerebral control of eye movements. I. The relation- Genet 1994;7:521–524. ship between cerebral lesion sites and smooth-pursuit deficits. Brain 1996;119: 560. Stein RW, Kase CS, Hier DB, et al. Caudate hemorrhage. Neurology 1984;34: 473–490. 1549–1554. 594. Tusa RJ, Zee DS, Nerdman SJ. Effect of unilateral cerebral cortical lesions on 561. Vermersch AI, Muri RM, Rivaud S, et al. Saccade disturbances after bilateral ¨ ocular motor behavior in monkeys: saccades and quick phases. J Neurophysiol lentiform nucleus lesions in humans. J Neurol Neurosurg Psychiatry 1996;60: 1986;56:1590–1625. 179–184. 595. Fletcher WA, Gellman RS. Saccades in humans with lesions of frontal eye fields 562. Lekwuwa GU, Barnes GR. Cerebral control of eye movements. II. Timing of (FEF). Society for Neuroscience Abstracts 1989;15:1203. anticipatory eye movements, predictive pursuit and phase errors in focal cerebral 596. Meador KJ, Loring DW, Lee GP, et al. Hemisphere asymmetry for eye gaze lesions. Brain 1996;119:491–505. mechanisms. Brain 1989;112:103–111. 563. Frankel M, Cummings JL. Neuro-ophthalmic abnormalities in Tourette’s syn- drome: functional and anatomic implications. Neurology 1984;34:359–361. 597. Lesser RP, Leigh RJ, Dinner DS, et al. Preservation of voluntary saccades after 564. Elston JS, Casgranje F, Lees AJ. The relationship between eye-winking tics, intracarotid injection of barbiturate. Neurology 1985;35:1108–1112. frequent eye-blinking and blepharospasm. J Neurol Neurosurg Psychiatry 1989; 598. Tijssen CC, Van Gisbergen JAM, Schulte BPM. Conjugate eye deviation: side, 52:477–480. site, and size of the hemispheric lesion. Neurology 1991;41:846–850. 565. Bollen EL, Roos RAC, Cohen AP, et al. Oculomotor control in Gilles de 1a 599. Tijssen CC, Van Gisbergen JAM. Conjugate eye deviation after hemispheric . J Neurol Neurosurg Psychiatry 1988;51:1081–1083. stroke: a contralateral saccadic palsy? Neuroophthalmology 1993;13:107–118. 566. Binyon S, Prendergast M. Eye-movement tics in children. Dev Med Child Neurol 600. Morrow MJ. Craniotopic defects of smooth-pursuit and saccadic eye movement. 1991;33:343–355. Neurology 1996;46:514–521. 567. Shawkat F, Harris CM, Jacobs M, et al. Eye movement tics. Br J Ophthalmol 601. Tijssen CC. Contralateral conjugate eye deviation in acute supratentorial lesions. 1992;76:697–699. Stroke 1994;25:1516–1519. 568. Dursun SM, Burke JG, Reveley MA. Antisaccade eye movement abnormalities 602. Blacker DJ, Wijdicks EF. Delayed complete bilateral ptosis associated with in Tourette syndrome: evidence for cortico-striatal network dysfunction? J Psy- massive infarction of the right hemisphere. Mayo Clin Proc 2003;78:836–839. chopharmacol 2000;14:37–39. 603. Cogan DG. Neurologic significance of lateral conjugate deviation of the eyes 569. Farber RH, Swerdlow NR, Clementz BA. Saccadic performance characteristics on forced closure of the lids. Arch Ophthalmol 1948;39:37–42. and the behavioural neurology of Tourette’s syndrome. J Neurol Neurosurg 604. Sullivan HC, Kaminski HJ, Maas EF, et al. Lateral deviation of the eyes on Psychiatry 1999;66:305–312. forced lid closure in patients with cerebral lesions. Arch Neurol 1991;48: 570. LeVasseur AL, Flanagan JR, Riopelle RJ, Munoz DP. Control of volitional and 310–311. reflexive saccades in Tourette’s syndrome. Brain 2001;124:2045–2058. 605. Sharpe JA, Lo AW, Rabinovitch HE. Control of the saccadic and smooth-pursuit 571. Munoz DP, Le Vasseur AL, Flanagan JR. Control of volitional and reflexive systems after cerebral hemidecortication. Brain 1979;102:387–403. saccades in Tourette’s syndrome. Prog Brain Res 2002;140:467–481. 606. Troost BT, Daroff RB, Weber RB, et al. Hemispheric control of eye movements. 572. Mostofsky SH, Lasker AG, Singer HS, et al. Oculomotor abnormalities in boys II. Quantitative analysis of smooth pursuit in a hemispherectomy patient. Arch with Tourette syndrome with and without ADHD. J Am Acad Child Adolesc Neurol 1972;27:449–452. Psychiatry 2001;40:1464–1472. 607. Cogan DG, Loeb DR. Optokinetic response and intracranial lesions. Arch Neurol 573. Straube A, Mennicken J-B, Riedel M, et al. Saccades in Gilles de la Tourette’s Psychiatry 1949;61:183–187. syndrome. Mov Disorders 1997;12:536–546. 608. Troost BT, Daroff RB, Weber RB, et al. Hemispheric control of eye movements. 574. Klein CH, Raschke A, Brandenbusch A. Development of pro- and antisaccades II. Quantitative analysis of smooth pursuit in a hemispherectomy patient. Arch in children with attention-deficit hyperactivity disorder (ADHD) and healthy Neurol 1972;27:449–452. controls. Psychophysiology 2003;40:17–28. 609. Estanol B, Romero R, de Viteri MS, et al. Oculomotor and oculovestibular 575. Munoz DP, Armstrong IT, Hampton KA, Moore KD. Altered control of visual functions in a hemispherectomy patient. Arch Neurol 1980;37:365–368. fixation and saccadic eye movements in attention-deficit hyperactivity disorder. 610. Sharpe JA, Lo AW. Voluntary and visual control of the vestibuloocular reflex J Neurophysiol 2003;90:503–514. after cerebral hemidecortication. Ann Neurol 1981;10:164–172. 576. Demer JL, Holds JB, Hovis LA. Ocular movements in essential blepharospasm. 611. Traccis S, Puliga MV, Ruiu MC, et al. Unilateral occipital lesion causing hemia- Am J Ophthalmol 1990;110:674–682. nopia affects the acoustic saccadic programming. Neurology 1991;41: 577. Aramideh M, Bour LJ, Koelman JHTM, et al. Abnormal eye movements in 1633–1638. SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 963

612. Zihl J. Eye movement patterns in hemianopic dyslexia. Brain 1995;118: 646. Bogousslavsky J. Impairment of visually evoked eye movements with a unilat- 891–912. eral parieto-occipital lesion. J Neurol 1987;234:160–162. 613. Meienberg O, Bu¨ttner-Ennever JA, Kraus-Ruppert R. Unilateral paralysis of 647. Ron S, Schmid R, Orpaz D. Applying a model of saccadic prediction to patients’ conjugate gaze due to lesion of the abducens nucleus: clinico-pathological case saccadic eye movements. Brain Behav Evol 1989;33:179–182. report. Neuroophthalmology 1981;2:47–52. 648. Muri RM, Gaymard B, Rivaud S, et al. Hemispheric asymmetry in cortical 614. Meienberg O, Harrer M, Wehren C. Oculographic diagnosis of hemineglect in control of memory-guided saccades. A transcranial magnetic stimulation study. patients with homonymous hemianopia. J Neurol 1986;233:97–101. Neuropsychologia 2000;38:1105–1111. 615. Meienberg O. Clinical examination of saccadic eye movements in hemianopia. 649. Shimozaki SS, Hayhoe MM, Zelinsky GJ, et al. Effect of parietal lobe lesions Neurology 1983;11:1315. on saccade targeting and spatial memory in a naturalistic visual search task. 616. Zangemeister WN, Meienenberg O, Stark L, et al. Eye-head coordination in Neuropsychologia 2003;41:1365–1386. homonymous hemianopia. J Neurol 1982;43:254. 650. Posner MI, Walker JA, Friedrich FJ, et al. Effects of parietal injury on covert 617. Segraves MA, Goldberg ME, Deng SY, et al. The role of striate cortex in the orienting of attention. J Neurosci 1984;4:1863–1874. guidance of eye movements in the monkey. J Neurosci 1987;7:3040–3058. 651. Pierrot-Deseilligny C. Saccades and smooth-pursuit impairment after cerebral 618. Kolmel HW, Nabel HJ. Optokinetic nystagmus in homonymous hemianopia due hemispheric lesions. Eur Neurol 1994;34:121–134. to a strictly occipital lesion. Eur Arch Psychiatry Neurol Sci 1989;99:202. 652. Walker R, Findlay JM. Saccadic eye movement programming in unilateral neg- 619. Rizzo M, Hurtig R. The effect of bilateral visual cortex lesions on the develop- lect. Neuropsychologia 1996;34:493–508. ment of eye movements and perception. Neurology 1989;6:413. 653. Ferber S, Danckert J, Joanisse M, et al. Eye movements tell only half the story. 620. Zee DS, Tusa RJ, Herdman SJ, et al. Effects of occipital lobectomy upon eye Neurology 2003;60:1826–1829. movements in primate. J Neurophysiol 1987;58:883–907. 654. Duhamel J-R, Goldberg ME, Fitzgibbon EJ, et al. Saccadic dysmetria in a patient 621. Brindley GS, Gautier-Smith PC, Lewin W. Cortical blindness and the functions with a right frontoparietal lesion: the importance of corollary discharge for accu- of the nongeniculate fibers of the optic tracts. J Neurol Neurosurg Psychiatry rate spatial-behavior. Brain 1992;115:1387–1402. 1969;9:264. 655. Vuilleumier P, Schwartz S. Modulation of visual perception by eye gaze direc- 622. Fox JC, Holmes G. Optic nystagmus and its value in the localization of cerebral tion in patients with spatial neglect and extinction. Neuroreport 2001;12: lesions. Brain 1926;3:371. 2101–2104. 623. Kjallman L, Frise´n L. The cerebral ocular pursuit pathways: a clinicoradiological 656. Gaymard B, Lynch J, Ploner CJ, et al. The parieto-collicular pathway: anatomical study of small-field optokinetic nystagmus. J Clin Neuroophthalmol 1986;6: location and contribution to saccade generation. Eur J Neurosci 2003;17: 209–217. 1518–1526. 624. Heide W, Koenig E, Dichgans J. Optokinetic nystagmus, self-motion sensation 657. Zihl J, Hebel N. Patterns of oculomotor scanning in patients with unilateral and their after-effects in patients with occipitoparietal lesions. Clin Vis Sci 1990; posterior parietal or frontal lobe damage. Neuropsychologia 1997;35:893–906. 156:133–143. 658. Niemeier M, Karnath HO. Exploratory saccades show no direction-specific defi- 625. Heide W, Kurzidim K, Ko¨mpf D. Deficits of smooth-pursuit eye movements cit in neglect. Neurology 2000;54:515–518. after frontal and parietal lesions. Brain 1996;119:1951–1969. 659. Wardak C, Olivier E, Duhamel JR. Saccadic target selection deficits after lateral 626. Morrow MJ, Sharpe JA. Retinoptic and directional deficits of smooth-pursuit intraparietal area inactivation in monkeys. J Neurosci 2002;22:9877–9884. initiation after posterior cerebral hemispheric lesions. Neurology 1993;43: 660. Leigh RJ, Tusa RJ. Disturbance of smooth pursuit caused by infarction of occipi- 595–603. toparietal cortex. Ann Neurol 1985;17:185–187. 627. Dursteler MR, Wurtz RH. Pursuit and optokinetic deficits following chemical 661. Carmichael EA, Dix MR, Hall-Pike CS. Lesions of the cerebral hemispheres and lesions of cortical areas MT and MST. J Neurophysiol 1988;60:940–965. their effects upon optokinetic and caloric nystagmus. Brain 1954;77:345–372. 628. Leigh RJ. The cortical control of ocular pursuit movements. Rev Neurol (Paris) 662. Straube A, Brandt T. Importance of the visual and vestibular cortex for self- 1989;5:612. motion perception in man (circularvection). Hum Neurobiol 1987;6:211–218. 629. Barton JJS, Sharpe JA, Raymond JE. Retinoptic and directional defects in motion 663. Bottini G, Karnath HO, Vallar G, et al. Cerebral representations for egocentric discrimination in humans with cerebral lesions. Ann Neurol 1995;5:675. space: functional-anatomical evidence from caloric vestibular stimulation and 630. Barton JJS, Sharpe JA, Raymond JE. Directional defects in pursuit and motion neck vibration. Brain 2001;124:1182–1196. perception in humans with unilateral cerebral lesions. Brain 1996;535:1550. 664. Kluge M, Beyenburg S, Fernandez G, Elger CE. Epileptic vertigo: evidence for 631. Barton JJS, Simpson T, Kiriakopoulos E, et al. Functional MRI of lateral occipi- vestibular representation in human frontal cortex. Neurology 2000;55: totemporal cortex during pursuit and motion perception. Ann Neurol 1996;7: 1906–1908. 398. 665. Brandt T, Dieterich M. The vestibular cortex. Its locations, functions, and disor- 632. Thurston SE, Leigh RJ, Crawford T, et al. Two distinct deficits of visual tracking ders. Ann NY Acad Sci 1999;871:293–312. caused by unilateral lesions of cerebral cortex in humans. Ann Neurol 1988;6: 666. Bucher SF, Dieterich M, Wiesmann M, et al. Cerebral functional magnetic reso- 266–273. nance imaging of vestibular, auditory, and nociceptive areas during galvanic 633. Morrow MJ, Sharpe JA. Smooth-pursuit initiation in humans with cerebral hemi- stimulation. Ann Neurol 1998;44:120–125. spheric lesions. Society for Neuroscience Abstracts 1989;14:612. 667. Dieterich M, Brandt T. Vestibular system: anatomy and functional magnetic 634. Lawden MC, Bagelmann H, Crawford TJ, et al. An effect of structured back- resonance imaging. Neuroimaging Clin North Am 2001;11:263–273. grounds on smooth-pursuit eye movements in patients with cerebral lesions. 668. Dieterich M, Brandt T. Brain activation studies on visual-vestibular and ocular Brain 1995;118:37–48. motor interaction. Curr Opin Neurol 2000;13:13–18. 635. Bogousslavsky J, Regli F. Pursuit gaze defects in acute and chronic unilateral 669. Schneider D, Schneider L, Claussen CF, Kolchev C. Cortical representation of parieto-occipital lesions. Eur Neurol 1986;25:10–18. the vestibular system as evidenced by brain electrical activity mapping of vestib- 636. Ghika J, Ghika-Schmid F, Bogousslasvky J. Parietal motor syndrome: a clinical ular late evoked potentials. Ear Nose Throat J 2001;80:251–260. description in 32 patients in the acute phase of pure parietal strokes studied 670. Kahane P, Hoffmann D, Minotti L, Berthoz A. Reappraisal of the human vestibu- prospectively. Clin Neurol Neurosurg 1998;100:271–282. lar cortex by cortical electrical stimulation study. Ann Neurol 2003;54:615–624. 637. Morrow MJ, Sharpe JA. Cerebral hemispheric localization of smooth-pursuit 671. Israe¨l I, Rivaud S, Gaymard B, et al. Cortical control of vestibular-guided sac- asymmetry. Neurology 1990;40:284–292. cades. Brain 1995;118:1169–1184. 638. Ko¨mpf D, Oppermann J. Vertical gaze palsy and thalamic dementia. Syndrome 672. Mu¨ri RM, Rivaud S, Timsit S, et al. Role of the medial temporal lobe in the of the posterior thalamosubthalamic paramedian artery. Neuroophthalmology control of memory-guided saccade. Exp Brain Res 1994;101:165–168. 1986;6:121–124. 673. Nielsen JM. Tornado epilepsy simulating Me´nie`re’s syndrome. Neurology 1959; 639. Pierrot-Deseilligny C, Rivaud S, Samson Y, et al. Some instructive cases con- 9:794–796. cerning the circuitry of ocular smooth pursuit in the brainstem. Neuroophthal- 674. Smith BN. Vestibular disturbances in epilepsy. Neurology 1960;10:465–469. mology 1989;9:31–42. 675. Barac B. Vertiginous epileptic attacks and so-called ‘‘vestibulogenic seizures.’’ 640. Baloh RW, Yee RD, Honrubia V. Optokinetic nystagmus and parietal lobe le- Epilepsia 1968;9:137–144. sions. Ann Neurol 1980;7:269–276. 676. Schneider RC, Calhoun HD, Crosby EC. Vertigo and rotational movement in 641. Ventre J, Faugier-Grimaud S. Effects of posterior parietal lesions (area 7) on cortical and subcortical lesions. J Neurol Sci 1968;6:493–516. VOR in monkeys. Exp Brain Res 1986;62:654–658. 677. Kogeorgos J, Scott DF, Swash M. Epileptic dizziness. Br Med J 1981;282: 642. Pierrot-Deseilligny C, Rivaud S, Penet C, et al. Latencies of visually guided 687–689. saccades in unilateral hemispheric cerebral lesions. Ann Neurol 1987;21: 678. Furman JMR, Crumrine PK, Reinmuth OM. Epileptic nystagmus. Ann Neurol 138–148. 1990;27:686–688. 643. Sundqvist A. Saccadic reaction-time in parietal-lobe dysfunction. Lancet 1979; 679. Wiest G, Zimprich F, Prayer D, et al. Vestibular processing in human paramedian 1:870. precuneus as shown by electrical cortical stimulation. Neurology 2004;62: 644. Lynch JC, McLaren JW. Deficits of visual attention and saccadic eye movements 473–475. after lesions of parieto-ocipital cortex in monkeys. J Neurophysiol 1989;61: 680. Latto R, Cowey A. Fixation changes after frontal eye-field lesions in monkey. 74–90. Brain Res 1971;30:25–36. 645. Heide W, Blankenburg M, Zimmerman E, et al. Cortical control of double-step 681. Tanaka H, Arai M, Kubo J, Hirata K. Conjugate eye deviation with head version saccades: implications for spatial orientation. Ann Neurol 1995;38:739–748. due to a cortical infarction of the frontal eye field. Stroke 2002;33:642–643. 964 CLINICAL NEURO-OPHTHALMOLOGY

682. Pessin MS, Adelman LS, Prager RJ, et al. ‘‘Wrong-way eyes’’ in supratentorlal 717. Catalano RA, Calhoun JH, Reinecke RD, et al. Asymmetry in congenital ocular hemorrhage. Ann Neurol 1981;9:79–81. motor apraxia. Can J Ophthalmol 1988;23:318–321. 683. Blanke O, Spinelli L, Thut G, et al. Location of the human frontal eye field as 718. Eustace P, Beigi B, Bowell R, et al. Congenital ocular motor apraxia: an inability defined by electrical cortical stimulation: anatomical, functional and electrophys- to unlock the vestibulo-ocular reflex. Neuroophthalmology 1994;14:167–174. iological characteristics. Neuroreport 2000;11:1907–1913. 719. Cogan DG. Ophthalmic manifestations of bilateral non-occipital cerebral lesions. 684. Sharpe JA. Adaptation to frontal lobe lesions. In: Keller EL, Zee DS, eds. Adap- Br J Ophthalmol 1965;49:281–297. tive Processes in Visual and Oculomotor Systems. Oxford, Pergamon, 1986: 720. Russo PA, Flynn MF, Veith J. Congenital ocular motor apraxia with torsional 239–246. oscillations: a case report. Optom Vis Sci 1995;72:925–930. 685. Magel-Leiby S, Buchtel HA, Welch A. Cerebral control of directed visual atten- 721. Shawkat FS, Kingsley D, Kendall B, et al. Neuroradiological and eye movement tion and orienting saccades. Brain 1990;113:237–276. correlates in children with intermittent saccade failure: ‘‘ocular motor apraxia.’’ 686. Braun DI, Weber H, Mergner T, et al. Saccadic reaction times in patients with Neuropediatrics 1995;26:298–305. frontal and parietal lesions. Brain 1992;115:1359–1386. 722. Le BI, Moreira MC, Rivaud-Pechoux S, et al. Cerebellar ataxia with oculomotor 687. Rivaud S, Muri RM, Gaymard B, et al. Eye movement disorders after frontal apraxia type 1: clinical and genetic studies. Brain 2003;126:2761–2772. eye field lesions in humans. Exp Brain Res 1994;102:110–120. 723. Le B, Bouslam N, Rivaud-Pechoux S, et al. Frequency and phenotypic spectrum 688. Deng S-Y, Goldberg ME, Segraves MA, et al. The effect of unilateral ablation of ataxia with oculomotor apraxia 2: a clinical and genetic study in 18 patients. of the frontal eye fields on saccadic performance in the monkey. In: Keller EL, Brain 2004;127:759–767. Zee DS, eds. Adaptive Processes in Visual and Oculomotor Systems. Oxford, 724. Tusa RJ, Hove MT. Ocular and oculomotor signs in Joubert syndrome. J Child Pergamon, 1986:201–208. Neurol 1999;14:621–627. 689. Guitton D, Buchtel HA, Douglas RM. Frontal lobe lesions in man cause difficul- 725. Prasad P, Nair S. Congenital ocular motor apraxia: sporadic and familial. J ties in suppressing reflexive glances and in generating goal-directed saccades. Neuroophthalmol 1994;14:102–104. Exp Brain Res 1985;58:455–472. 726. Gurer YK, Kukner S, Kunak B, et al. Congenital ocular motor apraxia in two 690. Pierrot-Deseilligny C, Rivaud S, Gaymard B, et al. Cortical control of reflexive siblings. Pediatr Neurol 1995;13:261–262. visually-guided saccades. Brain 1991;114:1473–1485. 727. Stell R, Bronstein AM, Plant GT, et al. Ataxia telangiectasia: a reappraisal of 691. Machado L, Rafal RD. Control of fixation and saccades during an anti-saccade the ocular motor features and their value in the diagnosis of atypical cases. Mov task: an investigation in humans with chronic lesions of oculomotor cortex. Disord 1989;4:320–329. Neuropsychology 2004;18:115–123 728. Patterson MC, Horowitz M, Abel RB, et al. Isolated horizontal supranuclear 692. Pierrot-Deseilligny C, Israe¨l I, Berthoz A, et al. Role of the differential frontal gaze palsy as a marker of severe systemic involvement in Gaucher’s disease. lobe areas in the control of the horizontal component of memory-guided saccades Neurology 1993;43:1993–1997. in man. Exp Brain Res 1993;95:166–171. 729. Gascon GG, Abdo N, Sigut D, et al. Ataxia-oculomotor apraxia syndrome. J 693. Pierrot-Deseilligny C, Muri RM, et al. Decisional role of the dorsolateral prefron- Child Neurol 1995;10:118–122. tal cortex in ocular motor behaviour. Brain 2003;126:1460–1473. 730. Nezu A. Neurophysiological study in Pelizaeus-Merzbacher disease. Brain Dev 694. Iba M, Sawaguchi T. Involvement of the dorsolateral prefrontal cortex of mon- 1995;17:175–181. keys in visuospatial target selection. J Neurophysiol 2003;89:587–599. 731. Aicardi J, Barbosa C, Andermann E, et al. Ataxia-ocular motor apraxia: a syn- 695. Milea D, Napolitano M, Dechy H, et al. Complete bilateral horizontal gaze drome mimicking ataxia-telangiectasia. Ann Neurol 1988;24:497–502. paralysis disclosing multiple sclerosis. J Neurol Neurosurg Psychiatry 2001;70: 732. Farr AK, Shalev B, Crawford TO, et al. Ocular manifestations of ataxia-telangi- 252–255. ectasia. Am J Ophthalmol 2002;134:891–896. 696. Gaymard B, Pierrot-Deseilligny C, Rivaud S. Impairment of sequences of mem- 733. Lewis RF, Crawford TO. Slow target–directed eye movements in ataxia-telangi- ory guided saccades after supplementary motor area lesions. Ann Neurol 1990; ectasia. Invest Ophthalmol Vis Sci 2002;43:686–691. 28:622–626. 734. Lewis RF. Ocular motor apraxia and ataxia-telangiectasia. Arch Neurol 2001; 697. Lu X, Matsuzawa M, Hikosaka O. A neural correlate of oculomotor sequences 58:1312. in supplementary eye field. Neuron 2002;34:317–325. 735. Lewis RF, Lederman HM, Crawford TO. Ocular motor abnormalities in ataxia 698. Pierrot-Deseilligny C, Rivaud S, Gaymard B, et al. Cortical control of memory- telangiectasia. Ann Neurol 1999;46:287–295. guided saccades in man. Exp Brain Res 1991;83:607–617. 736. Currie J, Ramsden B, McArthur C, et al. Validation of a clinical antisaccade eye 699. Husain M, Parton A, Hodgson TL, et al. Self-control during response conflict movement test in the assessment of dementia. Arch Neurol 1991;48:644–648. by human supplementary eye field. Nat Neurosci 2003;6:117–118. 737. Abel LA, Unverzagt F, Yee RD. Effects of stimulus predictability and interstimu- 700. Morrow MJ, Sharpe JA. Deficits of smooth-pursuit eye movement after unilat- lus gap on saccades in Alzheimer’s disease. Dement Geriatr Cogn Disord 2002; eral frontal lobe lesions. Ann Neurol 1995;37:443–451 701. Fukushima K. Frontal cortical control of smooth-pursuit. Curr Opin Neurobiol 13:235–243. 2003;13:647–654. 738. Lueck KL, Mendez MF, Perryman KM. Eye movement abnormalities during reading in patients with Alzheimer disease. Neuropsychiatry Neuropsychol 702. Moser A, Ko¨mpf D. Unilateral visual exploration deficit in a frontal lobe lesion. Neuroophthalmology 1990;10:39–44. Behav Neurol 2000;13:77–82. 703. Pierrot-Deseilligny C, Gautier JC, Loron P. Acquired ocular motor apraxia due 739. Schewe HJ, Uebelhack R, Vohs K. Abnormality in saccadic eye movement in to bilateral frontoparietal infarcts. Ann Neurol 1988;23:199–202. dementia. Eur Psychiatry 1999;14:52–53. 704. Dehaene I, Lammens M. Paralysis of saccades and pursuit: clinicopathologic 740. Shafiq-Antonacci R, Maruff P, Masters C, Currie J. Spectrum of saccade system study. Neurology 1991;41:414–415. function in Alzheimer disease. Arch Neurol 2003;60:1272–1278. 705. Schiller PH, True SD, Conway JL. Deficits in eye movements following frontal 741. Fletcher WA, Sharpe JA. Saccadic eye movement dysfunction in Alzheimer’s eye-field and superior colliculus ablations. J Neurophysiol 1980;44:1175–1189. disease. Ann Neurol 1986;20:464–471. 706. Rizzo M. ‘‘Balint’s syndrome’’ and associated visuospatial disorders. Baillieres 742. Daffner KR, Scinto LFM, Weintraub S, et al. Diminished curiosity in patients Clin Neurol 1993;2:415–437. with probable Alzheimer’s disease as measured by exploratory eye movements. 707. Tsutsui J, Takeda J, Ichihashi S, et al. Ocular motor apraxia and lesions of the Neurology 1992;42:320–328. visual association area. Neuroophthalmology 1980;1:149–154. 743. Scinto LFM, Daffner KR, Castro L, et al. Impairment of spatially directed atten- 708. Pierrot-Deseilligny C, Gray F, Brunet P. Infarcts of both inferior parietal lobules tion in patients with probable Alzheimer’s disease as measured by eye move- with impairment of visually guided eye movements, peripheral visual attention ments. Arch Neurol 1994;51:682–688. and optic ataxia. Brain 1986;109:81–97. 744. Hof PR, Bouras C, Constantinidis J, et al. Balint’s syndrome in Alzheimer dis- 709. Pierrot-Deseilligny C. Controle cortical des saccades. Rev Neurol (Paris) 1989; ease: specific disruption of the occipito-parietal visual pathway. Brain Res 1989; 145:596–604. 493:368–375. 710. Watson RT, Rapcsak SZ. Loss of spontaneous blinking in a patient with Balint’s 745. Fletcher WA, Sharpe JA. Smooth-pursuit dysfunction in Alzheimer’s disease. syndrome. Arch Neurol 1989;46:567–570. Neurology 1988;38:272–277. 711. Gottlieb D, Calvanio R, Levine DN. Reappearance of the visual percept after 746. Zaccara G, Gangemi PF, Muscas GC, et al. Smooth-pursuit eye movements: intentional blinking in a patient with Balint’s syndrome. J Clin Neuroophthalmol alterations in Alzheimer’s disease. J Neurol Sci 1992;112:81–89. 1991;11:62–65. 747. Bertoni JM, Label LS, Sackelleres JC, et al. Supranuclear gaze palsy in familial 712. Johnston JL, Sharpe JA, Morrow MJ. Spasm of fixation: a quantitative study. Creutzfeldt-Jakob disease. Arch Neurol 1983;40:618–622. J Neurol Sci 1992;107:166–171. 748. Grant MP, Cohen M, Petersen RB, et al. Abnormal eye movements in Creutz- 713. Gittinger JW Jr, Sokol S. The visual- in the diagnosis of congeni- feldt-Jakob disease. Ann Neurol 1993;34:192–197. tal ocular motor apraxia. Am J Ophthalmol 1982;93:700–703. 749. Helmchen C, Bu¨ttner U. Centripetal nystagmus in a case of Creutzfeldt-Jacob 714. Zee DS, Yee RD, Singer HS. Congenital ocular motor apraxia. Brain 1977;100: disease. Neuroophthalmology 1985;15:187–192. 581–589. 750. Smith SJM, Kocen RS. A Creutzfeldt-Jakob-like syndrome due to lithium toxic- 715. Fielder AR, Gresty MA, Dodd KL, et al. Congenital ocular motor apraxia. Trans ity. J Neurol Neurosurg Psychiatry 1988;51:120–123. Ophthalmol Soc UK 1986;105:589–598. 751. Gordon MF, Abrams RI, Rubin DB, et al. Bismuth subsalicylate toxicity as 716. Harris C, Shawkat F, Russell-Eggitt I, et al. Intermittent horizontal saccade a cause of prolonged encephalopathy with myoclonus. Mov Disord 1995;10: failure (‘‘ocular motor apraxia’’) in children. Br J Ophthalmol 1996;80:151–158. 220–222. SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 965

752. Farlow MR, Yee RD, Dlouhy SR, et al. Gerstmann-Straussler-Scheinker disease. 790. Wyllie E, Luders H, Morris HH, et al. Ipsilateral forced head and eye turning I. Extending the clinical spectrum. Neurology 1989;39:1446–1452. at the end of the generalized tonic-clonic phase of versive seizures. Neurology 753. Yee RD, Farlow MR, Suzuki DA, et al. Abnormal eye movements in Gerstmann- 1986;36:1212–1217. Stra¨ussler-Scheinker disease. Arch Ophthalmol 1992;110:68–74. 791. Ochs R, Gloor P, Quesney F, et al. Does headturning during a seizure have 754. Currie J, Benson E, Ramsden B, et al. Eye movement abnormalities as a predictor lateralizing or localizing significance? Neurology 1984;34:884–890. of the acquired immunodeficiency syndrome dementia complex. Arch Neurol 792. Gloor P, Quesney F, Ives J, et al. Significance of direction of head turning 1988;9:953. during seizures. Neurology 1987;37:1092. 755. Hamed LM, Schatz NJ, Galetta SL. Brainstem ocular motility defects and AIDS. 793. Bender MB. Oscillopsia. Arch Neurol 1965;13:204–213. Am J Ophthalmol 1988;106:437–442. 794. Vignaendra V, Lim CL. Epileptic discharges triggered by eye convergence. 756. Jabs DA, Green WR, Fox R, et al. Ocular manifestations of acquired immune Neurology 1978;28:589–591. deficiency syndrome. Ophthalmology 1989;92:1099. 795. Shanzer S, April R, Atkin A. Seizures induced by eye deviation. Arch Neurol 757. Keane JR. Neuro-ophthalmologic signs in AIDS: 50 patients. Neurology 1991; 1965;13:621–626. 1:845. 796. Fisher CM. The neurological examination of the comatose patient. Acta Neurol 758. Hedges TR III. Ophthalmoplegia associated with AIDS. Surv Ophthalmol 1994; Scand 1969;45(Suppl 36):1–56. 39:51. 797. Plum F, Posner JB. Diagnosis of Stupor and Coma, 3rd ed. Philadelphia, FA 759. Merril PT, Paige GD, Abrams RA, et al. Ocular motor abnormalities in human Davis, 1980. immunodeficiency virus infection. Ann Neurol 1991;30:130–138. 798. Buettner UW, Zee DS. Vestibular testing in comatose patients. Arch Neurol 760. Johnston JL, Miller JD, Nath A. Ocular motor dysfunction in HIV-1-infected 1989;46:561–563. subjects: a quantitative oculographic analysis. Neurology 1996;1:457. 799. Simon RP. Forced downward ocular deviation. Occurrence during oculovestibu- 761. Brekelmans GJF, Tijssen CC. Acquired ocular motor apraxia in an AIDS patient lar testing in sedative drug-induced coma. Arch Neurol 1978;35:456–458. with bilateral fronto-parietal lesions. Neuroophthalmology 1990;56. 800. Rosenberg ML. The eyes in hysterical states of unconsciousness. J Clin Neuro- 762. Tagliati M, Simpson D, Morgello S, et al. Cerebellar degeneration associated ophthalmol 1982;2:259–260. with human immunodeficiency virus infection. Neurology 1998;4:251. 801. Keane JR. Sustained upgaze in coma. Ann Neurol 1981;9:409–412. 763. Pfister HW, Einhaupl KM, Bu¨ttner U, et al. Dissociated nystagmus as a common 802. Nakada T, Kwee IL, Lee H. Sustained upgaze in coma. J Clin Neuroophthalmol sign of ocular motor disorders in HIV-infected patients. Eur Neurol 1989;7:280. 1984;4:35–37. 764. Nguyen N, Rimmer S, Katz B. Slowed saccades in the acquired immunodefi- 803. Keane JR. Spastic eyelids. Failure of levator inhibition in unconscious states. ciency syndrome. Am J Ophthalmol 1989;107:356–360. Arch Neurol 1975;32:695–698. 765. Sweeney JA, Brew BJ, Keilp JG, et al. Pursuit eye movement dysfunction in 804. Keane JR. Acute vertical ocular myoclonus. Neurology 1986;36:86–89. HIV-1 seropositive individuals. J Psychiatr Neurosci 1991;16:247–252. 805. Fisher CM. Ocular bobbing. Arch Neurol 1964;11:543–546. 766. Thurston SE, Leigh RJ, Osorio L. Epileptic gaze deviation and nystagmus. Neu- 806. Mehler MF. The clinical spectrum of ocular bobbing and ocular dipping. J Neurol rology 1985;35:1518–1521. Neurosurg Psychiatry 1988;51:725–727. 767. Kaplan PW, Tusa RJ. Neurophysiologic and clinical correlations of epileptic 807. Brusa A, Firpo MP, Massa S, et al. Typical and reverse bobbing: a case with nystagmus. Neurology 1993;43:2508–2514. localizing value. Eur Neurol 1984;23:151–155. 768. Gire C, Somma-Mauvais H, Nicaise C, et al. Epileptic nystagmus: electroclinical 808. Rosenberg ML, Calvert PC. Ocular bobbing in association with other signs of study of a case. Epileptic Disord 2001;3:33–37. midbrain dysfunction. Arch Neurol 1986;43:314. 769. Grant AC, Jain V, Bose S. Epileptic monocular nystagmus. Neurology 2002; 809. Goldschmidt TJ, Wall M. Slow upward ocular bobbing. J Clin Neuroophthalmol 59:1438–1441. 1987;7:241–243. 770. Hughes JR, Fino JJ. Epileptic nystagmus and its possible relationship with PGO 810. Titer EM, Laureno R. Inverse/reverse ocular bobbing. Ann Neurol 1988;23: spikes. Clin Electroencephalogr 2003;34:32–38. 103–104. 771. Kellinghaus C, Loddenkemper T, Luders HO. Epileptic monocular nystagmus. 811. Rosenberg ML, Calvert PC. Ocular bobbing in association with other signs of Neurology 2003;61:145–147. midbrain dysfunction. Arch Neurol 1986;43:314. 772. Moster ML, Schnayder E. Epileptic periodic alternating nystagmus. J Neurooph- 812. Noda S, Ide K, Umezaki M, et al. Repetitive divergence. Ann Neurol 1987;21: thalmol 1998;18:292–293. 109–110. 773. Goodfellow GW, Allison CL, Schiange DG. Unusual eye movements in a patient 813. Oesterle CS, Faulkner WJ, Clay R, et al. Eye bobbing associated with jaw with complex partial seizure disorder. Optometry 2002;73:160–165. movement. Ophthalmology 1982;89:63–67. 774. Garcia-Pastor A, Lopez-Esteban P, Peraita-Adrados R. Epileptic nystagmus: a 814. Ishikawa H, Ishikawa S, Mukuno K. Short-cycle periodic (ping-pong) gaze. case study video-EEG correlation. Epileptic Disord 2002;4:23–28. 775. Bogacz J, Bogacz D, Bogacz A. Oculomotor phenomena in petit-mal. Clin Neu- Neurology 1993;43:1067–1070. rophysiol 2000;111:959–963. 815. Johkura K, Komiyama A, Tobita M, Hasegawa O. Saccadic ping-pong gaze. J 776. Kent L, Blake A, Whitehouse W. Eyelid myoclonia with absences: phenomenol- Neuroophthalmol 1998;18:43–46. ogy in children. Seizure 1998;7:193–199. 816. Senelick RC. ‘‘Ping-pong’’ gaze. Periodic alternating gaze deviation. Neurology 777. Young GB, Brown JD, Bolton CF, et al. Periodic lateralized epileptiform dis- 1976;26:532–535. charges (PLEDs) and nystagmus retractorius. Ann Neurol 1977;2:61–62. 817. Larmande P, Dongmo L, Limodin J, et al. Periodic alternating gaze: a case 778. Brenner RP, Carlow TJ. PLEDS and nystagmus retractorius. Ann Neurol 1979; without any hemispheric lesion. Neurosurgery 1987;20:481–483. 5:403. 818. Leigh RJ, Hanley DF, Munschauer FE III, et al. Eye movements induced by 779. Nelson KR, Brenner RP, Carlow TJ. Divergent-convergent eye movements and head rotation in unresponsive patients. Ann Neurol 1984;15:465–473. transient eyelid opening associated with an EEG burst-suppression pattern. J 819. Simon RP, Aminoff MJ. Electrographic status epilepticus in fatal anoxic coma. Clin Neuroophthalmol 1986;6:43–46. Ann Neurol 1986;20:351–355. 780. Watanabe K, Negoro T, Matsumoto A, et al. Epileptic nystagmus associated 820. Larmande P, Henin D, Jan M, et al. Abnormal vertical eye movements in the with typical absence seizures. Epilepsia 1984;25:22–24. locked-in syndrome. Ann Neurol 1982;11:100–102. 781. Horita H, Hoashi E, Okuyama Y, et al. The studies of the attacks of abnormal 821. Meienberg O, Mumenthaler M, Karbowski K. Quadriparesis and nuclear oculo- eye movement in a case of infantile spasms. Folia Psychiatrica Neurol Jpn 1977; motor palsy with total bilateral ptosis mimicking coma. A mesencephalic 31:393–402. ‘‘locked-in syndrome’’? Arch Neurol 1979;36:708–710. 782. Rosenbaum DA, Siegel M, Rowan AJ. Contraversive seizures in occipital epi- 822. Uematsu D, Suematsu M, Fukuuchi Y, et al. Midbrain locked-in state with lepsy: case report and review of the literature. Neurology 1986;36:281–284. oculomotor subnucleus lesion. J Neurol Neurosurg Psychiatry 1985;48:952–956. 783. Tusa RJ, Kaplan PW, Main TC, et al. Ipsiversive eye deviation and epileptic 823. Rosenberg M, Sharpe J, Hoyt WF. Absent vestibulo-ocular reflexes and acute nystagmus. Neurology 1990;40:662–665. supratentorial lesions. J Neurol Neurosurg Psychiatry 1975;38:6–10. 784. Stolz SE, Chatrian G, Spence AM. Epileptic nystagmus. Epilepsia 1991;32: 824. Hanid MA, Silk DBA, Williams R. Prognostic value of the oculovestibular reflex 910–918. in fulminant hepatic failure. Br Med J 1978;1:1029. 785. Keating EG, Gooley SG. Disconnection of parietal and occipital access to the 825. Mueller-Jensen A, Neunzig HP, Emskotter T. Outcome prediction in comatose saccadic oculomotor system. Exp Brain Res 1988;70:385–398. patients: significance of reflex eye movement analysis. J Neurol Neurosurg Psy- 786. Kaplan PW, Lesser RP. Vertical and horizontal epileptic gaze deviation and chiatry 1987;50:389–392. nystagmus. Neurology 1989;39:1391–1393. 826. Levy DE, Plum F. Outcome prediction in comatose patients: significance of 787. Bruce CJ, Goldberg ME, Bushnell MC, et al. Primate frontal eye fields. II. reflex eye movement analysis. J Neurol Neurosurg Psychiatry 1988;51:318. Physiological and anatomical correlates of electrically evoked eye movements. 827. Singh BM, Strobos RJ. Retraction nystagmus elicited by bilateral simultaneous J Neurophysiol 1985;54:714–734. cold caloric stimulation. Ann Neurol 1980;8:79. 788. Quesney LF. Seizures of frontal lobe origin. In: Pedley TA, Meldrum BS, eds. 828. Dougherty JH, Rawlinson DG, Levy DE, et al. Hypoxic-ischemic brain injury Recent Advances in Epilepsy, 3. Edinburgh, Churchill Livingstone, 1986: and the vegetative state: clinical and neuropathologic correlation. Neurology 81–110. 1981;31:991–997. 789. Wyllie E, Luders H, Morris HH, et al. The lateralizing significance of versive 829. Nayyar M, Strobos RJ, Singh BM, et al. Caloric-induced nystagmus with isoelec- head and eye movements during epileptic seizures. Neurology 1986;36:606–611. tric electroencephalogram. Ann Neurol 1987;21:98–100. 966 CLINICAL NEURO-OPHTHALMOLOGY

830. Lempert T, von Brevern M. The eye movements of syncope. Neurology 1996; and recovery with thiamine in Wernicke syndrome. Am J Med Sci 2000;320: 46:1086–1088. 278–280. 831. Barnes D, McDonald WI. The ocular manifestations of multiple sclerosis: 2. 865. Mulder AH, Raemaekers JM, Boerman RH, Mattijssen V. Downbeat nystagmus Abnormalities of eye movements. J Neurol Neurosurg Psychiatry 1992;55: caused by thiamine deficiency: an unusual presentation of CNS localization of 863–888. large cell anaplastic CD 30-positive non-Hodgkin’s lymphoma. Ann Hematol 832. Frohman EM, Solomon D, Zee DS. Vestibular dysfunction and nystagmus in 1999;78:105–107. multiple sclerosis. Int J MS 1997;3:13–26. 866. Sharma S, Sumich PM, Francis IC, et al. Wernicke’s encephalopathy presenting 833. Gnanaraj L, Rao VJ. Partial unilateral third nerve palsy and bilateral internuclear with upbeating nystagmus. J Clin Neurosci 2002;9:476–478. ophthalmoplegia: an unusual presentation of multiple sclerosis. Eye 2000;14(Pt 867. Cogan DG, Witt ED, Goldman-Rakic PS. Ocular signs in thiamine-deficient 4):673–675. monkeys and in Wernicke’s disease in humans. Arch Ophthalmol 1985;103: 834. Frohman EM, Zhang H, Kramer PD, et al. MRI characteristics of the MLF in 1212–1220. MS patients with chronic internuclear ophthalmoparesis. Neurology 2001;57: 868. Witt ED, Goldman-Rakic PS. Intermittent thiamine deficiency in the rhesus 762–768. monkey. I. Progression of neurological signs and neuro-anatomical lesions. Ann 835. Frohman EM, O’Suilleabhain P, Dewey RB Jr, et al. A new measure of dysconju- Neurol 1983;13:376–395. gacy in INO: the first-pass amplitude. J Neurol Sci 2003;210:65–71. 869. Kenyon RV, Becker JT, Butters N, et al. Oculomotor function in Wernicke- 836. Aschoff JC, Conrad B, Kornhuber HH. Acquired pendular nystagmus with oscil- . I. Saccadic eye movements. Int J Neurosci 1984;25:53–65. lopsia in multiple sclerosis: a sign of cerebellar nuclei disease. J Neurol Neuro- 870. Kenyon RV, Becker JT, Butters N. Oculomotor function in Wernicke-Korsakoff surg Psychiatry 1974;37:570–577. syndrome. II. Smooth-pursuit eye movements. Int J Neurosci 1984;25:67–69. 837. Lopez LI, Bronstein AM, Gresty MA, et al. Clinical and MRI correlates in 27 871. Ramulu P, Moghekar A, Chaudhry V, et al. Wernicke’s encephalopathy. Neurol- patients with acquired nystagmus. Brain 1996;119:465–472. ogy 2002;59:846. 838. Meienberg O, Muri R, Rabineau PA. Clinical and oculographic examinations 872. Miranda AF, Ishii S, DiMauro S, et al. Cytochrome c oxidase deficiency in of saccadic eye movements in the diagnosis of multiple sclerosis. Arch Neurol Leigh’s syndrome: genetic evidence for a nuclear DNA-encoded mutation. Neu- 1986;43:438–443. rology 1989;39:697–702. 839. Frohman TC, Frohman EM, O’Suilleabhain P, et al. Accuracy of clinical detec- 873. DiMauro S, Servidei S, Zeviani M, et al. Cytochrome c oxidase deficiency in tion of INO in MS: corroboration with quantitative infrared oculography. Neurol- Leigh syndrome. Ann Neurol 1987;22:498–506. ogy 2003;61:848–850. 874. Adams PL, Lightowlers RN, Turnbull DM. Molecular analysis of cytochrome 840. Gresty MA, Findley LJ, Wade P. Mechanism of rotatory eye movements in c oxidase deficiency in Leigh’s syndrome. Ann Neurol 1997;41:268–270. opsoclonus. Br J Ophthalmol 1980;64:923–925. 875. Howard RO, Albert DM. Ocular manifestations of subacute necrotizing encepha- 841. Francis DA, Heron JR. Ocular flutter in suspected multiple sclerosis: a presenting lomyelopathy (Leigh’s disease). Am J Ophthalmol 1972;74:386–393. paroxysmal manifestation. Postgrad Med J 1985;61:333–334. 876. Dooling EC, Richardson EP Jr. Ophthalmoplegia and Ondine’s curse. Arch Oph- 842. Ashe J, Hain TC, Zee DS, et al. Microsaccadic flutter. Brain 1991;114:461–472. thalmol 1977;95:1790–1793. 843. Solingen LD, Baloh RW, Myers L, et al. Subclinical eye movement disorders 877. Blanco-Barca MO, Eiris-Punal J, Soler-Regal C, Castro-Gago M. [Duplication in patients with multiple sclerosis. Neurology 1977;27:614–619. of the PLP gene and the classical form of Pelizaeus-Merzbacher disease]. Rev 844. Reulen JPH, Sanders EACM, Hogenhuis LAM. Eye movement disorders in Neurol 2003;37:436–438. multiple sclerosis and . Brain 1983;106:121–140. 878. Trobe JD, Sharpe JA, Hirsh DK, et al. Nystagmus of Pelizaeus-Merzbacher 845. Milea D, Napolitano M, Dechy H, et al. Complete bilateral horizontal gaze disease: a magnetic search-coil study. Arch Neurol 1991;48:87–91. paralysis disclosing multiple sclerosis. J Neurol Neurosurg Psychiatry 2001;70: 879. Ouahchi K, Arita M, Kayden H, et al. Ataxia with isolated vitamin E deficiency 252–255. is caused by mutations in the alpha-tocopherol transfer protein. Nature Genet 846. Keane JR. Sustained upgaze in coma. Ann Neurol 1981;9:409–412. 1995;9:141–145. 847. Frohman EM, Solomon D, Zee DS. Vestibular dysfunction and nystagmus in 880. Yee RD, Cogan DC, Zee DS. Ophthalmoplegia and dissociated nystagmus in multiple sclerosis. Int J MS 1987;3:13–26. abetalipoproteinemia. Arch Ophthalmol 1976;94:571–575. 848. Downey DL, Stahl JS, Bhidayasiri R, et al. Saccadic and vestibular abnormalities 881. Hawkins RA, Mazziotta JC, Phelps ME. Wilson’s disease studied with FDG in multiple sclerosis: sensitive clinical signs of brainstem and cerebellar involve- and positron emission tomography. Neurology 1987;37:1707–1711. ment. Ann NY Acad Sci 2002;956:438–440. 882. Lennox G, Jones R. Gaze distractibility in Wilson’s disease. Ann Neurol 1989; 849. Frohman EM, Goodin DS, Calabresi PA, et al. The utility of MRI in suspected 25:415–417. MS: report of the Therapeutics and Technology Assessment Subcommittee of 883. Kirkham TH, Kamin DF. Slow saccadic eye movements in Wilson’s disease. J the American Academy of Neurology. Neurology 2003;61:602–611. Neurol Neurosurg Psychiatry 1974;37:191–194. 850. Starck M, Albrecht H, Po¨llmann W, et al. Drug therapy for acquired pendular 884. Keane JR. Lid-opening apraxia in Wilson’s disease. J Clin Neuroophthalmol nystagmus in multiple sclerosis. J Neurol 1997;244:9–16. 1988;8:31–33. 851. Hoyt CS. Nystagmus and other abnormal ocular movements in children. Pediatr 885. Kushner MJ, Parrish M, Burke A, et al. Nystagmus in motor neuron disease: Clin North Am 1987;34:1415–1423. clinicopathological study of two cases. Ann Neurol 1984;16:71–77. 852. Ahn JC, Hoyt WF, Hoyt CS. Tonic upgaze in infancy. A report of three cases. 886. Averbuch-Heller L, Helmchen C, Horn AK, et al. Slow vertical saccades in Arch Ophthalmol 1989;107:57–58. motor neuron disease: correlation of structure and function. Ann Neurol 1998; 853. Donat JFG, Donat JR, Lay KS. Changing response to caloric stimulation with 44:641–648. gestational age in infants. Neurology 1980;30:776–778. 887. Evdokimidis I, Constantinidis TS, Gourtzelidis P, et al. Frontal lobe dysfunction 854. Weissman BM, DiScenna AO, Leigh RJ. Maturation of the vestibulo-ocular in amyotrophic lateral sclerosis. J Neurol Sci 2002;195:25–33. reflex in normal infants during the first 2 months of life. Neurology 1989;39: 888. Vaphiades MS, Husain M, Juhasz K, Schmidley JW. Motor neuron disease 534–538. presenting with slow saccades and dementia. ALS Other Motor Neuron Disord 855. Harding AE, Young EP, Schon F. Adult onset supranuclear ophthalmoplegia, 2002;3:159–162. cerebellar ataxia, and neurogenic proximal muscle weakness in a brother and 889. Abel LA, Williams IM, Gibson KL, et al. Effects of stimulus velocity and sister: another hexosaminidase. A deficiency syndrome. J Neurol Neurosurg acceleration on smooth pursuit in motor neuron disease. J Neurol 1995;242: Psychiatry 1987;50:687–690. 419–424. 856. Neville BGR, Lake BD, Stephens R, et al. A neurovisceral storage disease with 890. Gizzi M, DiRocco A, Sivak M, et al. Ocular motor function in motor neuron vertical supranuclear ophthalmoplegia, and its relationship to Niemann-Pick dis- disease. Neurology 1992;42:1037–1046. ease. Brain 1973;96:97–120. 891. Okuda B, Yamamoto T, Yamasaki M, et al. Motor neuron disease with slow 857. Cogan DG, Chu FC, Reingold D, et al. Ocular motor signs in some metabolic eye movements and vertical gaze palsy. Acta Neurol Scand 1992;85:71–76. diseases. Arch Ophthalmol 1981;99:1802–1808. 892. Gruber H, Zeitlhofer J, Prager J, et al. Complex oculomotor dysfunctions in 858. Shawkat FS, Carr L, West P, et al. Vertical saccade palsy: a presenting sign of Kugelberg-Welander disease. Neuroophthalmology 1983;3:125–128. Niemann-Pick type IIS. Eur J Neurol 1994;1:93–95. 893. Thurston SE, Leigh RJ, Abel LA, et al. Slow saccades and hypometria in anticon- 859. Natowicz MR, Stoler JM, Prence EM, et al. Marked heterogeneity in Niemann- vulsant toxicity. Neurology 1984;34:1593–1596. Pick disease, type C: clinical and ultrastructural findings. Clin Ped 1995; 894. Remler BF, Leigh RJ, Osorio I, et al. The characteristics and mechanisms of 190–197. visual disturbances associated with anticonvulsant therapy. Neurology 1990;40: 860. Vivian AJ, Harris CM, Kriss A, et al. Oculomotor signs in infantile Gaucher 791–796. disease. Neuroophthalmology 1993;13:151–155. 895. Rascol O, Hain TC, Brefel C, et al. Antivertigo medications and drug-induced 861. Cogan DG, Victor M. Ocular signs of Wernicke’s disease. Arch Ophthalmol vertigo. Drugs 1995;50:777–791. 1954;51:204–211. 896. Padoan S, Korttila K, Magnusson M, et al. Effect of intravenous diazepam and 862. Ghez C. Vestibular paresis: a clinical feature of Wernicke’s disease. J Neurol thipental on voluntary saccades and pursuit eye movements. Acta Otolaryngol Neurosurg Psychiatry 1969;32:134–139. 1992;112:579–588. 863. Deramo VA, Jayamanne DG, Auerbach DB, Danesh-Meyer H. Acute bilateral 897. Cohen B, Helwig D, Raphan T, et al. Baclofen and velocity storage: a model ophthalmoplegia in a young woman. Surv Ophthalmol 2000;44:513–517. of the effects of the drugs on the vestibulo-ocular reflex in the monkey. J Physiol 864. Kumar PD, Nartsupha C, West BC. Unilateral internuclear ophthalmoplegia (Lond) 1987;393:703–725. SUPRANUCLEAR AND INTERNUCLEAR OCULAR MOTILITY DISORDERS 967

898. Pereira CB, Strupp M, Eggert T, et al. Nicotine-induced nystagmus: three-dimen- 920. Malm C, Lying-Tunell U. Cerebellar dysfunction related to toluene sniffing. sional analysis and dependence on head position. Neurology 2000;55: Acta Neurol Scand 1980;62:188–190. 1563–1566. 921. Lazar RB, Ho SU, Melen O, et al. Multifocal central nervous system damage 899. Pereira CB, Strupp M, Holzleitner T, Brandt T. Smoking and balance: correlation caused by toluene abuse. Neurology 1983;33:1337–1340. of nicotine-induced nystagmus and postural body sway. Neuroreport 2001;12: 922. Maas EF, Ashe J, Spiegel PS, et al. Acquired pendular nystagmus in toluene 1223–1226. addiction. Neurology 1991;41:282–285. 900. Sibony PA, Evinger C, Manning K, et al. Nicotine and tobacco-induced nystag- 923. Morata TC, Nyle´n P, Johnson A, et al. Auditory and vestibular functions after mus. Ann Neurol 1990;28:198. single or combined exposure to toluene: a review. Arch Toxicol 1995;69: 901. Booker JL. End-position nystagmus as an indicator of ethanol intoxication. Sci 431–443. Justice 2001;41:113–116. 924. Hunnewell J, Miller NR. Bilateral internuclear ophthalmoplegia and pendular 902. Citek K, Ball B, Rutledge DA. Nystagmus testing in intoxicated individuals. nystagmus in a patient with chronic toluene toxicity from glue-sniffing. Neurol- Optometry 2003;74:695–710. ogy, 1997. 903. Crevits L. Effect of alcohol on saccades. Clin Experiment Ophthalmol 2002; 925. Sibony PA, Evinger C, Manning KA. Tobacco-induced primary-position upbeat 30:450–451. nystagmus. Ann Neurol 1987;21:53–58. 904. Fetter M, Haslwanter T, Bork M, Dichgans J. New insights into positional alco- 926. Sibony PA, Evinger C, Manning KA. The effects of tobacco smoking on smooth- hol nystagmus using three-dimensional eye-movement analysis. Ann Neurol pursuit eye movements. Ann Neurol 1988;23:238–241. 1999;45:216–223. 927. Roos YBWEM, de Jongh FE, Crevits L. The effect of smoking on ocular sac- 905. Wegner AJ, Fahle M. Alcohol and visually guided saccades: gap effect and cadic latency time. Neuroophthalmology 1993;13:75–79. predictability of target location. Psychopharmacology (Berl) 1999;146:24–32. 928. Thaker GK, Ellsberry R, Moran M, et al. Tobacco smoking increases square- 906. Zingler VC, Strupp M, Krafczyk S, et al. Does alcohol cancel static vestibular wave jerks during pursuit eye movements. Biol Psychiatry 1991;29:82–88. compensation? Ann Neurol 2004;55:144–145. 929. Roos YBWEM, de Jongh FE, Crevits L. The effects of smoking on anti-saccades. 907. Dehaene I, Van Vleymen B. Opsoclonus induced by phenytoin and diazepam. Ann Neurol 1987;21:216. Neuroophthalmology 1995;15:3–8. 908. Flechtner K-M, Mackert A, Thies K, et al. Lithium effect on smooth-pursuit 930. Kim JI, Somers JT, Stahl JS, Bhidayasiri R, Leigh RJ. Vertical nystagmus in eye movements of healthy volunteers. Biol Psychiatr 1992;32:932–938. normal subjects: effects of head position, nicotine and scopolamine. J Vestib 909. Corbett JJ, Jacobson DM, Thompson HS, et al. Downbeating nystagmus and Res 2000;10:291–300. other ocular motor defects caused by . Neurology 1989;39: 931. Demer JL, Volkow ND, Ulrich I, et al. Eye movements in cocaine abusers. 481–487. Psychiat Res 1989;29:123–136. 910. Cannon SC, Robinson DA. Loss of the neural integrator of the oculomotor 932. Scharf D. Opsoclonus-myoclonus following the intranasal usage of cocaine. J system from brain stem lesions in monkey. J Neurophysiol 1987;5:1383–1409. Neurol Neurosurg Psychiatry 1989;52:1447–1448. 911. Schneider JA, Mirra SS. Neuropathologic correlate of persistent neurological 933. Elkardoudi-Pijnenburg Y, Van Vliet AGM. Opsoclonus, a rare complication of deficit in lithium intoxication. Ann Neurol 1994;36:928–931. cocaine misuse. J Neurol Neurosurg Psychiatry 1996;60:592. 912. Green JF, King DJ, Trimble KM. Antisaccade and smooth pursuit eye move- 934. Brandt T. Bilateral vestibulopathy revisited. Eur J Med Res 1996;1:361–368. ments in healthy subjects receiving sertraline and lorazepam. J Psychopharmacol 935. Halmagyi GM, Fattore CM, Curthoys IS, et al. Gentamicin vestibulotoxicity. 2000;14:30–36. Otolaryngol Head Neck Surg 1994;111:571–574. 913. Taylor JR, Selhorst JB, Houff SA, et al. Chlordecone intoxication in man. I. 936. Fishel-Ghodsian N, Prezant TR, Bu X, et al. Mitochondrial ribosomal RNA Clinical observations. Neurology 1978;28:626–630. gene mutation in a patient with sporadic aminoglycoside ototoxicity. J Otolaryn- 914. Maccario M, Seelinger D, Snyder R. Thallotoxicosis with coma and abnormal gol 1993;14:399–403. eye movements (opsoclonus): clinical and EEG correlations. Electroencephalogr 937. Prezant TR, Agapian JV, Bohlman MC, et al. Mitochondrial ribosomal RNA Clin Neurophysiol 1975;38:98–99. mutation associated with both antibiotic-induced and nonsyndromic deafness. 915. Hodgson MJ, Furman J, Ryan C, et al. Encephalopathy and vestibulopathy fol- Nature Genet 1993;4:289–294. lowing short-term hydocarbon exposure. J Occup Med 1989;31:51–54. 938. Longridge NS. Topical gentamicin vestibular toxicity. J Otolaryngol 1994;23: 916. O¨ dkvist LM, Mo¨ller C, Thuomas K. Otoneurologic disturbances caused by sol- vent polution. Otolarygol Head Neck Surg 1992;106:687–692. 444–446. 917. Larsby B, Tham R, Eriksson B, et al. I. Effects of trichloroethylene on the human 939. Kitsigianis G-A, O’Leary DP, Davis LL. Active head-movement analysis of vestibulo-oculomotor system. Acta Orolaryngol (Stockh) 1986;101:193–199. cisplatin-induced vestibulotoxicity. Otolaryngol Head Neck Surg 1988;98: 918. Moller C, O¨ dkvist LM, Thell J, et al. Otoneurological findings in psycho-organic 82–87. syndrome caused by industrial solvent exposure. Acta Otolaryngol (Stockh) 940. Myers SF, Blakely BW, Schwan S. Is cis-platinum vestibulotoxic? Otolaryngol 1989;107:5–12. Head Neck Surg 1993;108:322–328. 919. Rosenberg NL, Kleinschmidt-DeMasters BJ, Davis KA, et al. Toluene abuse 941. Nakayama M, Riggs LC, Matz GJ. Quantitative study of vestibulotoxicity in- causes diffuse central nervous system white matter changes. Ann Neurol 1988; duced by gentamicin or cisplatin in the guinea pig. Laryngoscope 1996;106: 23:611–614. 162–167.