Chapter 10
Supranuclear Disorders of Eye Movements
R. JOHN LEIGH and ROBERT B. DAROFF
Main Menu   Table Of Contents

Search

INTRODUCTION
BRAINSTEM CONTROL OF HORIZONTAL GAZE
BRAINSTEM CONNECTIONS FOR VERTICAL AND TORSIONAL MOVEMENTS
BRAINSTEM CONTROL OF VERGENCE EYE MOVEMENTS
CEREBELLAR INFLUENCES ON EYE MOVEMENTS
CEREBRAL AND BASAL GANGLIONIC CONTROL OF EYE MOVEMENTS
REFERENCES

INTRODUCTION
A good deal is now known about the control of gaze, from the level of genes and molecules up to the level of how populations of neurons control ocular motor behaviors.1–3 In this chapter, we attempt to summarize this body of knowledge into a scheme that can be applied at the bedside. An important starting point is a systematic examination of each functional class of eye movements (Table 1). Eye movements evolved to serve vision, and each functional class has distinctive properties that suit it to a specific purpose. Since the neurobiologic substrate of each functional class has been identified, specific disorders of eye movements often aid topological diagnosis. The approach here will be “bottom-up,” starting with the brainstem inputs to the ocular motoneurons that reside in the third, fourth, and sixth cranial nerve nuclei. Subsequently, the contributions of the cerebellum and cerebral hemispheres are reviewed. At each stage in this journey from ocular motoneuron to visual cortex, a summary of the relevant anatomy and physiology is followed by descriptions of the effects of discrete lesions and the characteristics of distinctive syndromes.

TABLE 1. Functional Classes Of Eye Movements

Class of eye movementMain function
Visual fixationHolds the image of a stationary object on the fovea
Vestibulo-ocular reflexHolds images of the seen world steady on the retina during brief head rotations
OptokineticHolds images of the seen world steady on the retina during sustained head rotations
Smooth pursuitHolds the image of a moving target close to the fovea
Nystagmus quick phasesReset the eyes during prolonged rotation, and direct gaze toward the oncoming visual scene
SaccadesBring images of objects of interest onto the fovea
VergenceMoves the eyes in opposite directions so that images of a single object are placed simultaneously on both foveas

Adapted from Leigh RJ, Zee DS: The Neurology of Eye Movements, 3rd ed. Oxford University Press, New York, 1999.)

 

Prior to starting on this tour, we must recognize that when the brain initiates any type of eye movement, it must take into account the properties of the eyeball, its suspensory ligaments, and fascia. Although it is easy to take for granted that the brain will make adjustments, for example, when the eye turns to look out of the corner of the orbit, such eccentric “gaze-holding” demands an appropriate tonic contraction of the extraocular muscles, to prevent the eye from drifting back toward primary position. This gaze-holding function (sometimes referred to as the achievement of a neural integrator) is really a separate functional property affecting all eye movements that also has its own identified neural substrate.

A recent revolution in the field has been the rediscovery of pulleys that guide the tendons of the extraocular muscles.4 Thus, each muscle has an outer orbital layer that inserts into the muscles pulley; the inner global layer passes through the pulley to insert on the globe. With the exception of the superior oblique, all muscle pulleys move, which may be how Listing's law is imposed on eye rotations. Listing's law states that any eye position can be reached from primary position by rotation of the eye about a single axis lying in the equatorial plane. By delegating control of the geometry of eye rotations to the muscles and pulleys, the brain seems to simplify its role in controlling three dimensional eye rotations. Different populations of ocular motoneurons may supply the global and orbital layers, and their proprioception may differ5; these are topics of current research.

Back to Top
BRAINSTEM CONTROL OF HORIZONTAL GAZE

MOTOR COMMANDS FOR CONJUGATE (VERSIONAL) HORIZONTAL MOVEMENTS

The abducens nucleus is of central importance in the control of horizontal gaze because it governs conjugate movements of both the ipsilateral lateral rectus and the contralateral medial rectus muscles. It houses two populations of neurons: (1) abducens motoneurons, which supply the lateral rectus muscle and (2) abducens internuclear neurons, which project up the contralateral medial longitudinal fasciculus (MLF) to contact medial rectus motoneurons of the oculomotor nucleus (Fig. 1).2,6 Thus, axons of the abducens nerve and axons of the abducens internuclear neurons that run in the MLF,7 together encode conjugate horizontal eye movements, and the mechanism for yoking of horizontal movements—Hering's law—has its neural basis in the abducens nucleus and its two populations of neurons. The abducens motoneurons and internuclear neurons show only minor morphological differences,8 but only the motoneurons contain acetylcholine.9 Both types of abducens neurons receive the same afferent input.

Fig. 1. Anatomic scheme for the synthesis of signals for horizontal eye movements. The abducens nucleus (CN VI) contains abducens motoneurons that innervate the ipsilateral lateral rectus muscle (LR), and abducens internuclear neurons that send an ascending projection in the contralateral medial longitudinal fasciculus (MLF) to contact medial rectus (MR) motoneurons in the contralateral third nerve nucleus (CN III). From the horizontal semicircular canal, primary afferents on the vestibular nerve project mainly to the medial vestibular nucleus (MVN), where they synapse and then send an excitatory connection to the contralateral abducens nucleus and an inhibitory projection to the ipsilateral abducens nucleus. Saccadic inputs reach the abducens nucleus from ipsilateral excitatory burst neurons (EBN) and contralateral inhibitory burst neurons (IBN). Eye position information (the output of the neural integrator) reaches the abducens nucleus from neurons within the nucleus prepositus hypoglossi (NPH) and adjacent MVN. The medial rectus motoneurons in CN III also receive a command for vergence eye movements. Putative neurotransmitters for each pathway are shown: Ach: acetylcholine; asp: aspartate; glu: glutamate; gly: glycine. CN VI: abducens nerve; MVN: medial vestibular nucleus. (Adapted from Leigh RJ, Zee DS: The Neurology of Eye Movements, 3rd ed. Oxford University Press, New York, 1999.)

How do signals for each functional class of eye movement (Table 1) project to the abducens nucleus? The abducens nucleus receives vestibular and optokinetic inputs from the vestibular nuclei.10 Saccadic commands originate from burst neurons of the pontine and medullary reticular formation.11 A descending smooth pursuit pathway probably projects to the abducens nucleus via the vestibular and cerebellar fastigial nuclei (see Pursuit System, below).2 The signals that are important for gaze-holding (neural integrator) function reach the abducens nucleus from the nucleus prepositus hypoglossi and the medial vestibular nuclei.12

EFFECTS OF LESIONS OF THE PONS AND MEDULLA ON HORIZONTAL GAZE

Lesions of the abducens nucleus produce paralysis of both the ipsilateral lateral rectus and contralateral medial rectus for all conjugate eye movements,13–16 but vergence movements are spared. Clinical lesions that affect only the abducens nucleus are rare, and usually there is also involvement of adjacent structures including the MLF, paramedian pontine reticular formation (PPRF), and particularly, the facial nerve fascicle.

Lesions of the medial longitudinal fasciculus produce internuclear ophthalmoplegia (INO), which is characterized by paresis of adduction for conjugate movements on the side of the lesion.2,17 Another common feature of INO is dissociated nystagmus, characterized by abduction overshoot of the eye contralateral to the lesion (see Chapter 11). Dissociated nystagmus is probably an adaptive phenomenon; as the brain attempts to point the paretic adducting eye at the visual target, a series of saccades is only evident in the non-paretic, abducting eye.18 The MLF is paramedian in the pons and becomes slightly more lateral in the midbrain as it approaches the oculomotor nuclear complex. INO is often accompanied by skew deviation, probably because of disruption of the otolithic-ocular connections.19 In addition, the vertical vestibulo-ocular reflex (VOR) may show deficient downward eye movements when tested with “head impulses,” because the INO carries some but not all vertical vestibular signals.20 Bilateral INO is usually associated with gaze-evoked vertical nystagmus, impaired vertical pursuit, and decreased vertical vestibular responses.21 Small-amplitude saccadic intrusions may interrupt fixation.22 The most frequent cause of INO in young adults, particularly when bilateral, is multiple sclerosis; other causes are summarized in Table 2.

TABLE 2. Causes Of Internuclear Ophthalmoplegia

  1. Multiple sclerosis (commonly bilateral)
  2. Brainstem infarction (commonly unilateral), including complication of arteriography and hemorrhage
  3. Brainstem and fourth ventricular tumors
  4. Arnold-Chiari malformation, and associated hydrocephalus and syringobulbia
  5. Infection: bacterial, viral and other forms of meningo-encephalitis; in association with AIDS
  6. Hydrocephalus, subdural hematoma, supratentorial arteriovenous malformation
  7. Nutritional disorders: Wernicke's encephalopathy and pernicious anemia
  8. Metabolic disorders: hepatic encephalopathy, maple syrup urine disease, abetalipoproteinemia, Fabry's disease
  9. Drug intoxications: phenothiazines, tricyclic antidepressants, narcotics, propranolol, lithium, barbiturates
  10. Cancer: either the result of carcinomatous infiltration, or remote effect
  11. Head trauma, and cervical hyperextension or manipulation
  12. Degenerative conditions: progressive supranuclear palsy
  13. Syphilis
  14. Pseudo-internuclear ophthalmoplegia of myasthenia gravis, and Fisher's syndrome

AIDS, acquired immune deficiency syndrome.
Adapted from Leigh RJ, Zee DS: The Neurology of Eye Movements, 3rd ed. Oxford University Press, New York, 1999.)

 

Strictly speaking, the term internuclear ophthalmoplegia should refer to absolute adduction paralysis during horizontal versions, and the term internuclear ophthalmoparesis should be used to describe those cases where adduction past the midline is present, but with limitation of amplitude or decreased velocity. When INO is subtle, two maneuvers may be useful for demonstrating it: the optokinetic and ocular dysmetria signs,23 which rely on demonstrating impairment of innervation of the medial rectus muscle compared with its yoke, the contralateral lateral rectus, during horizontal saccades. With the optokinetic tape moving to the side of the involved medial rectus muscle, the nystagmus response in that eye is reduced (especially the adduction saccade) compared with the opposite eye. The ocular dysmetria sign necessitates repetitive horizontal saccadic refixations that disclose slow, hypometric refixation of the adducting eye (medial rectus) and concomitant overshoot of the abducting eye (lateral rectus). However, recent studies indicate that even experienced clinicians may miss INO,24 and measurement of saccades to compare adducting and abducting movements is worthwhile when MS is suspected.25–27 Magnetic resonance imaging (MRI) with proton-density high-resolution views of the brainstem may also aid in establishing MLF lesions.28

Most cases of INO are orthotropic (or exophoric) in primary position without symptomatic diplopia unless accompanied by skew deviation. Occasionally, patients with bilateral INO show exotropia, designated the WEBINO (wall-eyed bilateral INO) syndrome. The explanation for this finding is unclear because monkeys with a lidocaine-induced internuclear ophthalmoplegia show an increased accommodative vergence to accommodation ratio, implying that the MLF actually carries signals that inhibit vergence.17 Thus, the cause of exotropia in INO is uncertain and we have observed it in patients with preserved convergence. Furthermore, the classification of INO into anterior and posterior types depending on the integrity of the convergence mechanism is not of localizing value in identifying the rostral-caudal location of the MLF lesion. The so-called posterior internuclear ophthalmoplegia of Lutz,29 in which abduction (but not adduction) is impaired is rare. It has been difficult to conceptualize because the abducens contains neurons that innervate both lateral and medial rectus muscles for all conjugate eye movements.30 One possible explanation concerns oculomotor internuclear neurons, which project from the third nerve nucleus down to the contralateral abducens nucleus; inactivation of these internuclear neurons causes impaired abduction.31

A combined lesion of one MLF and the adjacent abducens nucleus or PPRF produces paralysis of all conjugate movements except for abduction of the eye contralateral to the side of the lesion—”one-and-a-half” syndrome.32–34 This occurs with brainstem infarction, hemorrhage, multiple sclerosis, and pontine glioma; rarer causes are brainstem arteriovenous malformation, basilar artery aneurysm, and posterior fossa tumor. When the lesion is acute, the patient may be profoundly exotropic.35 The deviated eye may demonstrate marked nystagmus and is always on the side opposite that of the brainstem lesion.

Discrete lesions of the paramedian pontine reticular formation (PPRF)36,37—which mainly corresponds to the nucleus pontis centralis caudalis and contains saccadic burst neurons—cause loss of saccades and quick phases of nystagmus to the side of the lesion. Vertical saccades are misdirected obliquely away from the side of the lesion.38 Selective impairment of horizontal saccades occurs in degenerative conditions, such as some variants of olivopontocerebellar atrophy.39 Other causes of slow or absent saccades are summarized in Table 3. Infarction of the paramedian pons usually but not always also involves adjacent fibers conveying vestibular and pursuit inputs to the abducens nucleus.38 Pontine disease may cause a unilateral defect of smooth pursuit by affecting the dorsolateral pontine nuclei and their projections to the cerebellum (see below). More rostral brainstem lesions may cause ipsilateral smooth pursuit deficits, whereas caudal brainstem lesions tend to cause contralateral deficits2; this is discussed further below.

TABLE 3. Causes Of Slow Saccades

  1. Olivopontocerebellar atrophy and spinocerebellar degenerations
  2. Huntington's disease
  3. Progressive supranuclear palsy
  4. Parkinson's disease (advanced cases) and diffuse Lewy body disease
  5. Whipple's disease
  6. Lipid storage diseases
  7. Wilson's disease
  8. Drug intoxications: anticonvulsants, benzodiazepines
  9. Tetanus
  10. In dementia: Alzheimer's disease (stimulus-dependent) and in association with AIDS
  11. Lesions of the paramedian pontine reticular formation
  12. Internuclear ophthalmoplegia
  13. Peripheral nerve palsy, diseases affecting the neuromuscular junction and extraocular muscle, restrictive ophthalmopathy

AIDS, acquired immune deficiency syndrome.
Adapted from Leigh RJ, Zee DS. The Neurology of Eye Movements, 3rd ed. Oxford University Press, New York, 1999.)

 

Bilateral experimental lesions of the nucleus prepositus hypoglossi and adjacent medial vestibular nucleus abolish the gaze-holding mechanism (neural integrator) for eye movements in the horizontal plane.40,41 In health, the neural integrator generates the eye position command necessary to hold the eye steady in an eccentric orbital position. After damage to the neural integrator network (which includes connections with the vestibular cerebellum), the eye cannot be held in an eccentric position in the orbit and it drifts back to primary position. Corrective quick phases then produce gaze-evoked nystagmus (see Chapter 11). Thus, disease affecting the vestibular nuclei and nucleus prepositus hypoglossi cause vestibular imbalance—manifest as nystagmus or skew deviation—and impairment of gaze-holding. This occurs with lateral medullary infarction (Wallenberg's syndrome) in which the spontaneous nystagmus is usually horizontal or mixed horizontal-torsional with the slow phases directed toward the side of the lesion; the nystagmus may reverse direction in eccentric positions, suggesting coexistent involvement of the gaze-holding mechanism. Skew deviation, with an ipsilateral hypotropia, cyclodeviation (lower eye is more extorted), and ipsilateral head tilt (when present together, called the ocular tilt reaction) reflect an imbalance of otolithic inputs.42 In addition, patients with Wallenberg's syndrome show a characteristic lateropulsion in which the eyes deviate conjugately toward the side of the lesion if the lids are closed or with saccades.43 A hypothetical explanation for lateropulsion is that infarction of the inferior cerebellar peduncle leads to increased inhibition of the fastigial nucleus by the cerebellar vermis44; this is discussed further below in the section dealing with cerebellar influences on eye movements.

Back to Top
BRAINSTEM CONNECTIONS FOR VERTICAL AND TORSIONAL MOVEMENTS
Vertical and torsional saccadic commands and the vertical gaze-holding signal are synthesized in the midbrain, while vestibular and pursuit signals ascend to the midbrain from the lower brainstem. The ocular motoneurons that control vertical and torsional eye movements lie in the oculomotor and trochlear nuclei. How do signals for each functional class of eye movements project to these motoneurons?

In the prerubral fields of the mesencephalon, rostral to the tractus retroflexus and caudal to the mammillothalamic tract, lies a nucleus that is important for the generation of vertical saccades (Fig. 2). This structure, now called the rostral interstitial nucleus of the medial longitudinal fasciculus (riMLF),45,46 previously designated the nucleus of the prerubral fields, and the nucleus of the fields of Forel, contains burst neurons for vertical saccades and quick phases, and torsional quick phases. Each riMLF contains neurons that burst for upward and downward eye movements but for torsional quick phases in only one direction. Thus, the right riMLF discharges for quick phases that are directed clockwise with respect to the subject.47 In addition, each riMLF is connected to its counterpart by the posterior commissure, and by a more rostrally located ventral commissure that lies ventral to the aqueduct. The riMLF projects to motoneurons innervating elevator muscles bilaterally but to motoneurons innervating depressor muscles ipsilaterally.48,49 Furthermore, each burst neuron in the riMLF appears to send axon collaterals to motoneurons supplying yoke muscle pairs (e.g., superior rectus and inferior oblique); this is the neural substrate whereby Hering's law applies to saccades in the vertical plane.

Fig. 2. Anatomic schemes for the synthesis of upward, downward, and torsional eye movements. From the vertical semicircular canals, primary afferents on the vestibular nerve (vn) synapse in the vestibular nuclei (VN) and ascend in the medial longitudinal fasciculus (MLF) and brachium conjunctivum (not shown) to contact neurons in the trochlear nucleus (CN IV), oculomotor nucleus (CN III), and the interstitial nucleus of Cajal (INC). For clarity, only excitatory vestibular projections are shown. The rostral interstitial nucleus of the medial longitudinal fasciculus (riMLF), which lies in the prerubral fields, contains saccadic burst neurons. It receives an inhibitory input from omnipause neurons of the nucleus raphe interpositus (rip), which lie in the pons (for clarity, this projection is only shown for upward movements). Excitatory burst neurons in riMLF project to the motoneurons of CN III and CN IV, and also send an axon collateral to INC. Each riMLF neuron sends axon collaterals to yoke-pair muscles (Hering's law). Projections to the elevator subnuclei (innervating the superior rectus and inferior oblique muscles) may be bilateral due to axon collaterals crossing at the level of the CN III nucleus. Projections of inhibitory burst neurons are less well understood, and are not shown. The INC provides a gaze-holding signal, and projects to vertical motoneurons via the posterior commissure. Signals contributing to vertical smooth pursuit and eye-head tracking reach CN III from the y-group via the brachium conjunctivum and a crossing ventral tegmental tract. Neurotransmitters: asp = aspartate; glu = glutamate; gly = glycine. (Adapted from Leigh RJ, Zee DS: The Neurology of Eye Movements, 3rd ed. Oxford University Press, New York, 1999.)

The riMLF also projects to the interstitial nucleus of Cajal (INC), which has an important role in vertical gaze-holding (the neural integrator).50 In addition, the INC also receives inputs from the vestibular nuclei. The INC projects to motoneurons of the vertical ocular motor subnuclei. The projections of the INC to the elevator ocular motor nuclei (superior rectus and inferior oblique) pass dorsally over the aqueduct in the posterior commissure and this might explain why lesions of the posterior commissure limit mainly upward eye movements.50–52 The INC also contains neurons that project to motoneurons of the neck and trunk muscles and appears to coordinate combined torsional-vertical movements of the eyes and head.53 The neural signals necessary for vertical vestibular and smooth pursuit eye movements ascend from the medulla and pons to the midbrain. The MLF is the most important route for these projections but the brachium conjunctivum and other pathways are also involved.54

EFFECTS OF MIDBRAIN LESIONS ON VERTICAL AND TORSIONAL GAZE

Unilateral, experimental lesions of the riMLF cause a mild defect in vertical saccades, because each nucleus contains burst neurons for both upward and downward movements. On the other hand, unilateral riMLF lesions produce a specific defect of torsional quick phases.47 For example, with a lesion of the right riMLF, torsional quick phases, clockwise from the point of view of the subject (extorsion of the right eye and intorsion of the left eye) are lost. Vertical saccadic deficits with unilateral lesions of the riMLF in humans are rare; such deficits probably reflect involvement of the commissural pathways of the riMLF that makes the lesion, in effect, bilateral.55 In general, bilateral lesions are required to produce clinically apparent deficits of vertical eye movements.56 Bilateral experimental lesions of the riMLF in monkeys cause a vertical saccadic deficit, that may be more pronounced for downward eye movements57; this may be because upward saccadic projections are bilateral whereas downward saccadic projections are unilateral.48,49,58 With bilateral riMLF lesions, vertical gaze-holding, vestibular eye movements and possibly pursuit are preserved, as are horizontal saccades, with the deficits limited to either downward or both upward and downward saccades.58,59 Certain metabolic and degenerative disorders may lead to selective slowing or absence of vertical saccades (Table 3).

Unilateral experimental lesions of the INC are reported to impair gaze-holding function in the vertical plane.60 In addition, there is skew deviation (ipsilateral hypertropia), extorsion of the contralateral eye, intorsion of the ipsilateral eye, and contralateral head tilt.19 Stimulation near the INC in the monkey produces an ocular tilt reaction that consists of an ipsilateral head tilt and a synkinetic ocular reaction: depression and extorsion of the eye ipsilateral to the stimulation and elevation and intorsion of the contralateral eye.61,62 This ocular tilt reaction is similar to that produced by stimulation of the contralateral utricular nerve or an ipsilateral lesion of the vestibular nucleus (see Wallenberg's syndrome, above). Bilateral lesions of the medial longitudinal fasciculus (bilateral INO) impair vertical vestibular and pursuit movements but spare vertical saccades.63 In addition, partial loss of the vertical eye position signal causes vertical gaze-evoked nystagmus.

Lesions of the posterior commissure cause loss of upward gaze52; usually all types of eye movement are affected, although the VOR and Bell's phenomenon may sometimes be spared. Experimental inactivation of the posterior commissure with lidocaine impairs vertical gaze-holding (neural integrator) function.51 Other findings with posterior commissure lesions include slowing of vertical saccades below the horizontal meridian, disorders of convergence, convergence-retraction nystagmus, light-near dissociation of the pupils, and eyelid retraction (see Dorsal Midbrain Syndrome, below).

SUSTAINED VERTICAL DEVIATIONS

A number of disparate conditions may lead to tonic upward or downward eye deviations. One dramatic example is the sustained upward deviation of oculogyric crisis (see below, under parkinsonian disorders). Otherwise, tonic up-gaze deviation is seen in unconscious patients, especially after hypoxic-ischemic brain injury, perhaps reflecting loss of cerebellar Purkinje cells that normally balance vestibular and gaze-holding mechanisms. This notion is supported by the observation that patients who survive such insults develop downbeating nystagmus.64 Tonic downward deviation may occur transiently in neonates and does not necessarily indicate a neurologic abnormality; in such cases, the eyes can be easily driven above the horizontal meridian by the vertical doll's-head maneuver.65 In comatose patients, sustained downward deviations, with small unreactive pupils, usually reflect bilateral thalamic infarction or hemorrhage.66 The setting sun sign of infantile hydrocephalus is an analogous clinical phenomenon, with associated pathologic eyelid retraction. The reversibility of this sign with ventricular decompression indicates a dynamic mechanism secondary to acutely increased intracranial pressure.67 Tonic downward deviation may occur with lesions that compress the posterior commissure but also in metabolic encephalopathies.68 These sustained downward deviations are to be distinguished from “ocular bobbing,” an oscillatory disorder consisting of an abrupt downward jerk followed by a slow upward drift to midposition (see Chapter 11). Slow, spontaneously alternating skew deviations may occur with lesions of the dorsal midbrain.69

DORSAL MIDBRAIN SYNDROME

Dorsal lesions in the rostral midbrain produce a distinctive constellation of neuro-ophthalmologic signs involving supranuclear control of vertical gaze, eyelids, pupils, accommodation, and vergence. This entity encompasses many eponymic and anatomical designations, such as Parinaud, Koerber-Salus-Elschnig, sylvian aqueduct, pretectal, and posterior commissural syndromes.70 Pineal area tumors71 and midbrain infarction58,59 are the most common causes. Other etiologies include congenital aqueductal stenosis, multiple sclerosis, syphilis, arteriovenous malformations, midbrain hemorrhage,72 encephalitis, midbrain or third ventricle tumors, herniation of the uncus,73 and lesions resulting from stereotactic surgery.74 The lesions are usually large and involve one or more of the sites previously described as causing up-gaze palsies. The pupils are mid-position or large and demonstrate light-near dissociation; that is, the light reaction is smaller than the near response. Pathologic lid retraction (Collier's sign) and lid lag are common findings. Paralysis of upward gaze is the hallmark sign. Upward saccades are affected initially, whereas pursuit is relatively spared. Both types of eye movements, as well as vestibulo-ocular responses and Bell's phenomenon, may ultimately become paralyzed. Attempts at upward saccades evoke convergence-retraction nystagmus that persists as long as the refixation effort is maintained. The nystagmus is best elicited with down-going optokinetic targets providing a stimulus for repetitive upward saccades; each fast phase is replaced by a convergence or retraction movement. There is debate as to whether it is a true form of nystagmus,75 or a series of opposed adducting saccades, followed by slow divergence movements (see Chapter 11).76 This is in contrast to normal subjects who show transiently convergence during downward saccades and divergence during upward saccades.77 Rarely, a divergence-retraction nystagmus may occur in patients with dorsal midbrain syndrome.

Cerebral hemispheric lesions associated with focal motor seizures may cause retraction nystagmus temporally associated with periodic lateralized epileptiform discharges (PLEDs) on the electroencephalogram (EEG).78

SKEW DEVIATION

Skew deviation may be defined as a vertical misalignment of the visual axes caused by a disturbance of prenuclear inputs. The misalignment may be the same in all positions of gaze (concomitant), may vary with gaze position, or may even alternate (e.g., right hypertropia on right gaze, left hypertropia on left gaze).79–81 When skew deviation is nonconcomitant, it should be differentiated from vertical extraocular muscle palsy by the coexistence of signs of central neurologic dysfunction, and from trochlear nerve palsy by a negative head-tilt test. In some patients, skew deviation is accompanied by ocular torsion and head tilt—the ocular tilt reaction (OTR)19 (Fig. 3), which may be tonic (sustained)82,83 or paroxysmal.84,85 Such patients also show a deviation of the subjective vertical.19Rarely, skew deviation slowly alternates or varies in magnitude over the course of minutes.69 These patients have midbrain lesions. The periodicity of the phenomena is reminiscent of periodic alternating nystagmus (see Chapter 11) and the two phenomena may coexist.86

Fig. 3. Graviceptive pathways from the otoliths and vertical semicircular canals mediating the vestibular reactions in the roll plane. The projections from the otoliths and the vertical semicircular canals to the ocular motor nuclei (trochlear nucleus IV, oculomotor nucleus III, abducens nucleus VI) and the supranuclear centers of the interstitial nucleus of Cajal (INC), and the rostral interstitial nucleus of the medial longitudinal fasciculus (riMLF) are shown. The subserve the vestibulo-ocular reflex (VOR) in three planes. The VOR is part of a more complex vestibular reaction that also involves the vestibulospinal connections via the medial and lateral vestibulospinal tracts for head and body posture control. Furthermore, connections to the assumed vestibular cortex (area 2v and 3a and the parietoinsular vestibular cortex, PIVC) via the vestibular nuclei of the thalamus (Vim, Vce) are depicted. Graviceptive vestibular pathways for the roll plane cross at the pontine level. Ocular tilt reaction (OTR) is depicted schematically on the right in relation to the level of the lesion (i.e., ipsiversive OTR with peripheral and pontomedullary lesions) contraversive OTR with pontomesencephalic lesions. In vestibular thalamus lesions, the tilts of the subjective visual vertical may be contraversive or ipsiversive; in vestibular cortex lesions, they are preferably contraversive. (From Brandt T, Dieterich M: Vestibular syndromes in the roll plane: topographic diagnosis from brainstem to cortex. Ann Neurol 1994;36:337–347.)

Skew deviation is seen with a variety of disorders of the brainstem and cerebellum,79 and as a reversible finding with raised intracranial pressure. Current evidence suggests that skew deviation occurs whenever peripheral or central lesions cause an imbalance of otolithic inputs.19 Skew deviation also occurs with cerebellar lesions.81 The characteristics of the ocular tilt reaction caused by lesions of the vestibular labyrinth, vestibular nuclei, MLF, and the INC are summarized in Figure 10-3. In cases when the skew deviation is paroxysmal, the mechanism is thought to be irritative. This interpretation is supported by observations in one patient; stimulation of in the region of INC caused an ocular tilt reaction, with episodes of contralateral hypertropia and ipsilateral head tilt.62

Back to Top
BRAINSTEM CONTROL OF VERGENCE EYE MOVEMENTS

MOTOR COMMANDS FOR VERGENCE MOVEMENTS

Neurophysiologic studies in monkeys have shown that almost all oculomotor neurons subserving the medial rectus and the majority of neurons in the abducens nucleus discharge for both conjugate (version) and disjunctive (vergence) eye movements, although there is evidence that different neurons play relatively smaller or larger roles in conjugate versus vergence eye movements.87–90 Premotor commands for vergence are found on neurons in the mesencephalic reticular formation, 1 to 2 mm dorsal and dorsolateral to the oculomotor nucleus.89,91 Three main types of neurons are present; those that discharge in relation to vergence angle, vergence velocity,92 and both vergence angle and velocity. Evidence suggests that the nucleus reticularis tegmenti pontis (NRTP) houses neurons that are important for generating the vergence position signal (the vergence integrator).93 Abducens and oculomotor interneurons (each of which has projections to the other nucleus via the MLF) contribute to the coordination of conjugate and vergence commands.17,31

Most natural vergence movements are made in combination with a saccade, as we shift our fixation point between visual targets located at different distances and in different directions.77 Clinicians traditionally test vergence with a smooth movement of a visual target that is brought along the patient's midsagittal plane toward the nose. These two types of vergence—saccadic vergence and pursuit vergence—depend, in part, on different anatomic circuits.2

DISORDERS OF VERGENCE

Midbrain lesions may disrupt convergence, accommodation, and pupillary constriction—the near reflex. In addition, they produce vertical gaze disturbances and so-called convergence-retraction nystagmus (see Dorsal Midbrain Syndrome, above). Convergence paralysis or paresis may be secondary to various organic processes such as encephalitis, multiple sclerosis, and occlusive vascular disease involving the rostral midbrain. The distinction between organic and functional convergence palsy is based primarily on the assumption that accommodation and pupillary constrictions are provoked during attempted convergence in the former, but not in the latter. Senescence and lack of effort are the most common causes of poor convergence. Overall, the most common disorders of vergence are congenital, causing strabismus in childhood. In many such cases, abnormality of the accommodative-convergence synkinesis can be demonstrated. In adults, the commonest disorders are spasm of the near reflex and divergence paresis (see below).

Testing of vergence eye movements has focused on movements that are either made with a saccade, or in response to a smoothly target moving in the depth plane. Vergence movements made with saccades may be slow in progressive supranuclear palsy (PSP; see Parkinsonian Disorders, below),94 whereas patients with lesions involving the nucleus reticularis tegmenti pontis may have selective impairment of smooth vergence movements.95 Applying this approach with reliable measurements of eye movements, is likely to provide new information on disorders of vergence.

SPASM OF THU NEAR REFLEX

Spasm of the near reflex (accommodative or convergence spasm) is most commonly psychogenic. Affected patients may appear to have unilateral or bilateral abducens paresis.96 However, the pupillary miosis accompanying failure of abduction is a telltale sign. Volitional excessive near effort is uncomfortable and rarely can be maintained for more than seconds to minutes. Symptoms include blurred vision, eye strain, dizziness, headache, and diplopia. Rarely, a tonic near reflex may occur with organic disease, including generalized seizures, head trauma, the dorsal midbrain syndrome, craniocervical junction abnormalities, intoxications, and Wernicke's encephalopathy.2,97

DIVERGENCE PARESIS OR PARALYSIS

Paresis (insufficiency) of divergence is a rare clinical syndrome characterized by orthophoria at near but relatively concomitant esotropia at distance, although abduction is normal. Diplopia while viewing distant objects is the principal complaint, although other vague symptoms may be emphasized. A prerequisite for the diagnosis of divergence paresis is the demonstration of normal abducting saccades, and this criteria has seldom been met. In many patients, the question of decompensation of a long-standing esophoria bias is usually raised, for example, in the setting of concussion or cervical hyperextension injury. Tests of vergence amplitudes and stereopsis may be equivocal and the possibility of mild abduction defects remains. Mild abducens nerve dysfunction occurs after head trauma or raised intracranial pressure. The possibility of raised intracranial pressure must be excluded. Divergence insufficiency has been reported in PSP, seizure disorder, acute brainstem dysfunction, vascular disease, and after viral infections.98,99 Most patients do well with base-out prisms incorporated in distant spectacle correction and only rarely is ocular muscle surgery required.

Back to Top
CEREBELLAR INFLUENCES ON EYE MOVEMENTS
The cerebellum optimizes eye movements enabling them to provide clear and stable vision. Two separate parts of the cerebellum play an important role in the control of eye movements: the vestibulocerebellum (flocculus, paraflocculus, nodulus, and ventral uvula) and the dorsal vermis of the posterior lobe, and the fastigial nuclei.

The flocculi and ventral paraflocculi (tonsils), are paired structures that, in the human brain, lie ventral to the inferior cerebellar peduncle and next to the eighth cranial nerve. The flocculus receives bilateral inputs from the vestibular nuclei and the nucleus prepositus hypoglossi, and inputs from the dorsal cap of the contralateral inferior olive.100 The main efferent pathways of the flocculus are to the ipsilateral superior, medial, and y-group of vestibular nuclei.101 The flocculus and ventral paraflocculus contain Purkinje cells that discharge during smooth pursuit and combined eye-head tracking to encode gaze-velocity.102 These floccular Purkinje cells may also contribute to vestibular eye movements during self-rotation,103 and have an important role in the adaptive control of the VOR.104

The nodulus, which is the midline portion of the flocculo-nodular lobe, lying immediately caudal to the inferior medullary velum, and the adjacent ventral uvula receive afferents from the vestibular nuclei and inferior olive.105 The nodulus and uvula project to the vestibular nuclei and probably to other structures via the fastigial nucleus. Together, these structures affect the temporal response of the VOR, so that a sustained, constant-velocity rotation induces nystagmus that out-lasts by two or three times the duration of displacement of the cupula of the labyrinthine semicircular canals (a process called velocity storage).106

Lobules IV to VII of the vermis (the culmen, folium, tuber, declive, and part of the pyramis) receive mossy fiber inputs from the PPRF, NRTP, dorsolateral pontine nuclei, vestibular nuclei and nucleus prepositus hypoglossi, and climbing fiber inputs from the inferior olive.107 The projection from NRTP may relay information from the frontal eye fields necessary for the planning of saccades. The dorsal vermis projects to the fastigial nucleus, which also receives inputs from the inferior olive.108 The fastigial nucleus sends contralateral projections via the uncinate fasciculus (which runs in the dorsolateral border of the brachium conjunctivum) to the PPRF, and riMLF. Current concepts are that feedback of motor commands (efference copy) steers the eye to its target during saccades. This feedback circuit includes the dorsal vermis and fastigial nuclei, and timing of the discharge of neurons at these sites influences initial acceleration of the eye and deceleration when the eye approaches its target.109–111

EFFECTS OF CEREBELLAR LESIONS ON EYE MOVEMENTS

Experimental lesions of the flocculus and paraflocculus in monkeys produce a syndrome that is similar to that encountered clinically in patients with the Arnold-Chiari malformation.2,112 This includes impaired smooth pursuit and eye-head tracking,113 and impaired gaze-holding (deficient neural integrator). The gaze-holding deficit most probably reflects the contribution that the cerebellum makes to enhance the brainstem neural integrator, which lies in the medial vestibular nuclei and the nucleus prepositus hypoglossi. Another deficit caused by floccular lesions is loss of the ability to adapt the properties of the VOR in response to visual demands104,114; consequently, patients with disease involving the vestibular cerebellum may have vestibulo-ocular eye movements that are hypoactive or hyperactive, and they may complain of impaired vision and oscillopsia with head movements. Downbeat nystagmus (see Chapter 11) is another common finding, and probably reflects an imbalance in central vestibular connections due to loss of floccular modulation of the vertical VOR.

In monkeys, lesions of the nodulus and uvula maximize the velocity-storage effect115; maneuvers that would usually minimize the effect, such as pitching the head forward during postrotational nystagmus, are rendered ineffectual. Similar effects are seen in patients with midline cerebellar tumors that involve the nodulus.116 In addition, when monkeys that have nodular lesions are placed in darkness, they may develop periodic alternating nystagmus.115 Patients with periodic alternating nystagmus often have lesions involving the nodulus and ventral uvula.117

Experimental lesions of the dorsal vermis produce dysmetria of saccades, especially hypometria.118 Experimental inactivation of the caudal fastigial nuclei with the GABA agonist muscimol causes hypermetria of ipsilateral saccades and hypometria of contralateral saccades.119 This mechanism of saccadic dysmetria may account for the ipsipulsion of saccades encountered in patients with Wallenberg's syndrome; loss of inputs from an infarcted inferior cerebellar peduncle cause an increase in Purkinje cell activity that inhibits the ipsilateral fastigial nucleus.44 Disease processes that involve the fastigial nucleus in humans often produce a severe form of saccadic dysmetria in which the eye may repetitively overshoot a stationary target, so-called macrosaccadic oscillations (see Chapter 11); this also is reported with a genetic disorder (spinocerebellar ataxia with saccadic intrusions).120 The posterior vermis also contributes to smooth pursuit; it has reciprocal connections with the dorsolateral pontine nuclei, and neurons encode target velocity during pursuit. Lesions of the posterior vermis impair smooth pursuit.121 Unilateral lesions of the fastigial nucleus cause a defect of contralateral smooth pursuit, but bilateral lesions may leave smooth pursuit preserved, reflecting directional (acceleration-deceleration) influences of this nucleus on tracking.122

COMMON VASCULAR SYNDROMES OF THE CEREBELLUM

The cerebellum receives its blood supply from the three branches of the posterior circulation: the posterior-inferior cerebellar artery (PICA), anterior-inferior cerebellar artery (AICA), and superior cerebellar artery (SCA). Occlusion in these vessels often also produces brainstem infarction, making precise clinicopathologic correlation difficult. Infarction in the distribution of the distal PICA may cause acute vertigo and nystagmus that often simulates an acute peripheral vestibular lesion.123 These symptoms probably reflect a central vestibular imbalance, created by asymmetrical infarction in the vestibulocerebellum that normally has a tonic inhibitory effect on the vestibular nuclei. Affected patients may have gaze-evoked nystagmus that differentiates this cerebellar lesion from an acute peripheral vestibulopathy. The AICA supplies portions of the vestibular nuclei, adjacent dorsolateral brainstem and inferior lateral cerebellum (including the flocculus); in addition the AICA is the origin of the labyrinthine artery in most individuals. Consequently, ischemia in the AICA distribution may cause vertigo, vomiting, hearing loss, facial palsy, and ipsilateral limb ataxia.124 Infarction in the territory of the SCA causes ataxia of gait and limbs, and vertigo. A characteristic finding is saccadic contrapulsion. This consists of an overshooting of contralateral saccades and an undershooting of ipsilateral saccades; attempted vertical saccades are oblique, with a horizontal component away from the side of the lesion.125–127 Thus, the saccadic disorder is the opposite of the saccadic ipsipulsion seen in Wallenberg's syndrome and probably reflects interruption of crossed outputs from the fastigial nucleus running in the uncinate fasciculus adjacent to the superior cerebellar peduncle.

Back to Top
CEREBRAL AND BASAL GANGLIONIC CONTROL OF EYE MOVEMENTS
Current notions of the cerebral hemispheric control of eye movements are founded on information from several lines of investigation, each with its own strengths and weaknesses.3,128,129 Thus, caution is required, for example, in extrapolating the effects of discrete cortical lesions in monkeys to the effects of disease in humans, because the cortical architecture shows differences.130 Nonetheless, experimental studies in the rhesus monkey have provided insights into how neurons encode and program eye movements. Application of functional imaging techniques—positron emission tomography (PET) and functional magnetic resonance imaging (fMRI)—as well as transcranial magnetic stimulation, have identified similar cortical mechanisms in humans. Other advances have been made with test paradigms to identify specific defects of eye movement control in patients with well delineated cerebral lesions, such as infarcts. Thus, the scheme to be presented should be viewed as a working hypothesis that future studies are likely to clarify and refine. The approach here will be first, to lay out the scheme and salient evidence to support it and, second, to use it to account for the classic findings described with cerebral disease.

THE SACCADIC SYSTEM

Several Roles for Saccades

To appreciate the role of the cerebral hemispheres in the control saccadic movements, one must understand that rapid eye movements serve several distinct functions.3

  1. At the lowest level are the fast phases of nystagmus, which reset the eyes after vestibular or optokinetic slow-phase movements.
  2. Spontaneous saccades occur at a frequency of approximately 20 per minute and serve the purpose of visual search of the environment, although they also occur in darkness.
  3. Reflexive or nonvolitional saccades, characterized by short reaction times, occur in response to new visual, auditory, or tactile cues.
  4. Voluntary (intentional or volitional) saccades carry the eyes to a predetermined location corresponding to the position of a visual target or sound; they direct the fovea at a goal. Such voluntary saccades can also be made in a predictive fashion when the target is moving in a regular pattern: the eye movement anticipates the change in target position. Voluntary saccades can also be made toward imagined or remembered target locations, or in response to commands (e.g., “Look right”). A special case of voluntary saccades is the antisaccade, when the subject is required to respond to a visual stimulus by looking to the corresponding position in the opposite visual hemifield.
  5. Saccades can be voluntarily suppressed when it is necessary to maintain steady foveal fixation.

Thus, saccades respond to the full range of sensory inputs (visual, vestibular, auditory, somatosensory) and to internal cognitive factors. Each of these distinct functions of saccades depends on different cortical and subcortical areas; by examining each function, at the bedside or in the laboratory, it is possible to evaluate the separate neural substrates.

Several Cortical Areas that Contribute to Saccade Generation

In addition to the classic frontal and parietal eye fields (PEF), several other cortical areas contribute to the control of saccades; these include the supplementary eye fields (SEF), and dorsolateral prefrontal cortex (Fig. 4). How do sensory and cognitive signals reach these areas?

Fig. 4. A hypothetical scheme of the major structures that project to the brainstem saccade generator (premotor burst neurons in PPRF and riMLF). Also shown are projections from cortical eye fields to superior colliculus. FEF, frontal eye fields; SEF,-supplementary eye fields; DLPC, dorsolateral prefrontal cortex; IML, intramedullary lamina of thalamus; PEF, parietal eye fields (LIP); PPC, posterior parietal cortex; SNpr, substantia nigra, pars reticulata. Not shown are the pulvinar, which has connections with the superior colliculus and both the frontal and parietal lobes, and certain projections, such as that from the superior colliculus to nucleus reticularis tegmenti pontis (NRTP). B: Probable location of cortical areas important for eye movements in human brain. FEF, frontal eye field; LIP, lateral intraparietal area; MST, medial superior temporal visual area; MT, middle temporal visual area (V5); SEF, supplementary eye field; SMA, supplementary motor area; V1, primary visual cortex. (Adapted from Leigh RJ, Zee DS: The Neurology of Eye Movements, 3rd ed. Oxford University Press, New York, 1999.)

The different attributes of vision reaching striate cortex (Brodmann area 17, visual area V1), such as the fine features, color, spatial location, and motion of a viewed object, are mainly processed separately in different cortical areas.131 Thus, striate cortex projects to many separate secondary visual areas, each of which is mainly concerned with analysis of certain features of vision. Important for the planning of saccades is the inferior parietal lobule. In monkeys, area 7a is important for localizing an object in space by taking into account the location of a stimulus within the visual field and the direction in which the eyes are pointing.132 Neurons here discharge after saccades, and seem more concerned with visuospatial integration than with eye movements per se. In humans, the homologue of area 7a is probably close to the intraparietal sulcus (posterior parietal cortex [PPC]). In the posterior insula and adjacent superior temporal gyrus lies an area of multisensory vestibular cortex (MVC) that may contribute information of head or body movements.133 Together, this information is important for planning accurate saccadic eye movements and eye-head movements to objects in extrapersonal space. Also important for eye movements are the middle temporal (MT, or V5) and medial superior temporal (MST) visual areas that, in humans, probably lie at the junction of temporal, parietal, and occipital lobes posterior to the superior temporal sulcus, at the junction of Brodmann areas 19, 37, and 39, close to the intersection of the ascending limb of the inferior temporal sulcus and the lateral occipital sulcus.131,134 In monkeys, area MT contains neurons that encode the speed and direction of moving visual targets; experimental lesions here cause a selective defect of motion vision.135 Area MST, to which MT projects, contains neurons that combine visual motion information with eye movement.136 Areas MT and MST are important for generating smooth pursuit (see below), and also provide important information to cortical areas concerned with saccade generation because we often make saccades toward moving targets.

The PEF in monkeys lie in the lateral intraparietal area (LIP), within the intraparietal sulcus. Neurons here discharge before saccades.132,137 LIP projects to frontal eye fields and the superior colliculus, but not directly to the PPRF or riMLF.138 Experimental lesions of LIP delay the onset of saccades made to visual stimuli.139 In humans, the homologue of LIP, the PEF, probably lies within or close to the horizontal portion of the intraparietal sulcus.140 Parietal lesions in humans also cause an increase in the response time from presentation of a visual stimulus until the generation of a saccade (increased saccadic latency). Unilateral parietal disease may cause bilateral increases in saccadic latency, and the effect is more marked for right hemisphere lesions.141 This defect in the initiation of saccades often coexists with disorders of hemivisual attention; however, the saccadic defect may occur alone, supporting the notion that the human homologue of LIP is distinct from areas concerned with directing visual attention. The most impressive disturbance of saccades with parietal lobe lesions occurs when patients are required to respond to a double-step stimulus, in which the target jumps twice before there is time for the subjects to respond.142,143 If the target first jumps into the contralateral hemifield and then into the ipsilateral field, patients cannot make accurate saccades to the final target position, even though it lies in the intact hemifield. This suggests that the parietal lobe plays an important role in computing target position from both visual stimuli and an efference copy of eye movements (in this case, the change in eye position due to the first saccade).

The frontal eye fields (FEF) in monkeys lies along the posterior bank of the arcuate sulcus, corresponding to part of Brodmann area 8. The FEF receive input from visual areas MT and MST, dorsolateral prefrontal cortex, PEF, SEF, and from the intralaminar thalamic nuclei.144 Neurons in the monkey FEF discharge before volitional but not reflexive saccades.145 Furthermore, the FEF seem important in suppressing saccades and maintaining steady fixation. Pharmacologic inactivation of the FEF in monkeys causes a contralateral ocular motor scotoma with abolition of all reflex visual and voluntary saccades with sizes and directions corresponding to the site of injection on the FEF map.146 The homologue of FEF in humans probably lies in the precentral gyrus and sulcus, close to the intersection with the superior frontal sulcus.147 The FEF project caudally to the superior colliculus, caudate nucleus, and brainstem reticular formation (see below). Disease involving the FEF in humans causes bilaterally increased latency of saccades to visual targets if the previous (fixation) target remains visible. If, however, the fixation target disappears before the new target is presented (gap paradigm), then there is no increase in the response time.148 This evidence suggests that the FEF are important for disengagement of fixation. In addition, unilateral FEF lesions cause increased saccadic latency to remembered target locations bilaterally.149 Also impaired is the ability to generate predictive saccades, and antisaccades that are directed into the visual hemifield opposite to that of the test stimulus.129 Thus, the FEF seem important for the programming of saccades concerned with intentional exploration of the visual environment.

The SEF, in monkeys, are located in the dorsomedial frontal lobes, in the anterior part of the supplementary motor area.150 The SEF receive input from FEF, parietal cortex, and intralaminar thalamic nuclei, and project to the FEF, caudate nucleus, superior colliculus and brainstem reticular formation.151 In humans, the SEF lies in the posteromedial portion of the superior frontal gyrus, anterior to the supplementary motor area (SMA) in the upper part of the paracentral sulcus.152 One important difference between SEF and FEF is that the former but not the latter is important for generating saccades to specified spatial (rather than retinal) locations.149 Another function that is impaired by disease affecting the SEF, especially on the left side, is the ability to make a sequence of saccades to several target lights that are turned on, and remain on in a specified order.153 The SEF may be aided by anterior cingulate in learning new motor sequences.154 One patient with a discrete lesion affecting one SEF had difficulty in changing the direction of his saccadic eye movements, especially when he had to reverse the direction of a previously established pattern of response.152 The SEF and the adjacent pre-supplementary motor area (pre-SMA) seem to work together to coordinate sequential movements. 155

The dorsolateral prefrontal cortex (PFC; Brodmann area 46) makes an important contribution to voluntary control of saccades. Although not a classic eye field, it is important for making accurate motor responses (including saccades) to the remembered spatial locations of targets.156,157 Experimental lesions of PFC, including pharmacologic blockade of D1 dopamine receptors,158 interfere with this working memory area so that targets that are out of sight, are also “out of mind.” Patients with disease involving PFC show impaired accuracy of saccades to remembered target locations, increased errors on the antisaccade task, impaired predictive saccades, and impaired visual search.3,129

CORTICAL PROJECTIONS TO BRAINSTEM SACCADE CIRCUITS

Parallel descending pathways connect these cortical regions with brainstem and cerebellar structures concerned with the generation of saccades. No direct projection exists from cortical neurons to ocular motoneurons or even to saccade-generating burst neurons in the PPRF and riMLF.159–161 Instead, several intermediate structures play important roles, including the caudate nucleus, substantia nigra, superior colliculus, and brainstem reticular formation. These are summarized in Figure 4.

Parietal area LIP projects to the superior colliculus, and to the FEF, but not to the brainstem reticular formation concerned with saccade generation.138 The FEF projections run in the internal capsule; clinical lesions involving the anterior portion of the internal capsule and adjacent deep frontal region are reported to increase saccadic latency.162 Below the level of the internal capsule, several separate pathways can be discerned: one to the caudate nucleus; a second to the intralaminar thalamic nuclei and the ipsilateral superior colliculus; and a third, the pedunculopontine pathway that runs in the most medial aspect of the cerebral peduncle, and projects to the NRTP, which in turn, projects to the cerebellum.160,161,163 A partial ocular motor decussation may occur between the levels of the trochlear and abducens nuclei,164 although this has not been confirmed. The SEF also project to the caudate nucleus, where convergence with FEF projections occurs, and to the superior colliculus and pontine omnipause neurons.165 The dorsolateral prefrontal cortex projects to parts of the caudate nucleus, and to the superior colliculus.166 The caudate nucleus sends inhibitory projections to the nondopaminergic, substantia nigra pars reticulata (SNpr), which in turn, inhibits neurons in the superior colliculus. Thus, this basal ganglia system is composed of two serial, inhibitory links: a caudo-nigral inhibition that is only phasically active and a nigro-collicular inhibition that is tonically active.167 If the frontal cortex causes caudate neurons to fire, then the nigro-collicular inhibition is removed and the superior colliculus is able to activate a saccade. Disease affecting the caudate nucleus could impair the ability to make saccades to complex tasks. On the other hand, disease affecting the SNpr might disinhibit the superior colliculus, causing excessive, inappropriate saccades. Both deficits occur in patients with disorders affecting the basal ganglia, such as Huntington's disease.168

THE SUPERIOR COLLICULUS

Much of what know about the superior colliculus is based on studies in monkeys. Saccades can be produced by electrical stimulation in its ventral layers.169 In this “motor map,” the direction and size of the saccade are mainly functions of the site of stimulation, not its strength. Small saccades are elicited rostrally, large saccades caudally. Stimulation of the rostral pole of the motor map inhibits saccades; this fixation zone of the superior colliculus suppresses saccades.170 Electrophysiologic activity of more caudally placed collicular neurons indicates that they encode the overall gaze displacement,171 which is consistent with the notion that they indirectly drive burst neurons in the brainstem reticular formation. The superior colliculus has reciprocal connections with the mesencephalic reticular formation.172,173 The importance of the superior colliculus is demonstrated by the finding that inactivation of collicular burst neurons blocks the effects of frontal eye field stimulation.174 Lesions restricted to the superior colliculus in humans are uncommon, but existing reports indicate that unilateral collicular lesions impair the ability to initiate contralateral saccades at short-latency (express saccades)175 a finding consistent with animal lesion studies.176 Degenerative disorders, such as PSP (discussed below), that involve the superior colliculus, also affect the adjacent central mesencephalic reticular formation and riMLF, and initially cause slow vertical saccades; subsequent involvement of the pons and PPRF leads to saccadic slowing in all directions.177

Thus, current evidence indicates that the superior colliculus is important for initiating saccades and specifying saccade direction and size. Similar to cortical areas, superior colliculus neurons encode this information in a “place map.” The location of discharging collicular burst cells remains constant throughout the eye movement, encoding the overall gaze displacement.171

The Role of Descending Parallel Projections

The relative importance of descending pathways for saccade generation has been elucidated by studies of the effects of restricted, experimental lesions. Chronic lesions of the superior colliculus, in monkeys, cause relatively minor deficits such as an increase in saccadic latency.178 Chronic FEF lesions, in monkeys, also cause minor deficits that affect saccades to remembered targets.179 In contrast, combined lesions of the FEF and superior colliculi produce a severe and lasting deficit of eye movements, with a restricted range of movement.180 Severe deficits of saccadic and pursuit eye movements also follow combined, bilateral lesions of parietal-occipital and frontal cortex in monkeys.139 With unilateral, combined parieto-frontal lesions, saccades to visual targets in contralateral hemispace are impaired;181 with hemidecortication, the deficit is more enduring.182 In humans, combined bilateral lesions of frontal and parietal cortex cause loss of ability to make voluntary saccades—ocular motor apraxia.183

THE PURSUIT SYSTEM

As indicated above, neurons in the secondary visual areas MT encode the speed and direction of a moving target.184 MT projects to the adjacent visual area MST, where neurons encode not only visual information, but also a corollary discharge (efference copy) of the eye movement command.136 Adjacent parietal cortex also influences smooth pursuit, probably by enhancing attention on the moving target. The FEF also contribute to both the initiation and maintenance of smooth pursuit.185 From MST, projections descend to the dorsolateral pontine nuclei (DLPN),186 and subsequently to the flocculus, paraflocculus, and vermis of the cerebellum.187 The flocculus projects to the ipsilateral vestibular nuclei,101 which, in turn, project to the ocular motor nuclei; the vermis projects to the underlying fastigial nucleus.

The effects of experimental or discrete clinical stimuli can be interpreted with reference to the above scheme (Fig. 5). Lesions of striate cortex abolish smooth pursuit of targets presented in the blind hemifield, but pursuit is normal in both directions in the seeing hemifield.189 Lesions of MT do not cause a conventional visual field defect but produce a selective abnormality in target speed estimation; both saccades and smooth pursuit are consequently impaired.188 Lesions of MST in monkeys cause a disturbance much like the ipsilateral pursuit deficit encountered with unilateral cortical lesions in humans (Fig. 6); horizontal pursuit is impaired for targets moving toward the side of the lesion.135,189–192 Experimental lesions of DLPN cause a similar impairment of ipsilateral smooth pursuit.193,194 Lesions of the cerebellar flocculus, paraflocculus, and dorsal vermis also impair ipsilateral smooth pursuit.195 However, experimental or clinical lesions of the vestibular or fastigial nuclei produce a defect of contralateral smooth pursuit.196,197 Lesions of the PPRF that impair or abolish saccades do not necessarily affect smooth pursuit.36,37 Hence, the PPRF lesions that affect smooth pursuit probably do so because of involvement of axons that traverse it.

Fig. 5. A hypothetical scheme for horizontal smooth pursuit. A hypothetical anatomic scheme for smooth pursuit eye movements. Signals encoding retinal image motion pass via the lateral geniculate nucleus (LGN) to striate cortex (V1), and extrastriate areas. MT (V5), middle temporal visual area; MST, medial superior temporal visual area; PP, posterior parietal cortex; FEF, frontal eye fields; SEF, supplementary eye fields. The nucleus of the optic tract (NOT) and accessory optic system (AOS) receive visual motion signals from the retina but also from extrastriate cortical areas. Cortical areas concerned with smooth pursuit project to the cerebellum via pontine nuclei, including the dorsolateral pontine nuclei (DLPN). The cerebellar areas concerned with smooth pursuit project to ocular motor neurons via fastigial, vestibular, and y-group nuclei; the pursuit pathway for fastigial nucleus efferents has not yet been defined. The NOT projects back to LGN. The NOT and AOS may influence smooth pursuit through their projections to the pontine nuclei, and indirectly, via the inferior olive. (Adapted from Leigh RJ, Zee DS: The Neurology of Eye Movements, 3rd ed. Oxford University Press, New York, 1999.)

Fig. 6. Electro-oculography of unilateral saccadic (cogwheel) pursuit in a patient with left cerebral hemispherectomy. The upper two tracings represent position and velocity, respectively, from the right eye; the lower two tracings are from the left eye. Upward deflection represents rightward eye movement. The pursuit movement to the right (upslope) is smooth, with only occasional interspersed saccades, while pursuit toward the left (downslope) is slower than the target and requires catch-up saccades to regain target foveation. The multiple saccades are particularly evident in the velocity tracings.

A large number of conditions bilaterally impair smooth pursuit eye movements so that the eyes consequently “fall behind” the moving target and require “catch-up” saccades to regain fixation. Such saccadic or cogwheel pursuit occurs with age, sedative drugs, inattention,198 fatigue, and impaired consciousness, as well as in diffuse cerebral,199 cerebellar, or brainstem disease. A pursuit defect, difficult to distinguish from that of inattention, is found in schizophrenia.200

VESTIBULO-OCULAR SYSTEM

The VOR, during brief, natural head rotations, is probably unaffected by cortical lesions201; only when vestibular responses are supplemented by pursuit eye movements—to enhance or suppress the response—may asymmetric deficits become evident. The posterior insula and adjacent superior temporal gyrus house an area of multisensory vestibular cortex (MVC) contributes information of head or body movements133; lesions affecting this region impair patients' ability to judge the direction of earth-vertical.19

GAZE-HOLDING

Gaze-holding—sustaining the eyes conjugately in eccentric positions in the orbits—is dependent on vestibular nuclei, nucleus prepositus hypoglossi, and their cerebellar connections.2,41 However, the cerebral hemispheres, particularly the parietal lobes, influence this ability by encoding the location of a target in space. Parietal neurons respond not simply to the location of a visual stimulus on the retina, but also to the direction of gaze.202 Thus, unilateral parietal lobe lesions are manifest not only by hemispatial neglect, but also by a “gaze preference” ipsilateral to the side of the lesion.182

COMMON DISTURBANCES OF GAZE WITH UNILATERAL HEMISPHERIC LESIONS

With acute lesions, the patient's eyes are often deviated conjugately toward the side of the lesion. Such deviations are more common after large strokes involving right post-Rolandic cortex203 visual hemineglect is a common accompaniment.204 Conjugate deviation of the patient's eyes away from the side of the lesion (wrong-way deviation) occurs with subfrontal or thalamic lesions.32,205

Acutely, patients may not voluntarily direct their eyes toward the side of the intact hemisphere, but vestibular stimulation usually produces a full range of horizontal movement (with the slow phase), in contrast to gaze palsies associated with pontine lesions.70 Sometimes, in addition to ipsiversive deviation of the eyes, there is a small amplitude nystagmus with ipsilateral quick phases; the slow phases of this nystagmus may reflect unopposed pursuit drives directed away from the side of the lesion. Vertical saccades may be dysmetric with an inappropriate horizontal component toward the side of the lesion. Because both hemispheres are normally activated to elicit a purely vertical saccade, the loss of one hemisphere may be the cause of such oblique saccades.

Chronically, there usually is no resting deviation of the eyes, and persistence of gaze deviation is associated with only a prior lesion in the contralateral hemisphere.206 Attempted forced eyelid closure may cause a contralateral conjugate eye movement, the mechanism of which is not understood.207 In primary position, a small-amplitude nystagmus may be present (best seen during ophthalmoscopy), with slow phases directed toward the side of the intact hemisphere208; it may represent an imbalance in smooth pursuit tone. Horizontal pursuit is impaired for tracking of targets moving toward the side of the lesion209,210; a convenient way to demonstrate this asymmetry of pursuit is with a hand-held optokinetic drum or tape.211 The response is decreased when the stripes are moved toward the side of the lesion. At the bedside this optokinetic response is usually judged by the frequency and amplitude of quick phases. A variety of defects in the control of saccades can be brought out by special testing, as outlined above. However, when prolongation of saccadic reaction time is evident at the bedside, it may reflect defects in visual detection, visual attention, or abnormal programming of saccades.

ACUTE BILATERAL FRONTO-PARIETAL LESIONS: OCULAR MOTOR APRAXIA

Large bihemispheric lesions produce a disturbance of ocular motility called acquired ocular motor apraxia. It is characterized by loss of voluntary control of eye movements, with preservation of reflex movements, including the VOR and quick phases of nystagmus. Acquired apraxia limited to the vertical plane implies bilateral lesions at the mesencephalic-diencephalic junction.212

When acquired ocular motor apraxia is associated with acute optic ataxia and disturbance of visual attention, the eponym Balint's syndrome is used.183,213 Voluntary movements of the eyes are limited in the horizontal and vertical planes. Gaze shifts are usually achieved by combined movements of eyes and head.183 Both slow and quick phases of vestibular nystagmus are largely preserved, confirming the intact brainstem mechanisms. The defect of voluntary eye movements probably reflects disruption of descending pathways both from both the frontal and the PEF, so that the superior colliculus and brainstem reticular formation lack their supranuclear inputs; similar results have been produced experimentally in monkeys.180

Sometimes the term “spasm of fixation” is applied to such patients with difficulties in voluntarily shifting gaze. Holmes214 described a group of such patients, and noted that if the visual scene was a homogenous white screen, then voluntary eye movements became possible. As reviewed above, one effect of FEF lesions is difficulties in disengagement of fixation. Thus, it appears that spasm of fixation might have a real physiologic basis, but before applying the term to patients, it would seem necessary to demonstrate difference in saccadic latency between trials in which the fixation stimulus disappeared before the new visual stimulus appeared (gap stimulus) and trials in which the fixation stimulus remained visible.215 Brainstem disorders that affect the saccade-generating reticular formation should not be included in this category. Another factor in evaluating such patients is the nature and size of the visual stimulus. We have seen a patient with extreme difficulty making saccades on command who did so only after prolonged tight eyelid closure; there was slightly less difficulty with small refixation targets, and with larger targets the saccades were initiated with more ease until, finally, normal command saccades occurred as the patient alternately refixated on close faces.

Congenital oculomotor apraxia is characterized by absence of horizontal saccadic refixations and smooth pursuit movement, although some patients show normal smooth pursuit. During childhood, fast phases of nystagmus are absent during either optokinetic or caloric stimulation. This implicates a basic saccadic, rather than apraxic, disorder and the condition is better designated congenital saccadic palsy or congenital gaze palsy, depending on the integrity of smooth pursuit eye movements. The most striking feature of this syndrome is head thrusts used to accomplish ocular refixations. The head thrust may not be present in early infancy but develops between 4 and 8 months of age.216,217 These head movements are more exaggerated than those in patients with acquired saccadic palsy. The head moves toward the position of the eccentric new target, and the eyes, responding to the active vestibulo-ocular system, rotate conjugately in the opposite direction (Fig. 7). The eyelids often close at the onset of the head movement, which may promote saccade generation. Head rotation may overshoot the intended target, enabling the contraversively deviated eyes to fixate the target. While maintaining fixation, the head slowly moves back until the eyes are straight forward. With advancing age, the head thrust becomes less prominent and saccadic eye movements, albeit abnormal, may emerge.218 In one patient who was reexamined in adulthood, head thrusts were noted only rarely.219 This syndrome may be familial, and significant developmental central nervous system (CNS) defects such as hypoplasia of the corpus callosum and cerebrum may be associated.220–222 Apraxia of vertical gaze is usually not congenital but acquired.

Fig. 7. Congenital ocular motor apraxia. A: Patient looking to the right. B: The child's attention is drawn to the camera, the head turning to the left, with contraversion of the eyes. C: Overshoot of the head with extreme contraversion of the eyes. D: Fixation and realignment of the head and eyes. (From Felker GV, Ide CH, Hart WM: Congenital ocular motor apraxia. EENT Monthly, March 1973).

PARKINSONIAN DISORDERS AND PROGRESSIVE SUPRANUCLEAR PALSY (PSP)

A broad range of parkinsonian disorders are now recognized, and examination of eye movements, especially saccades, often contributes to accurate clinical diagnosis.

Idiopathic Parkinson's Disease

Most patients with Parkinson's disease (PD) show only minor abnormalities on clinical examination that may also occur in healthy elderly subjects. Thus, steady fixation may be disrupted by saccadic intrusions (square-wave jerks),223–225 but these are also seen in healthy elderly subjects. Moderate restriction of the range of upward gaze is common in elderly individuals,226 with or without parkinsonism, and may be caused by changes in the orbital tissues.227 Smooth pursuit may be impaired in PD as in some healthy normals.228 Convergence insufficiency occurs commonly.229 Patients with advanced PD may show slowing of vertical saccades, but PSP should always be considered in such cases. One characteristic sign in PD is hypometria that becomes more marked when patients are asked to rapidly perform self-paced refixations between two continuously visible targets (e.g., index fingers of the examiner's right and left hand positioned to test large horizontal saccades).

Laboratory studies have shown that the phenomenon of hypometria during self-paced saccades is not simply the result of persistence of the visual targets, because saccades made in anticipation of the appearance of a target light at a remembered location are also hypometric.230,231 Patients with PD also have difficulty in generating sequences of memory-guided saccades.232–234 However, saccades made reflexively to novel visual stimuli are normal in size and promptly initiated.235 Moreover, visually guided adaptation of saccades is preserved whereas memory-guided adaptation of saccades is impaired.236 Thus, patients with PD seem to have difficulty generating internally guided saccades to accurately shift gaze.235,237 Despite this hypometria, patients are still able to shift their gaze with a series of saccades to the location of a briefly flashed target; this finding indicates a retained ability to encode the location of objects in extrapersonal space.225,230

During tracking of a predictable, sinusoidal target motion, eye speed is less than target speed, leading to catch-up saccades.225,238 In addition, the catch-up saccades are hypometric; thus, the cumulative tracking eye movement is less than that of the target.239 Despite these impairments, the phase relationship between eye and target movement is normal,237 which implies a normal predictive smooth tracking strategy. This is in contrast to saccadic tracking of predictive target jumps that, as described above, is deficient.

Caloric and low-frequency rotational vestibular responses, in darkness, may be hypoactive in patients with PD.240,241 However, at higher frequencies of head rotation, and especially during visual fixation, the VOR adequately compensates for head perturbations, which explains the lack of oscillopsia complaints in patients with PD.

Treatment of PD with L-dopa does not generally improve the ocular motor deficits except that saccades tend to become larger.238,242 Newly diagnosed patients with idiopathic PD may show improved smooth pursuit after starting dopaminergic therapy.242 Memory-guided saccades are impaired after pallidotomy for Parkinson's disease243 but improve with subthalamic nucleus stimulation.244 Pallidotomy increases saccadic intrusions on steady fixation (square wave jerks).224,245 In parkinsonism caused by methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) toxicity, saccadic latency is reduced and saccadic accuracy improved by dopaminergic agents.246 Similarly, in monkeys that received MPTP, saccadic abnormalities, including increased latency, increased duration, decreased rate of spontaneous saccades, and inappropriate saccades, all are reversed by dopaminergic therapy.247,248

Oculogyric crisis, first encountered as a feature of postencephalitic parkinsonism, now usually occurs as a side effect of drugs, especially neuroleptic agents.249 A typical attack is ushered in by feelings of fear or depression, which give rise to an obsessive fixation or a thought. The eyes usually deviate upward and sometimes laterally; they rarely deviate downward. During the period of upward deviation, the movements of the eyes in the upper field of gaze appear normal. Affected patients have great difficulty in looking down, except if they combine a blink and downward saccade. Thus, the ocular disorder may reflect an imbalance of the vertical eccentric gaze-holding mechanism. Anticholinergic drugs promptly terminate the thought disorder and ocular deviation, a finding suggesting that the disorders of thought and eye movements are linked by a pharmacologic imbalance common to both.249 Oculogyric crises have also been described after striatocapsular infarction, and with bilateral putaminal hemorrhage.250 Oculogyric crises are distinct from the brief upward ocular deviations that may occur in Tourette's syndrome,251 Rett's syndrome,252 children with benign paroxysmal tonic up-gaze,65 and in patients with tardive dyskinesia.253 Episodic brief spells of tonic up-gaze have also been reported after bilateral lentiform lesions.254

PSP is a degenerative disease of later life characterized by abnormal eye movements, frequent falls, axial rigidity, difficulties with swallowing and speech, and mental slowing.255 Median survival time is approximately 6 years. The disturbance of eye movements is usually present early in the course, but occasionally develops late, or is not noted by the patient's physicians.256 Patients may present with the complaints of blurred vision, double vision, or photophobia,255 and have often been fitted with several different spectacle refractions, without improvement. These visual complaints are caused by the loss of the ability to voluntarily shift gaze downward so that patients have difficulty looking down to see a plate of food, tying their shoes, or walking down stairs with confidence.

The initial ocular motor findings are of slowing of vertical saccades and quick phases, either down, or up, or both. Sometimes vertical saccades take a curved or oblique trajectory (“round the houses”).228,257 Vertical smooth pursuit is relatively preserved but of decreased gain.258 Larger targets may elicit greater responses259 and should be used to evaluate the range of eye movements in patients in whom neck stiffness makes testing the VOR technically difficult. Full-field optokinetic stimuli may induce responses that are useful for analysis.260 Combined eye-head tracking may also be relatively spared. As the disease progresses, the range of movements with vertical saccades and pursuit declines, and eventually no voluntary vertical eye movements are possible. Usually the VOR is preserved until late in the disease (although a characteristic rigidity of the neck may make the vertical doll's head maneuver difficult to elicit).

Steady fixation is disrupted by horizontal square-wave jerks,223,228 which are more common than in other parkinsonian disorders. Horizontal saccades are hypometric but initially normal in speed.177 In some patients, the involvement of horizontal saccades may resemble internuclear ophthalmoplegia, although vestibular stimulation may overcome the limitation of adduction.261 Horizontal smooth pursuit is impaired, partly because of square-wave jerks. Convergence eye movements are commonly impaired.94 Late in the disease, there is complete ophthalmoplegia. Patients with absent quick phases but intact vestibular eye movements may also show tonic deviation of the eyes in the orbit during body rotation; if the head is free to move, it may also deviate opposite to the direction of body rotation.262

A range of eyelid abnormalities occur in PSP: blepharospasm, lid-opening apraxia, eye-closing apraxia, lid retraction, and lid lag.228 Patients usually show an inability to suppress a blink to a bright light, which is a visual Meyerson's (glabella) sign.262 Bell's phenomenon is usually absent.

Laboratory studies of saccades in PSP have demonstrated that vertical saccades are slower than horizontal saccades of similar size.177,228 The latency of horizontal saccades in PSP is prolonged in some patients, but others retain the ability to make short-latency or express saccades.263 Patients with PSP also make errors when they are required to look in the opposite direction to a suddenly appearing target (the antisaccade task). Both the presence of express saccades and errors on the antisaccade task suggest defects in frontal lobe function and, although frontal neuropathological changes are mild, positron emission scanning indicates profound frontal hypometabolism.264

Vestibular eye movements during horizontal rotation, either in darkness, or during fixation of a stationary target, confirms that PSP patients show similar slow phases to controls but abnormal quick phases, commonly leading to contraversive deviation of gaze.258

Smooth pursuit is usually impaired both horizontally and vertically.228 In the vertical plane, no corrective catch-up saccades occur. The combined impairment of vertical saccades and pursuit constitutes a voluntary gaze palsy. During large-field, vertical optokinetic stimulation, patients with PSP show tonic deviation of the eyes in the direction of stripe motion, with small or absent resetting quick phases.260 When patients with PSP shift their fixation point between distant and near targets, the vergence movement is slowed compared to normal subjects.94

Pathologically, PSP is a diffuse disorder, and although the midbrain bears the brunt of the early pathology, accounting for the relative vulnerability of vertical saccades, other parts of the brainstem, cerebellum, and forebrain are affected.265 Treatment is unsatisfactory, having little effect on the ocular motor syndrome or the general neurologic disorder

CONTRIBUTION OF EYE MOVEMENT EXAMINATION IN THE DIFFERENTIAL DIAGNOSIS OF PARKINSONISM

Careful examination of eye movements, especially saccades, can often contribute to accurate diagnosis in parkinsonism. Thus, most patients with PD have normal eye movements for their age, whereas most patients with PSP do not. A number of other parkinsonian disorders may mimic PSP. These include multiple infarcts affecting the basal ganglia, internal capsule, and midbrain (in the distribution of the perforating vessels arising from the proximal portions of the posterior cerebral artery)266; infiltrative disorders such as lymphoma; and paraneoplastic syndromes.58 Diseases causing the dorsal midbrain syndrome, such as tumor and hydrocephalus, can also present like PSP, with vertical gaze palsy.

Whipple's disease can closely mimic PSP, with vertical saccadic gaze palsy,267,268 in addition to oculomasticatory myorhythmia, a pendular vergence oscillation with concurrent contractions of the masticatory muscles; rarely limb muscles also show rhythmic contractions.269 Whipple's disease can be diagnosed using polymerase chain reaction (PCR) analysis of involved tissues,270 and be treated with antibiotics.271 Pure akinesia is characterized by profound slowness of speech, handwriting, and gait, and affected patients may suffer episodes during which they stand “frozen” for hours on end.272 Tremor, limb rigidity, akinesia, or responsiveness to levodopa are absent. Such patients may show slow and hypometric vertical saccades, and the disorder may be a restricted form of PSP with a longer, more benign course.

Cortical-basal ganglionic degeneration (CBGD) may cause a defective range of vertical eye movements, but it does not cause marked slowing of saccades. The main ocular motor defect is an increased saccadic reaction time (latency), which is evident at the bedside.228,273 The other features of this degeneration—focal dystonia, apraxia, alien hand syndrome, myoclonus, asymmetric akinetic-rigid syndrome, with late onset of balance disturbances—are more important in making the diagnosis.274,275 Multiple system atrophy (MSA) causes a parkinsonian syndrome with characteristic autonomic findings. Some patients may show slowing of vertical saccades as well as hypometria.177,228 Cerebellar eye movement findings, including downbeat nystagmus during positional testing,276 impaired smooth ocular and eye-head pursuit may occur. Diffuse Lewy body disease, characterized by parkinsonism and fluctuating dementia with florid visual hallucinations, may sometimes be associated with a vertical gaze paralysis.277,278 Slow saccades, with a supranuclear gaze palsy are reported in Creutzfeldt-Jakob disease279; cerebellar findings, including periodic alternating nystagmus, rebound nystagmus and centripetal nystagmus (slow phases directed eccentrically) on lateral gaze also may occur.279,280

HUNTINGTON'S DISEASE

Huntington's disease (HD) is caused by a genetic defect of the IT15 gene (huntingtin) on chromosome 4, leading to increased CAG triplet repeat length. Disturbances of voluntary gaze are common.168,281–283 Saccades are initiated with difficulty and with prolonged latencies, especially when made to command. An obligatory blink or head thrust may be used to start the eye moving.284 Saccades may be slow horizontally or vertically. Longitudinal studies have documented progressive slowing of saccades with prolongation of reaction time.285 Saccades may be slower in patients who become symptomatic at an earlier age, and who have inherited the disease from their father.168

Fixation is disrupted in some patients by saccadic intrusions.283 Smooth pursuit may be impaired, but often is relatively spared compared with saccades. Eccentric gaze holding and the VOR are usually normal. Late in the disease, rotational stimulation causes the eyes to deviate tonically because of impaired quick phases.

Individuals studied at a presymptomatic point in their disease may have normal eye movements.281,286,287 Thus, testing of eye movements cannot be regarded as a reliable method for determining which offspring of affected patients will go on to develop the disease. Improvement of saccadic eye movement abnormalities in HD has been reported with sulpiride.288

FUNCTIONAL DISTURBANCES OF GAZE

Spasm of the near reflex (convergence spasm), which may mimic a lateral rectus paresis, has been discussed above. Rarely, patients may present with a psychogenic paralysis of horizontal gaze. Pursuit and saccades are usually symmetrically restricted. Sophisticated neuro-ophthalmologic testing using passive head rotations, optokinetic stimuli, and pursuit stimulation with the large mirror test should demonstrate the full ability to make fast and slow eye movements and uncover the psychogenic nature of the disturbance.289 In the mirror test, the patient is asked to look into a large hand-held mirror that is then tilted right and left or up and down. The patient's eyes readily moved with the moving reflection of the visual environment.

TABLE 4. Ocular Motor Disorders Caused By Lesions Of The Cerebellum

Vestibulocerebellum

Flocculus, Paraflocculus, and Ventral Uvula:
  1. Gaze-evoked nystagmus (leaky neural integrator)
  2. Impaired smooth pursuit (and eye-head tracking)
  3. Downbeat nystagmus
  4. Rebound nystagmus
  5. Uncalibrated vestibulo-ocular reflex
Nodulus:
  1. Prolongation of durations of vestibular nystagmus in response to rotational stimuli
  2. Periodic alternating nystagmus (when in darkness).

Dorsal vermis, fastigial nucleus, and their projections

Dorsal vermis lesions: Saccadic dysmetria, with hypometria of contralateral saccades
Fastigial Nucleus: Hypermetria of ipsilateral saccades (ipsipulsion), hypometria of contralateral saccades and impairment of contralateral smooth pursuit
Uncinate fasciculus and superior cerebellar peduncle: contralateral hypermetria (contrapulsion)

 

Back to Top
REFERENCES

1. Kaminski HJ, Leigh RJ, eds: Neurobiology of Eye Movements—From Molecules to Behavior. New York: New York Academy of Sciences, 2002

2. Leigh R, Zee D: The Neurology of Eye Movements. Third ed. New York: Oxford University Press, 1999

3. Leigh RJ, Kennard C: Using saccades as a research tool in clinical neurosciences. Brain 27:1–18, 2004

4. Demer JL: Pivotal role of orbital connective tissues in binocular alignment and strabismus. Invest Ophthalmol Vis Sci 45:729–738, 2004

5. Büttner-Ennever JA, Horn AK, Graf W, et al:et al: Modern concepts of brainstem anatomy: from extraocular motoneurons to proprioceptive pathways. Ann NY Acad Sci 956:75–84, 2002

6. Highstein SM, Baker R: Excitatory termination of abducens internuclear neurons on medial rectus motoneurons: relationship to syndrome of internuclear ophthalmoplegia. J Neurophysiol 41:1647–1661, 1978

7. King W, Lisberger S, Fuchs A: Responses of fibers in medial longitudinal fasciculus (MLF) of alert monkeys during horizontal and vertical conjugate eye movements evoked by vestibular or visual stimuli. J Neurophysiol 39:1135–1149, 1976

8. McCrea R, Strassman A, Highstein S: Morphology and physiology of abducens motoneurons and internuclear neurons intracellularly injected with horseradish peroxidase in alert squirrel monkeys. J Comp Neurol 243:291–308, 1986

9. Spencer R, Baker R. Histochemical localization of acetycholinesterase in relation to motor neurons and internuclear neurons of the cat abducens nucleus. J Neurocytol 15:137–154, 1986

10. McCrea R, Strassman A, May E, et al:et al: Anatomical and physiological characteristics of vestibular neurons mediating the horizontal vestibulo-ocular reflex of the squirrel monkey. J Comp Neurol 264:547–570, 1987

11. Horn AKE, Büttner-Ennever JA, Suzuki Y, et al: Histological identification of premotor neurons for horizontal saccades in monkey and man by parvalbumin immunostaining. J Comp Neurol 359:350–363, 1997

12. Belknap D, McCrea R: Anatomical connections of the prepositus and abducens nuclei in the squirrel monkey. J Comp Neurol 268:13–28, 1988

13. Bronstein AM, Morris J, Du B, et al: Abnormalities of horizontal gaze. Clinical, oculographic and magnetic resonance imaging findings. I Abducens palsy. J Neurol Neurosurg Psychiatry 53:194–199, 1990

14. Hirose G, Furui K, Yoshioka A, et al: Unilateral conjugate gaze palsy due to a lesion of the abducens nucleus. J Clin Neuroophthalmol 13:54–58, 1993

15. Meienberg O, Büttner-Ennever JA, Kraus-Ruppert R: Unilateral paralysis of conjugate gaze due to lesion of the abducens nucleus. Clinico-pathological case report. Neuroophthalmology 2:47–52, 1981

16. Müri RM, Chermann JF, Cohen L, et al: Ocular motor consequences of damage to the abducens nucleus area in humans. J Neuroophthalmol 16:191–195, 1996

17. Gamlin PD, Gnadt JW, Mays LE: Lidocaine-induced unilateral internuclear ophthalmoplegia: effects on convergence and conjugate eye movements. J Neurophysiol 62:82–95, 1989

18. Zee DS, Hain TC, Carl JR: Abduction nystagmus in internuclear ophthalmoplegia. Ann Neurol 21:383–388, 1987

19. Brandt T, Dieterich M: Vestibular syndromes in the roll plane: topographic diagnosis from brainstem to cortex. Ann Neurol 36:337–347, 1994

20. Cremer P, Migliaccio A, Halmagyi GM, et al: Vestibulo-ocular reflex pathways in internuclear ophthalmoplegia. Ann Neurol 45:529–533, 1999

21. Ranalli P, Sharpe J: Vertical vestibulo-ocular reflex, smooth pursuit and eye-head tracking dysfunction in internuclear ophthalmoplegia. Brain 111:1299–1317, 1988

22. Herishanu YO, Sharpe JA. Saccadic intrusions in internuclear ophthalmoplegia. Ann Neurol 14:67–72, 1983

23. Smith JL, David NJ: Internuclear ophthalmoplegia. Two new clinical signs. Neurology 14:307–309, 1964

24. Frohman TC, Frohman EM, O'Suilleabhain P, et al: Accuracy of clinical detection of INO in MS: corroboration with quantitative infrared oculography. Neurology. 61:848–850, 2003

25. Frohman EM, Frohman TC, O'Suilleabhain P, et al: Quantitative oculographic characterisation of internuclear ophthalmoparesis in multiple sclerosis: the versional dysconjugacy index Z score. J Neurol Neurosurg Psychiatry. 73:51–55, 2002

26. Flipse JP, Straathof CS, Van der Steen J, et al: Binocular saccadic acceleration in multiple sclerosis. Neuro-ophthalmology 16:43-46, 1996

27. Ventre J, Vighetto A, Bailly G, et al: Saccade metrics in multiple sclerosis: versional velocity disconjugacy as the best clue? J Neurol Sci 102:144–149, 1991

28. Frohman EM, Zhang H, Kramer PD, et al: MRI characteristics of the MLF in MS patients with chronic internuclear ophthalmoparesis. Neurology. 57:762–768, 2001

29. Lutz A. Ueber die Bahnen der Blickwendung und deren Dissoziierung. Klin Monatsble Augenheilkd 70:213–235, 1923

30. Bogousslavsky J, Regli F, Ostinelli B, et al: Paresis of lateral gaze alternating with so-called posterior internuclear ophthalmoplegia. A partial paramedian pontine reticular formation-abducens nucleus syndrome. J Neurol 232:38–42, 1985

31. Clendaniel RA, Mays LE: Characteristics of antidromically identified oculomotor internuclear neurons during vergence and versional eye movements. J Neurophysiol 71:1111–1127, 1994

32. Fisher CM: Some neuro-ophthalmological observations. J Neurol Neurosurg Psychiatry 30:383–392, 1967

33. Pierrot-Deseilligny 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. Brain 104:665–699, 1981

34. Wall M, Wray SH: The one-and-a-half syndrome—A unilateral disorder of the pontine tegmentum: a study of 20 cases and review of the literature. Neurology 33:971-980, 1983

35. Sharpe JA, Rosenberg MA, Hoyt WF, et al: Paralytic pontine exotropia. A sign of acute unilateral pontine gaze palsy and internuclear ophthalmoplegia. Neurology 24:1076-1081, 1974

36. Hanson M, Hamid M, Tomsak R, et al: Selective saccadic palsy caused by pontine lesions: clinical, physiological, and pathological correlations. Ann Neurol 20:209-217, 1986

37. Henn V, Lang W, Hepp K, et al: Experimental gaze palsies in monkeys and their relation to human pathology. Brain 107:619-636, 1984

38. Johnston JL, Sharpe JA, Ranalli PJ, et al: Oblique misdirection and slowing of vertical saccades after unilateral lesions of the pontine tegmentum. Neurology 43:2238-2244, 1993

39. Zee DS, Optican LM, Cook JD, et al: Slow saccades in spinocerebellar degeneration. Arch Neurol 33:243-251, 1976

40. Aksay E, Gamkrelidze G, Seung HS, et al: In vivo intracellular recording and perturbation of persistent activity in a neural integrator. Nat Neurosci 4:184–193, 2001

41. Cannon SC, Robinson DA: Loss of the neural integrator of the oculomotor system from brainstem lesions in monkey. J Neurophysiol 57:1383–1409, 1987

42. Dieterich M, Brandt T: Wallenberg's syndrome: lateropulsion, cyclorotation, and subjective visual vertical in thirty-six patients. Ann Neurol 31:399–408, 1992

43. Kirkham TH, Guitton D, Gans M: Task-dependent variations of ocular lateropulsion in Wallenberg's syndrome. Can J Neurol Sci 8:21–22, 1981

44. Waespe W, Baumgartner R: Enduring dysmetria and impaired gain adaptivity of saccadic eye movements in Wallenberg's lateral medullary syndrome. Brain 115:1125–1146, 1992

45. Büttner-Ennever JA, Buttner U: A cell group associated with vertical eye movements in the rostral mesencephalic reticular formation of the monkey. Brain Res 151:31–47, 1978

46. Horn AKE, Büttner-Ennever JA: Premotor neurons for vertical eye-movements in the rostral mesencephalon of monkey and man: the histological idnetification by parvalbumin immunostaining. J Comp Neurol 392:413–427, 1998

47. Suzuki Y, Büttner-Ennever JA, Straumann D, et al: Deficits in torsional and vertical rapid eye movements and shift of Listing's plane after uni- and bilateral lesions of the rostral interstitial nucleus of the medial longitudinal fasciculus. Exp Brain Res 106:215–232, 1995

48. Moschovakis AK, Scudder CA, Highstein SM: Structure of the primate oculomotor burst generator. I. Medium-lead burst neurons with upward on-directions. J Neurophysiol 65:203–217, 1991

49. Moschovakis AK, Scudder CA, Highstein SM, et al: Structure of the primate oculomotor burst generator. II. Medium-lead burst neurons with downward on-directions. J Neurophysiol 65:218–229, 1991

50. Kokkoroyannis T, Scudder CA, Balaban CD, et al: Anatomy and physiology of the primate interstitial nucleus of Cajal I. Efferent projections. J Neurophysiol 75:725–739, 1996

51. Partsalis AM, Highstein SM, Moschovakis AK: Lesions of the posterior commissure disable the vertical neural integrator of the primate oculomotor system. J Neurophysiol 71:2582–, 19942585

52. Pasik P, Pasik T, Bender MB: The pretectal syndrome in monkeys. I. Disturbances of gaze and body posture. Brain 92:521–534, 1969

53. Fukushima F. The interstitial nucleus of Cajal and its role in the control of movements of head and eyes. Prog Neurobiol 29:107–192, 1987

54. McCrea R, Strassman A, Highstein S: Anatomical and physiological characteristics of vestibular neurons mediating the vertical vestibulo-ocular reflexes of the squirrel monkey. J Comp Neurol 264:571–594, 1987

55. Ranalli PJ, Sharpe JA, Fletcher WA: Palsy of upward and downward saccadic, pursuit, and vestibular movements with a unilateral midbrain lesion: pathophysiologic correlations. Neurology 38:114–122, 1988

56. Bender MB: Brain control of conjugate horizontal and vertical eye movements. A survey of the structural and functional correlates. Brain 103:23–69, 1980

57. Kompf D, Pasik T, Pasik P, et al: Downward gaze in monkeys. Stimulation and lesion studies. Brain 102:527–558, 1979

58. Bhidayasiri R, Plant GT, Leigh RJ: A hypothetical scheme for the brainstem control of vertical gaze. Neurology 54:1985–1993, 2000

59. Büttner-Ennever JA, Buttner U, Cohen B, et al: Vertical gaze paralysis and the rostral interstitial nucleus of the medial longitudinal fasciculus. Brain 105:125–149, 1982

60. Crawford JD, Cadera W, Vilis T: Generation of torsional and vertical eye position signals by the interstitial nucleus of Cajal. Science 252:1551–1553, 1991

61. Westheimer G, Blair SM: The ocular tilt reaction—A brainstem oculomotor routine. Invest Ophthalmol 14:833–839, 1975

62. Lueck CJ, Hamlyn P, Crawford TJ, et al: A case of ocular tilt reaction and torsional nystagmus due to direct stimulation of the midbrain in man. Brain 114:2069–2079, 1991

63. Evinger LC, Fuchs AF, Baker R: Bilateral lesions of the medial longitudinal fasciculus in monkeys: Effects on the horizontal and vertical components of voluntary and vestibular induced eye movements. Exp Brain Res 28:1–20, 1977

64. Keane JR: Sustained upgaze in coma. Ann Neurol 9:409–412, 1981

65. Hoyt CS, Mousel DK: Transient supranuclear disturbances of gaze in healthy neonates. Am J Ophthalmol 89:708–713, 1980

66. Fisher CM: The pathologic and clinical aspects of thalamic hemorrhage. Trans Am Neurol Assoc 84:56–59, 1959

67. Frohman LP, Kupersmith MJ: Reversible vertical ocular deviations associated with raised intracranial pressure. J Clin Neuroopthalmol 5:158–163, 1985

68. Keane JR, Rawlinson DG, Lu AT: Sustained downgaze deviation. Two cases without structural pretectal lesions. Neurology 26:594–595, 1976

69. Corbett JJ, Schatz NJ, Shults WT, et al: Slowly alternating skew deviation: description of a pretectal syndrome in three patients. Ann Neurol 10:540–546, 1981

70. Daroff RB, Hoyt WF: Supranuclear disorders of ocular control systems in man: clinical, anatomical and physiological correlations—1969. In: Bach-y-Rita P, Collins CC, Hyde JE, eds. The Control of Eye Movements. New York: Academic Press, 1971:175–235

71. Büttner-Ennever JA, Acheson JF, Büttner U, et al: Ptosis and supranuclear downgaze palsy. Neurology 39:385–389, 1989

72. Sand JJ, Biller J, Corbett JJ, et al: Partial dorsal mesencephalic hemorrhages: report of three cases. Neurology 36:529–533, 1986

73. Keane JR: Bilateral ocular motor signs after tentorial herniation in 25 patients. Arch Neurol 43:806–807, 1986

74. Seaber JH, Nashold BS: Comparison of ocular motor effects of unilateral stereotactic midbrain lesions in man. Neuroophthalmology 1:95–99, 1980

75. Rambold H, Kompf D, Helmchen C: Convergence retraction nystagmus: a disorder of vergence? Ann Neurol 50:677–681, 2001

76. Ochs AL, Stark L, Hoyt WF, et al: Opposed adducting saccades in convergence-retraction nystagmus. A patient with sylvian aqueduct syndrome. Brain 102:479–508, 1979

77. Zee DS, FitzGibbon EJ, Optican LM: Saccade-vergence interactions in humans. J Neurophysiol 68:1624–1641, 1992

78. Brenner RP, Carlow TJ: PLEDs and nystagmus retractorius. Ann Neurol 5:403, 1979

79. Keane JR: Ocular skew deviation. Arch Neurol 32:185–190, 1975

80. Moster ML, Schatz NJ, Savino PJ, et al: Alternating skew on lateral gaze (bilateral abducting hypertropia). Ann Neurol 23:190–192, 1988

81. Zee DS: Considerations on the mechanisms of alternating skew deviation in patients with cerebellar lesions. J Vestibular Res 6:1–7, 1996

82. Brandt T, Dieterich M: Pathological eye-head coordination in roll: tonic ocular tilt reaction in mesencephalic and medullary lesions. Brain 110:649–666, 1987

83. Halmagyi GM, Brandt T, Dieterich M, et al: Tonic contraversive ocular tilt reaction due to unilateral meso-diencephalic lesion. Neurology 40:1503–1509, 1990

84. Rabinovitch HE, Sharpe JA, Sylvester TO: The ocular tilt reaction. A paroxysmal dyskinesia associated with elliptical nystagmus. Arch Ophthalmol 95:1395–1398, 1977

85. Hedges TR, Hoyt WF: Ocular tilt reaction due to an upper brainstem lesion: paroxysmal skew deviation, torsion, and oscillation of the eyes with head tilt. Ann Neurol 11:537–540, 1982

86. Lewis JM, Kline LB: Periodic alternating nystagmus associated with periodic alternating skew deviation. J Clin Neuroopthalmol 13:115–117, 1983

87. Gamlin PDR, Mays LE: Dynamic properties of medial rectus motoneurons during vergence eye movements. J Neurophysiol 67:64–74, 1992

88. Mays LE, Gamlin PDR: Neuronal circuitry controlling the near response. Curr Opin Neurobiol 5:763–768, 1995

89. Mays LE: Neural control of vergence eye movements: convergence and divergence neurons in the midbrain. J Neurophysiol 51:1091–1108, 1984

90. Zhang Y, Mays LE, Gamlin PDR: Characteristics of near response cells projecting to the oculomotor nucleus. J Neurophysiol 67:944–960, 1992

91. Judge SJ, Cumming BG: Neurons in the monkey midbrain with activity related to vergence eye movements and accommodation. J Neurophysiol 55:915–930, 1986

92. Mays LE, Porter JD, Gamlin PDR, et al: Neural control of vergence eye movements: neurons encoding vergence velocity. J Neurophysiol 56:1007–1021, 1986

93. Gamlin PDR, Clarke RJ: Single-unit activity in the primate nucleus reticularis tegmenti pontis related to vergence and accommodation. J Neurophysiol 73:2115–2119, 1995

94. Kitthaweesin K, Riley DE, Leigh RJ: Vergence disorders in progressive supranuclear palsy. Ann NY Acad Sci 956:504–507, 2002

95. Rambold H, Neumann G, Helmchen C: Vergence deficits in pontine lesions. Neurology 62:1850–1853, 2004

96. Keane JR: Neuro-ophthalmological signs and symptoms of hysteria. Neurology 32, 1982

97. Guiloff RJ, Whiteley A, Kelly RE: Organic convergence spasm. Acta Neurol Scand 61:252–259, 1980

98. Krohel GB, Kristan RW, Simon JW, et al: Divergence paralysis. Am J Ophthalmol 94:506–510, 1982

99. Roper-Hall G, Burde RM: Diagnosis and management of divergence paralysis. Am Orthop J 37:113, 1987

100. Langer T, Fuchs A, Scudder C, et al: Afferents to the flocculus of the cerebellum in the rhesus macaque as revealed by retrograde transport of horseradish peroxidase. J Comp Neurol 235:1–25, 1985

101. Langer T, Fuchs A, Chubb M, et al: Floccular efferents in the rhesus macaque as revealed by autoradiography and horseradish peroxidase. J Comp Neurol 235:26–37, 1985

102. Belton T, McCrea RA: Role of the cerebellar flocculus region in the coordination of eye and head movements during gaze pursuit. J Neurophysiol 84:1614–1626, 2000

103. Waespe W, Henn V: Visual-vestibular interaction in the flocculus of the alert monkey. II. Purkinje cell activity. Exp Brain Res 43:349–360, 1981

104. Raymond JL, Lisberger SG, Mauk MD: The cerebellum: a neuronal learning machine? Science 272:1126–1131, 1996

105. Walberg F, Dietrichs E: The interconnection between the vestibular nuclei and the nodulus: a study of reciprocity. Brain Res 449:47–53, 1988

106. Cohen H, Cohen B, Raphan T, et al: Habituation and adaptation of the vestibulo-ocular reflex: a model of differential control by the vestibulo-cerebellum. Exp Brain Res 110:110–120, 1993

107. Yamada J, Noda H: Afferent and efferent connections of the oculomotor cerebellar vermis in the macaque monkey. Journal of Comparative Neurology 265:224–241, 1987

108. Ikeda Y, Noda H, Sugita S: Olivocerebellar and cerebelloolivary connections of the oculomotor region of the fastigial nucleus in the macaque monkey. J Comp Neurol 284:463–488, 1989

109. Thier P, Dicke PW, Haas R, et al: The role of the oculomotor vermis in the control of saccadic eye movements. Ann NY Acad Sci 978:50–62, 2002

110. Fuchs AF, Robinson FR, Straube A: Role of the caudal fastigial nucleus in saccade generation: I. Neuronal discharge patterns. J Neurophysiol 70:1723–1740, 1993

111. Buttner U, Glasauer S, Glonti L, et al: Multimodal signal integration in vestibular neurons of the primate fastigial nucleus. Ann NY Acad Sci 1004:241–251, 2003

112. Zee D, Yamazaki A, Butler P, et al: Effects of ablation of flocculus and paraflocculus on eye movements in primate. J Neurophysiol 46:878–899, 1981

113. Rambold H, Churchland A, Selig Y, et al: Partial ablations of the flocculus and ventral paraflocculus in monkeys cause linked deficits in smooth pursuit eye movements and adaptive modification of the VOR. J Neurophysiol 87:912–924, 2002

114. Lisberger S, Miles F, Zee D: Signals used to compute errors in monkey vestibuloocular reflex: possible role of flocculus. J Neurophysiol 52:1140–1153, 1984

115. Waespe W, Cohen B, Raphan T: Dynamic modification of the vestibulo-ocular reflex by the nodulus and uvula. Science 228:199–202, 1985

116. Hain T, Zee D, Maria B: Tilt suppression of vestibulo-ocular reflex in patients with cerebellar lesions. Acta Otolaryngol (Stockh) 105:13–20, 1988

117. Furman JM, Wall C III, Pang D: Vestibular function in periodic alternating nystagmus. Brain 113:1425–1439, 1990

118. Takagi M, Zee DS, Tamargo R: Effects of lesions of the oculomotor vermis on eye movements in primate: saccades. J Neurophysiol 80:1911–1930, 1998

119. Robinson FR, Straube A, Fuchs AF. Role of the caudal fastigial nucleus in saccade generation. II. Effects of muscimol inactivation. J Neurophysiol 70:1741–1758, 1993

120. Swartz BE, Bespavola IN, Burmeister M, et al: Pathogenesis of clinical signs in recessive cerebellar ataxia with saccadic intrusions and sensorimotor neuropathy (SCASI). Ann Neurol 54:824–828, 2003

121. Pierrot-Deseilligny C, Amarenco P, Roullet E, et al: Vermal infarct with pursuit eye movement disorders. J Neurol Neurosurg Psychiatry 53:519–521, 1990

122. Buttner U, Straube A, Spuler A: Saccadic dysmetria and “intact” smooth pursuit eye movements after bilateral deep cerebellar nuclei lesions. J Neurol Neurosurg Psychiatry 57:832–834, 1994

123. Duncan GW, Parker SW, Fisher CM. Acute cerebellar infarction in the PICA territory. Arch Neurol 32:364–368, 1975

124. Grad A, Baloh RW: Vertigo of vascular origin. Clinical and electronystagmographic features in 84 cases. Arch Neurol 46:281–284, 1989

125. Benjamin EE, Zimmerman CF, Troost BT: Lateropulsion and upbeat nystagmus are manifestations of central vestibular dysfunction. Arch Neurol 43:962–964, 1986

126. Ranalli P, Sharpe J: Contrapulsion of saccades and ipsilateral ataxia: a unilateral disorder of the rostral cerebellum. Ann Neurol 20:311–316, 1986

127. Uno A, Mukuno K, Sekiya H, et al: Lateropulsion in Wallenberg's syndrome and contrapulsion in the proximal type of the superior cerebellar artery syndrome. Neuroophthalmology 9:75–80, 1998

128. Pierrot-Deseilligny C, Muri RM, Ploner CJ, et al: Cortical control of ocular saccades in humans: a model for motricity. Prog Brain Res 142:3–17, 2003

129. Pierrot-Deseilligny C, Ploner CJ, Muri RM, et al: Effects of cortical lesions on saccadic: eye movements in humans. Ann N Y Acad Sci. 956:216–229, 2002

130. Andersen R: Visual and eye movement functions of the posterior parietal cortex. Ann Rev Neurosci 12:377–403, 1989

131. Zeki S: A Vision of the Brain. London: Blackwell Scientific, 1993

132. Barash S, Bracewell RM, Fogassi L, et al: Saccade-related activity in the lateral intraparietal area. I. Temporal proper comparison with area 7a. J Neurophysiol , ties 66:1095–1108, 1991

133. Dieterich M, Bense S, Lutz S, et al: Dominance for vestibular cortical function in the non-dominant hemisphere Cereb Cortex 13:1007, 2003

134. Ocraven KM, Rosen BR, Kwong KK, et al: Voluntary attention modulates fMRI activity in human MT-MST. Neuron 18:591–598, 1997

135. Dursteler M, Wurtz R: Pursuit and optokinetic deficits following chemical lesions of cortical areas MT and MST. J Neurophysiol 60:940–965, 1988

136. Newsome W, Wurtz R, Komatsu H: Relation of cortical areas MT and MST to pursuit eye movements. II. Differentiation of retinal from extraretinal inputs. J Neurophysiol 60:604–620, 1988

137. Goldberg ME, Bisley J, Powell KD, et al: The role of the lateral intraparietal area of the monkey in the generation of saccades and visuospatial attention. Ann NY Acad Sci 956:205–215, 2002

138. Lynch J, Graybiel A, Lobeck L: The differential projection of two cytoarchitectonic subregions of the inferior parietal lobule of macaque upon the deep layers of the superior colliculus. J Comp Neurol 235:241–254, 1985

139. Lynch JC: Saccade initiation and latency deficits after combined lesions of the frontal and posterior eye fields in monkeys. J Neurophysiol 68:1913–1916, 1992

140. Müri RM, Iba-Zizen MT, Derosier C, et al: Location of the human posterior eye field with functional magnetic resonance imaging. J Neurol Neurosurg Psychiatry 60:445–448, 1996

141. Pierrot-Deseilligny C, Rivaud S, Gaymard B, et al: Cortical control of reflexive visually-guided saccades. Brain 114:1473-1485, 1991

142. Duhamel J, Goldberg ME, FitzGibbon EJ, et al: Saccadic dysmetria in a patient with a right frontoparietal lesion: the importance of corollary discharge for accurate spatial-behavior. Brain 115:1387-1402, 1992

143. Heide W, Blankenburg M, Zimmermann E, et al: Cortical control of double-step saccades: Implications for spatial orientation. Ann Neurology 38:739-748, 1995

144. Huerta M, Krubitzer L, Kaas J. Frontal eye field as defined by intracortical microstimulation in squirrel monkeys, owl monkeys, and macaque monkeys. II. Cortical connections. J Comp Neurol 265:332-361, 1987

145. Bruce CJ, Goldberg ME. Primate frontal eye fields. I. Single neurons discharging before saccades. J Neurophysiol 53:603-635, 1985

146. Dias EC, Kiesau M, Segraves MA: Acute activation and inactivation of macaque frontal eye field with GABA-related drugs. J Neurophysiol 74:2744–2748, 1995

147. Rosano C, Krisky CM, Welling JS, et al: Pursuit and saccadic eye movement subregions in human frontal eye field: a high-resolution fMRI investigation. Cereb Cortex 12:107–115, 2002

148. Rivaud S, Müri RM, Gaymard B, et al: Eye movement disorders after frontal eye field lesions in humans. Exp Brain Res 102:110–120, 1994

149. Pierrot-Deseilligny C, Israël I, Berthoz A, et al: Role of the different frontal lobe areas in the control of the horizontal component of memory-guided saccades in man. Exp Brain Res 95:166–171, 1993

150. Schlag J, Schlag-Rey M: Evidence for a supplementary eye field. J Neurophysiol 57:179–200, 1987

151. Huerta M, Kaas J: Supplementary eye fields as defined by intracortical microstimulation: connections in macaques. J Comp Neurol 293:299–330, 1990

152. Husain M, Parton A, Hodgson TL, et al: Self-control during response conflict by human supplementary eye field. Nat Neurosci 6:117-118, 2003

153. Gaymard B, Rivaud S, Pierrot-Deseilligny C: Role of the left and right supplementary motor areas in memory-guided saccade sequences. Ann Neurol 34:404–406, 1993

154. Procyk E, Tanaka YL, Joseph JP: Anterior cingulate activity during routine and non-routine sequential behaviors in macaques. Nat Neurosci 3:502–508, 2000

155. Tanji J: Sequential organization of multiple movements: Involvement of cortical motor areas. Annu Rev Neurosci 24:631–651 , 2001

156. Goldmanrakic PS: Cellular basis of working memory. Neuron 14:477–485, 1995

157. Sawaguchi T, Iba M: Prefrontal cortical representation of visuospatial working memory in monkeys examined by local inactivation with muscimol. J Neurophysiol 86:2041–2053, 2001

158. Sawaguchi T: The effects of dopamine and its antagonists on directional delay-period activity of prefrontal neurons in monkeys during an oculomotor delayed-response task. Neurosci Res 41:115–128, 2001

159. Iwatsubo T, Kuzuhara S, Kanemitsu A, et al: Corticofugal projections to the motor nuclei of the brainstem and spinal cord in humans. Neurology 40:309–312, 1990

160. Stanton GB, Goldberg ME, Bruce CJ: Frontal eye field efferents in the macaque monkey. I. Subcortical pathways and topography of striatal and thalamic terminal fields. J Comp Neurol 271:473–492, 1988

161. Stanton GB, Goldberg ME, Bruce CJ: Frontal eye field efferents in the macaque monkey. II. Topography of terminal fields in midbrain and pons. J Comp Neurol 271:493–506, 1988

162. Pierrot-Deseilligny C, Rivaud S, Penet C, et al: Latencies of visually guided saccades in unilateral hemispheric cerebral lesions. Ann Neurol 21:138–148, 1987

163. Gonzalo-Ruiz A, Leichnetz G, Smith D: Origin of cerebellar projections to the region of the oculomotor complex, medial pontine reticular formation, and superior colliculus in new world monkeys: a retrograde horseradish peroxidase study. J Comp Neurol 268:508–526, 1988

164. Leichnetz G: The prefrontal cortico-oculomotor trajectories in the monkey. J Neurol Sci 49:387–396, 1981

165. Shook B, Schlag-Rey M, Schlag J: Direct projection from the supplementary eye field to the nucleus raphe interpositus. Exp Brain Res 73:215–218, 1988

166. Selemon L, Goldman-Rakic P: Common cortical and subcortical targets of the dorsolateral prefrontal and posterior parietal cortices in the rhesus monkey: evidence for a distributed neural network subserving spatially guided behavior. J Neurosci 8:4049–4068, 1988

167. Hikosaka O, Takikawa Y, Kawagoe R: Role of the basal ganglia in the control of purposive saccadic eye movements. Physiol Rev 80:953–978, 2000

168. Lasker AG, Zee DS, Hain TC, et al: Saccades in Huntington's disease: initiation defects and distractability. Neurology 37:364–370, 1987

169. Robinson DA. Eye movements evoked by collicular stimulation in the alert monkey. Vision Res 12:1795–1808, 1972

170. Everling S, Paré M, Dorris MC, et al: Comparison of the discharge charactersitics of brainstem omipause neurons and superior colliculus fixation neurons in monkey: implications for control of fixation and saccade behavior. J Neurophysiol 79:511–528, 1998

171. Bergeron A, Matsuo S, Guitton D. Superior colliculus encodes distance to target, not saccade amplitude, in multi-step gaze shifts. Nat Neurosci 6:404–413, 2003

172. Waitzman DM, Silakov VL, DePalma-Bowles S, et al: Effects of reversible inactivation of the primate mesencephalic reticular formation. I. Hypermetric goal-directed saccades. J Neurophysiol 83:2260–2284, 2000

173. Waitzman DM, Silakov VL, DePalma-Bowles S, et al: Effects of reversible inactivation of the primate mesencephalic reticular formation. II. Hypometric vertical saccades. J Neurophysiol 83:2285–2299, 2000

174. Hanes DP, Wurtz RH: Interaction of the frontal eye field and superior colliculus for saccade generation. J Neurophysiol 85:804–815, 2001

175. Pierrot-Deseilligny C, Rosa A, Masmoudi K, et al: Saccade deficits after a unilateral lesion affecting the superior colliculus. J Neurol Neurosurg Psychiatry 54:1106–1109, 1991

176. Schiller PH, Sandell JH, Maunsell JHR: The effect of frontal eye field and superior colliculus lesions on saccadic latencies in the rhesus monkey. J Neurophysiol 57:1033–, 19871049

177. Bhidayasiri R, Riley DE, Somers JT, et al: Pathophysiology of slow vertical saccades in progressive supranuclear palsy. Neurology 57:2070–2077, 2001

178. Schiller P, Sandell J, Maunsell J: The effect of frontal eye field and superior colliculus lesions on saccadic latencies in the rhesus monkey. J Neurophysiol 57:1033–1049, 1987

179. Deng S-Y, Goldberg ME, Segraves MA, et al: The effect of unilateral ablation of the frontal eye fields on saccadic performance in the monkey. In: Keller EL, Zee DS, eds. Adaptive Processes in Visual and Oculomotor Systems. Oxford:Pergamon, 1986:201–208

180. Schiller PH, True SD, Conway JL: Deficits in eye movements following frontal eye-field and superior colliculus ablations. J Neurophysiol 44:1175–1189, 1980

181. Lynch JC, McLaren J: Deficits of visual attention and saccadic eye movements after lesions of parietoocipital cortex in monkeys. J Neurophysiol 61:74–90, 1989

182. Tusa RJ, Zee DS, Herdman SJ: Effect of unilateral cerebral cortical lesions on ocular motor behavior in monkeys: saccades and quick phases. J Neurophysiol 56:1590–1625, 1986

183. Pierrot-Deseilligny C, Gautier JC, Loron P: Acquired ocular motor apraxia due to bilateral fronto-parietal infarcts. Ann Neurol 23:199-202, 1988

184. Maunsell J, Van Essen D: Functional properties of neurons in middle temporal visual area of the macaque monkey. I. Selectivity for stimulus direction, speed, and orientation. J Neurophysiol 49:1127-1147, 1983

185. Gottlieb J, MacAvoy M, Bruce C: Unit activity related to smooth pursuit eye movements in rhesus monkey frontal eye fields. Soc Neurosci Abstr 15:1203, 1989

186. Glickstein M, Cohen J, Dixon B, et al: Corticopontine visual projections in macaque monkeys. J Comp Neurol 190:209-229, 1980

187. Suzuki D, Noda H, Kase M: Visual and pursuit eye movement-related activity in posterior vermis of monkey cerebellum. J Neurophysiol 46:1120-1139, 1981

188. Newsome W, Pare E: A Selective impairment of motion perception following lesions of the middle temporal visual area (MT). J Neurosci 8:2201-2211, 1988

189. Morrow MJ, Sharpe JA: Retinoptic and directional deficits of smooth pursuit initiation after posterior cerebral hemispheric lesions. Neurology 43:595–603, 1993

190. Thurston SE, Leigh RJ, Crawford TJ, et al: Two distinct deficits of visual tracking caused by unilateral lesions of cerebral cortex in humans. Ann Neurol 23:266–273, 1988

191. Barton JJ, Sharpe JA, Raymond JE: Directional defects in pursuit and motion perception in humans with unilateral cerebral lesions. Brain 119:1535–1550, 1996

192. Barton JJ, Simpson T, Kiriakopoulos E, et al: Functional MRI of lateral occipitotemporal cortex during pursuit and motion perception. Ann Neurol 40:387–398, 1996

193. May JG, Keller EL, Suzuki DA: Smooth-pursuit eye movement deficits with chemical lesions in the dorsolateral pontine nucleus of the monkey. J Neurophysiol 59:952–977, 1988

194. Thier P, Bachor A, Faiss J, et al: Selective impairment of smooth-pursuit eye movements due to an ischemic lesion of the basal pons. Ann Neurol 29:443–448, 1991

195. Keller EL, Heinen SJ: Generation of smooth-pursuit eye movements: neuronal mechanisms and pathways. Neurosci Res 11:79–107, 1991

196. Johnston JL, Sharpe JA, Morrow MJ: Paresis of contralateral smooth pursuit and normal vestibular smooth eye movements after unilateral brainstem lesions. Ann Neurol 31:495–, 1992502

197. Büttner U, Straube A: The effect of cerebellar midline lesions on eye movements. Neuroophthalmology 15:75–82, 1995

198. Kaufman SR, Abel LA: The effects of distraction on smooth pursuit in normal subjects. Acta Otolaryngol (Stockh) 102:57–64, 1986

199. Leigh RJ, Tusa RJ: Disturbance of smooth pursuit caused by infarction of occipitoparietal cortex. Ann Neurol 17:185–187, 1985

200. Levin S, Luebke A, Zee DS, et al: Smooth pursuit eye movements in schizophrenics. Quantitative measurements with the search-coil technique. J Psychiatr Res 22:195–206, 1988

201. Huebner WP, Leigh RJ, Seidman SH, et al: An investigation of horizontal combined eye-head tracking in patients with abnormal vestibular and smooth pursuit eye movements. J Neurol Sci 116:152–164, 1993

202. Andersen RA, Bracewell RM, Barash S, et al: Eye position effects on visual, memory, and saccade-related activity in areas LIP and 7a of macaque. J Neurosci 10:1176–1196, 1990

203. De Renzi E, Colombo A, Faglioni P, et al: Conjugate gaze paresis in stroke patients with unilateral damage. An unexpected instance of hemispheric asymmetry. Arch Neurol 39:482–486, 1982

204. Kömpf D, Gmeiner H-J: Gaze palsy and visual hemineglect in acute hemisphere lesions. Neuroophthalmology 9:49–53, 1989

205. Sharpe JA, Bondar RL, Fletcher WA: Contralateral gaze deviation after frontal lobe haemorrhage. J Neurol Neurosurg Psychiatry 48:86–88, 1985

206. Steiner I, Melamed E: Conjugate eye deviation after acute hemispheric stroke: delayed recovery after previous contralateral frontal lobe damage. Ann Neurol 16:509–511, 1984

207. Sullivan HC, Kaminski HJ, Maas EF, et al: Lateral deviation of the eyes on forced lid closure in patients with cerebral lesions. Arch Neurol 48:310–311, 1991

208. Sharpe JA, Lo AW, Rabinovitch HE: Control of the saccadic and smooth pursuit systems after cerebral hemidecortication. Brain 102:387–403, 1979

209. Troost BT, Daroff RB, Weber RB, et al: Hemispheric control of eye movements. II. Quantitative analysis of smooth pursuit in a hemispherectomy patient. Arch Neurol 27:449–452, 1972

210. Sharpe JA, Morrow MJ: Cerebral hemispheric smooth pursuit disorders. Neuro-ophthalmology 11:87–98, 1991

211. Cogan DG, Loeb DR: Optokinetic response and intracranial lesions. Arch Neurol Psychiatry 61:183–187, 1949

212. Ebner R, Lopez L, Ochoa S, et al: Vertical ocular motor apraxia. Neurology 40:712–, 1990713

213. Husain M, Stein J: Rezso Balint and his most celebrated case. Arch Neurol 45:89–93, 1988

214. Holmes G: Spasm of fixation. Trans Ophthalmol Soc UK 50:253–262, 1930

215. Johnston JL, Sharpe JA, Morrow MJ: Spasm of fixation: a quantitative study. J Neurol Sci 107:166-171, 1992

216. Fielder AR, Gresty MA, Dodd KL, et al: Congenital ocular motor apraxia. Trans Ophthalmol Soc UK 105:589–598, 1986

217. Harris CM, Shawkat F, Russell-Eggitt I, et al: Intermittent horizontal saccade failure (‘ocular motor apraxia’) in children. Br J Ophthalmol 80:151–158, 1996

218. Zee DS, Yee RD, Singer HS: Congenital ocular motor apraxia. Brain 100:581–599, 1977

219. Cogan DG, Chu FC, Reingold DR, et al: A long-term follow-up of congenital ocular motor apraxia. Neuroophthalmology 1:145–147, 1980

220. Hsu HN, Yang ML, Lai HC: Familial congenital ocular motor apraxia. Chang Gung Med J 25:411–414, 2002

221. Phillips PH, Brodsky MC, Henry PM: Congenital ocular motor apraxia with autosomal dominant inheritance. Am J Ophthalmol 129:820–822, 2000

222. Shawkat FS, Kingsley D, Kendall B, et al: Neuroradiological and eye movement correlates in children with intermittent saccade failure: “Ocular motor apraxia.“rdquo; Neuropediatrics 26:298–305, 1995

223. Rascol O, Sabatini U, Simonetta-Moreau M, et al: Square wave jerks in Parkinsonian syndromes. J Neurol Neurosurg Psychiatry 54:599–602, 1991

224. O'Sullivan JD, Maruff P, Tyler P, et al: Unilateral pallidotomy for Parkinson's disease disrupts ocular fixation. J Clin Neurosci 10:181–185, 2003

225. White OB, Saint-Cyr JA, Tomlinson RD, et al: Ocular motor deficits in Parkinson's disease. II. Control of saccadic and smooth pursuit systems. Brain 106:571–587, 1983

226. Chamberlain W: Restriction of upward gaze with advancing age. Am J Ophthalmol 71:341–346, 1971

227. Demer JL: Pivotal role of orbital connective tissues in binocular alignment and strabismus: the Friedenwald lecture. Invest Ophthalmol Vis Sci 45 :729–738, 2004

228. Rottach KG, Riley DE, DiScenna AO, et al: Dynamic properties of horizontal and vertical eye movements in parkinsonian syndromes. Ann Neurol 39:368–377, 1996

229. Repka MX, Claro MC, Loupe DN, et al: Ocular motility in Parkinson's disease. Pediatr Ophthalmol Strabismus 33:144–147, 1996

230. Crawford T, Goodrich S, Henderson L, et al: Predictive responses in Parkinson's disease: manual key presses and saccadic eye movements to regular stimulus events. J Neurol Neurosurg Psychiatry 52:1033–1042, 1989

231. Lueck CJ, Crawford TJ, Henderson L, et al: Saccadic eye movements in Parkinson's disease. II. Remembered saccades—towards a unified hypothesis? Q J Exp Psychol 45A:211–233, 1992

232. Nakamura T, Bronstein AM, Lueck CJ, et al: Vestibular, cervical and visual remembered saccades in Parkinson's disease. Brain 117:1423–1432, 1994

233. Vermersch AI, Rivaud S, Vidailhet M, et al: Sequences of memory-guided saccades in Parkinson's disease. Ann Neurol 35:487–490, 1994

234. Kimmig H, Haussmann K, Mergner T, et al: What is pathological with gaze shift fragmentation in Parkinson's disease? J Neurol 249:683–692, 2002

235. Vidailhet M, Rivaud S, Gouider-Khouja N, et al: Eye movements in parkinsonian syndromes. Ann Neurol 35:420–426, 1994

236. MacAskill MR, Anderson TJ, Jones RD: Saccadic adaptation in neurological disorders. Prog Brain Res 140:417–431, 2002

237. Bronstein AM, Kennard C: Predictive ocular motor control in Parkinson's disease. Brain 108:925–940, 1985

238. Rascol O, Clanet M, Montastruc JL, et al: Abnormal ocular movements in Parkinson's disease. Evidence for involvement of dopaminergic systems. Brain 112:1193–1214, 1989

239. Waterston JA, Barnes GR, Grealy MA, et al: Abnormalities of smooth eye and head movement control in Parkinson's disease. Ann Neurology 39:749–760, 1996

240. White OW, Saint-Cyr JA, Sharpe JA: Ocular motor deficits in Parkinson's disease. I. The horizontal vestibulo-ocular reflex and its regulation. Brain 106:555–570, 1983

241. Reichert WH, Doolittle J, McDowell FH: Vestibular dysfunction in Parkinson's disease. Neurology 32:1133–1138, 1982

242. Gibson JM, Pimlott R, Kennard C: Ocular motor and manual tracking in Parkinson's disease and the effect of treatment. J Neurol Neurosurg Psychiatry 50:853–860, 1987

243. Blekher T, Siemers E, Abel LA, et al: Eye movements in Parkinson's disease: before and after pallidotomy. Invest Ophthalmol Vis Sci 41:2177–2183, 2000

244. Rivaud-Pechoux S, Vermersch AI, Gaymard B, et al: Improvement of memory guided saccades in parkinsonian patients by high frequency subthalamic nucleus stimulation. J Neurol Neurosurg Psychiatry. 68:381–384, 2000

245. Averbuch-Heller L, Stahl JS, Hlavin ML, et al: Square wave jerks induced by pallidotomy in parkinsonian patients. Neurology 52:185–188, 1999

246. Hotson JR, Langston EB, Langston JW. Saccade responses to dopamine in human MTPT-induced parkinsonism. Ann Neurol 20:456–463, 1986

247. Schultz W, Romo R, Scarnati E, et al: Saccadic reaction times, eye-arm coordination and spontaneous eye movements in normal and MTPT-treated monkeys. Exp Brain Res 78:253–267, 1989

248. Brooks BA, Fuchs AF, Finochio D: Saccadic eye movement deficits in the MTPT monkey model of Parkinson's disease. Brain Res 383:402–407, 1986

249. Leigh RJ, Foley JM, Remler BF, et al: Oculogyric crisis: a syndrome of thought disorder and ocular deviation. Ann Neurol 22:13–17, 1987

250. Liu GT, Carrazana EJ, Macklis JD, et al: Delayed oculogyric crises associated with striatocapsular infarction. J Clin Neuro-ophthalmol 11:198–201, 1991

251. Frankel M, Cummings JL: Neuro-ophthalmic abnormalities in Tourette's syndrome. Functional and anatomical implications. Neurology 34:359–361, 1984

252. FitzGerald PM, Jankovic J, Glaze DG, et al: Extrapyramidal involvement in Rett's syndrome. Neurology 40:293–295, 1990

253. FitzGerald PM, Jankovic J: Tardive oculogyric crisis. Neurology 39:1434–1437, 1989

254. Kim JS, Kim HK, Im JH, et al: Oculogryic crisis and abnormal magnetic resonance imaging signals in bilateral lentiform nuclei. Mov Disord 11:756–758, 1996

255. Nath U, Ben-Shlomo Y, Thomson R, et al: Clinical features and natural history of progressive supranuclear palsy. A clinical cohort study. Neurology 60:910–916, 2003

256. Davis PH, Bergeron C, McLachlan DR: Atypical presentation of progressive supranuclear palsy. Ann Neurol 17:337–343, 1985

257. Quinn N: The “round the houses” sign in progressive supranuclear palsy. Ann Neurol 40:951 , 2003

258. Das V, Leigh R: Visual-vestibular interaction in progressive supranuclear palsy. Vision Res 40:2077–2081, 2000

259. Seemungal B, Faldon M, Revesz T, et al: Influence of target size on vertical gaze palsy in a pathologically proven case of progressive supranuclear palsy. Mov Disord 18:818–822, 2003

260. Garbutt S, Han Y, Kumar A, et al: Vertical optokinetic nystagmus and saccades in normal human subjects. Invest Ophthalmol Vis Sci 44:3833–3841, 2003

261. Mastaglia FL, Grainger KMR: Internuclear ophthalmoplegia in progressive supranuclear palsy. J Neurol Sci 25:303–308, 1975

262. Kuniyoshi S, Riley DE, Zee DS, et al: Distinguishing progressive supranuclear palsy from other forms of Parkinson's disease: evaluation of new signs. Ann NY Acad Sci 956:484–, 2002486

263. Pierrot-Deseilligny C, Rivaud S, Fournier E, et al: Lateral visually-guided saccades in progressive supranuclear palsy. Brain 112:471–487, 1989

264. Goffinet AM, DeVolder AG, Gillian C, et al: Positron tomography demonstrates frontal lobe hypometabolism in progressive supranuclear palsy. Ann Neurol 25:131–139, 1989

265. Litvan I, Dickson D, Büttner-Ennever J, et al: Research goals in progressive supranuclear palsy. Movement Disorders. Mov Disord 15:446–458, 2000

266. Moses HI, Zee DS: Multi-infarct PSP. Neurology 37:1819, 1987

267. Averbuch-Heller L, Paulson G, Daroff R, et al: Whipple's disease mimicking progressive supranuclear palsy: the diagnostic value of eye movement recording. J Neurol Neurosurg Psychiatry 66:532–535, 1999

268. Knox DL, Green WR, Troncosa JC, et al: Cerebral ocular Whipple's disease: a 62 year-old odyssey from death to diagnosis. Neurology 45:617–625, 1995

269. Schwartz MA, Selhorst JB, Ochs AL, et al: Oculomasticatory myorhythmia: a unique movement disorder occurring in Whipple's disease. Ann Neurol 20:677–683, 1986

270. Lowsky R, Archer GL, Fyles G, et al: Diagnosis of Whipple's disease by molecular analysis of peripheral blood. N Engl J Med 331:1343–1346, 1994

271. Fleming JL, Wiesner RH, Shorter RG: Whipple's disease: clinical, biochemical, and histopathological features and assessment of treatment in 29 patients. Mayo Clin Proc 63:539–551, 1988

272. Riley DE, Fogt N, Leigh RJ. The syndrome of ‘pure akinesia’ and its relationship to progressive supranuclear palsy. Neurology 44:1025–1029, 1994

273. Vidailhet M, Rivaud-Pechoux S: Eye movement disorders in corticobasal degeneration. Adv Neurol 82:161–167, 2000

274. Riley DE, Lang AE, Lewis A, et al: Cortical-basal ganglionic degeneration. Neurology 40:1203–1212, 1990

275. Gibb WRG, Luthert PJ, Marsden CD: Corticobasal degeneration. Brain 112:1171–, 19891192

276. Bertholon P, Bronstein AM, Davies RA, et al: Positional down beating nystagmus in 50 patients: cerebellar disorders and possible anterior semicircular canalithiasis. J Neurol Neurosurg Psychiatry 72:366–372 , 2002

277. Brett F, Henson C, Staunton H: Familial diffuse Lewy body disease, eye movement abnormalities, and distribution of pathology. Arch Neurol 59:464–467, 2003

278. de Bruin VM, Lees AJ, Daniel SE: Diffuse Lewy body disease presenting with supranuclear gaze palsy, parkinsonism, and dementia: a case report. Mov Disord 7:355–358, 1992

279. Grant MP, Cohen M, Petersen RB, et al: Abnormal eye movements in Creutzfeldt-Jakob disease. Ann Neurol 34:192–197, 1993

280. Helmchen C, Büttner U: Centripetal nystagmus in a case of Creutzfeldt-Jakob disease. Neuroophthalmology 15:187–192, 1995

281. Collewijn H, Went LN, Tamminga EP, et al: Oculomotor defects in patients with Huntington's disease and their offspring. J Neurol Sci 86:307–320, 1988

282. Lasker AG, Zee DS: Ocular motor abnormalities in Huntington's disease. Vision Res 37:3639–3645, 1997

283. Leigh RJ, Newman SA, Folstein SE, et al: Abnormal ocular motor control in Huntington's disease. Neurology 33:1268–1275, 1983

284. Zangemeister WH, Mueller-Jensen A: The coordination of gaze movements in Huntington's disease. Neuroophthalmology 5:193–206, 1985

285. Rubin AJ, King WM, Reinbold KA, et al: Quantitative longitudinal assessment of saccades in Huntington's disease. J Clin Neuroopthalmol 13:59–66, 1993

286. Kirkwood SC, Siemers E, Bond C, et al: Confirmation of subtle motor changes among presymptomatic carriers of the Huntington disease gene. Arch Neurol 57:1040–1044, 2000

287. Rothlind JC, Brandt J, Zee D, et al: Verbal memory and oculomotor control are unimpaired in asymptomatic adults with the genetic marker for Huntington's disease. Arch Neurol 50:799–802, 1993

288. Reveley MA, Dursun SM, Andrews H: Improvement of abnormal saccadic eye movements in Huntington's disease by sulpiride: A case study. J Psychopharmacol 8:262–265, 1994

289. Troost BT, Troost EG: Functional paralysis of horizontal gaze. Neurology 29:82, 1979

Back to Top