Download Learning in the oculomotor system: from molecules to behavior

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Optogenetics wikipedia , lookup

Feature detection (nervous system) wikipedia , lookup

Artificial neural network wikipedia , lookup

Donald O. Hebb wikipedia , lookup

Neuroeconomics wikipedia , lookup

Perceptual learning wikipedia , lookup

Types of artificial neural networks wikipedia , lookup

Clinical neurochemistry wikipedia , lookup

Neuropsychopharmacology wikipedia , lookup

Long-term depression wikipedia , lookup

Neural engineering wikipedia , lookup

Learning wikipedia , lookup

Recurrent neural network wikipedia , lookup

Premovement neuronal activity wikipedia , lookup

Machine learning wikipedia , lookup

Learning theory (education) wikipedia , lookup

Process tracing wikipedia , lookup

Psychological behaviorism wikipedia , lookup

Concept learning wikipedia , lookup

Metastability in the brain wikipedia , lookup

Neural correlates of consciousness wikipedia , lookup

Development of the nervous system wikipedia , lookup

Neuroplasticity wikipedia , lookup

Cerebellum wikipedia , lookup

Transsaccadic memory wikipedia , lookup

Neuroscience in space wikipedia , lookup

Nonsynaptic plasticity wikipedia , lookup

Superior colliculus wikipedia , lookup

Activity-dependent plasticity wikipedia , lookup

Eyeblink conditioning wikipedia , lookup

Transcript
770
Learning in the oculomotor system: from molecules to behavior
Jennifer L Raymond
A combination of system-level and cellular–molecular
approaches is moving studies of oculomotor learning rapidly
toward the goal of linking synaptic plasticity at specific sites in
oculomotor circuits with changes in the signal-processing
functions of those circuits, and, ultimately, with changes in eye
movement behavior. Recent studies of saccadic adaptation
illustrate how careful behavioral analysis can provide constraints
on the neural loci of plasticity. Studies of vestibulo-ocular
adaptation are beginning to examine the molecular pathways
contributing to this form of cerebellum-dependent learning.
Addresses
Department of Neurobiology, Stanford University School of Medicine,
Sherman Fairchild Building, Room 251, Stanford, California 94305, USA;
e-mail: [email protected]
Current Opinion in Neurobiology 1998, 8:770–776
http://biomednet.com/elecref/0959438800800770
© Current Biology Ltd ISSN 0959-4388
Abbreviations
LTD
long-term depression
VOR
vestibulo-ocular reflex
From an experimental standpoint, oculomotor learning
offers many advantages. Eye movements are easily quantified, and oculomotor learning can be induced within a
single experimental session. The anatomy and physiology
of the oculomotor system are conserved across a wide
range of species. The topography of the oculomotor circuitry allows electrical stimulation at several different sites
to be used to elicit eye movements. Moreover, detailed
studies of identified neurons of the oculomotor system can
be performed in vivo during learning and in reduced preparations. By capitalizing on these advantages, previous
research has described the neural circuits mediating the
different types of eye movement behavior in considerable
detail and has analyzed the computational or signal-processing functions of different loci in those circuits [1•,2•].
These detailed anatomical and physiological studies pave
the way for efforts to identify the synaptic sites of plasticity and the cellular–molecular changes associated with
learning, and to determine how those cellular–molecular
changes are induced, how they alter neural signal processing in oculomotor circuits, and how the altered signal
processing ultimately results in a learned change in eye
movement behavior.
Introduction
Studies of the neural mechanisms of learning generally
have proceeded on two separate and parallel levels. At
one level, neuroanatomical structures involved in particular types of learning have been identified. At another
level, molecular pathways involved in synaptic plasticity
have been identified at a number of sites in the brain. A
major challenge in neuroscience over the coming years
will be to bridge these levels of analysis to create a comprehensive understanding of the interrelationship
between cellular–molecular and circuit properties in the
regulation of behavior by learning. The oculomotor system is an experimental system particularly well suited to
this challenge.
Eye movements can be categorized into several types
of reflexive and voluntary movements — the vestibuloocular reflex (VOR), the optokinetic reflex, ocular
following, smooth pursuit, vergence, saccades, and gaze
holding [1•,2 •]. These eye movements function to
acquire visual targets and to stabilize images on the retina for improved visual acuity. The accuracy required for
these functions is maintained by motor learning throughout development, aging, and injury. Some form of
learning has been demonstrated in each type of eye
movement. (The various forms of learning in the oculomotor system have been collectively referred to as
oculomotor plasticity. For clarity, I use an alternate term,
oculomotor learning, to refer to behavioral changes in
eye movements, and I reserve the term plasticity to refer
to changes in the nervous system.)
The present review highlights recent advances in studies of
two forms of oculomotor learning — saccadic adaptation and
VOR adaptation. The studies of saccadic adaptation illustrate
how behavioral analysis can constrain sites of plasticity by
establishing the functional relationship between plasticity and
the signal-processing functions of the oculomotor pathways.
For VOR adaptation, anatomical sites of plasticity have been
localized, and initial attempts have been made to identify the
synaptic loci and cellular–molecular pathways involved in this
form of learning. In addition, three new experimental preparations are described that should provide powerful tools for
future attempts to link systems-level and cellular–molecular
analyses of oculomotor learning. Finally, general issues in
learning that can be explored using the integrated approaches
available in the oculomotor system are discussed.
Learned changes in saccade gain
A saccade is a rapid change in the position of the eye that
functions to bring visual targets onto the fovea. The gain of a
saccade is defined as the ratio of saccade amplitude to target
displacement. Adaptive changes in saccade gain can be
induced by experimental manipulations that cause saccades
to consistently miss their targets. For example, weakening
the eye muscles causes saccades to consistently fall short of
visual targets. If the muscles of one eye are weakened,
monocular viewing with the weak eye results in a gradual
increase in the gain of saccades in both eyes, and monocular
viewing with the normal eye reverses the increase in gain [3].
A second paradigm for inducing an increase (or decrease) in
saccade gain is to train with a visual target that consistently
Learning in the oculomotor system: from molecules to behavior Raymond
jumps forward (or backward) during saccades toward the target [4]. With either saccadic adaptation paradigm, changes in
gain can take place within a few hundred saccades [5••] and,
therefore, are easily induced within a single experimental
session. In the absence of training to reverse adaptation, the
changes can last at least 20 h in the dark [6••].
Recent studies of saccadic adaptation illustrate how careful behavioral analysis can provide constraints on the
neural loci of plasticity. For example, saccadic adaptation
is specific for the amplitude and direction in three-dimensional space of the saccades elicited during training (see
e.g. [6••,7•,8–11]). This specificity indicates that plasticity
must take place at a site in the brain in which saccades
with different amplitudes or directions are encoded in
separate, or at least partially separate, neural pathways. On
the other hand, saccadic adaptation induced using visual
targets generalizes to saccades made to auditory targets,
indicating that plasticity must take place at a site where
commands for saccades elicited by auditory and visual targets are carried by the same neural pathways [10]. In at
least some cases, the patterns of generalization versus
specificity depend on the particular training paradigm.
For example, saccadic adaptation can generalize across
orbital position, or it can exhibit orbital position specificity, depending on the training paradigm [5••,10,11].
Patterns of specificity and generalization are, on the
whole, similar in human and monkey, although a few differences have been reported. For example, in monkeys,
changes in the gain of saccadic eye movements induced
with the head fixed transfer to both the eye- and headmovement components of head-free gaze saccades [12••],
whereas in humans, no generalization to the head-movement component has been reported [13].
This type of behavioral analysis constrains the site(s) of
plasticity to loci in which the neural responses exhibit
patterns of specificity and generalization that match
those of the learned changes in behavior. One locus in the
circuit for saccades that seems to meet the constraints
defined by the behavioral experiments is the superior colliculus. However, recordings from the superior colliculus
during saccadic adaptation revealed no change in the
responses of saccade-related burst neurons to visual targets [14,15••], suggesting that saccadic adaptation does
not take place upstream of the colliculus. Moreover,
adaptation does not generalize to eye movements produced by electrical stimulation of the superior colliculus
[16,17], suggesting that saccadic adaptation does not
occur downstream of the colliculus. Together, these
results suggest that saccadic adaptation occurs in a pathway parallel to the one through the superior colliculus.
A likely candidate for the site of plasticity mediating saccadic adaptation is the cerebellum. Neurons in the
cerebellar vermis and caudal fastigial nucleus burst during
saccades, and electrical stimulation of these mid-line
cerebellar structures elicits saccades. Moreover, mid-line
771
cerebellar lesions produce saccadic dysmetria and abolish
saccadic adaptation (reviewed in [18]). These results are
consistent with a contribution of the cerebellum to both
the induction and storage of saccadic adaptation.
Recordings of saccade-related signals in the cerebellum
before, during, and after saccadic adaptation should help
elucidate its role.
Learned changes in vestibulo-ocular reflex gain
Learned changes in the gain of another type of eye movement, the VOR, have been reported in a broad range of
species. Studies of this form of motor learning, called VOR
adaptation, are beginning to combine systems-level analyses with more cellular–molecular approaches.
The VOR is an eye movement driven by vestibular signals. The VOR functions to stabilize images on the
retina by producing eye movements in the opposite
direction from head movements. The gain of the VOR is
defined as the ratio of eye speed to head speed. As with
saccadic adaptation, experimentally manipulated optical
conditions can induce learned changes in the gain of the
VOR [19]: for example, if head movements are paired
with movements of the visual surround in the same
direction as the head movements, a gradual reduction in
the gain of the VOR is induced; whereas, if head movements are paired with movements of the visual surround
in the opposite direction from the head movements, a
gradual increase in the gain of the VOR is induced.
The circuit for the VOR contains several parallel pathways
from the vestibular afferents that encode head movements
to the motor neurons that move the eye. After years of controversy, it is now widely accepted that VOR adaptation is
associated with plasticity at two sites — in the vestibular
nuclei and in a pathway through the cerebellar cortex
[20,21•,22]. Why does an apparently simple change in
reflex gain require more than one site of plasticity? One
hypothesis suggests that plasticity in the vestibular nuclei
may be primarily responsible for the gain changes per se,
whereas plasticity in the cerebellar cortex may function to
regulate eye movement dynamics [20,23]. A second
hypothesis, suggested by recent lesion studies, is that the
cerebellar cortex may be more important for changes in the
low-frequency components of the VOR, whereas the
vestibular nuclei may be more important for changes in the
high-frequency components of the VOR [24,25•]. These
hypotheses need not be mutually exclusive.
The region of the cerebellar cortex that has been implicated in VOR adaptation is the floccular complex, which
comprises the flocculus and adjacent ventral paraflocculus. Physiological studies have not demonstrated any
qualitative differences between the responses of Purkinje
cells in the ventral paraflocculus and the responses of
Purkinje cells in the flocculus during the VOR before or
after adaptation (for a discussion, see [26]). However,
anatomical studies have reported some differences in the
772
Motor systems
afferent and efferent projections of the flocculus and ventral paraflocculus, which raise the possibility that the two
structures may make somewhat different contributions to
the VOR [27,28•,29•].
To understand VOR adaptation, it will be important to
specify more precisely which synapses in the vestibular
pathways through the cerebellar cortex and vestibular
nuclei are modified during learning and to identify the
cellular–molecular mechanisms that produce those
changes. Some clues are provided by in vitro studies,
which have begun to identify forms of synaptic plasticity
in the circuit for the VOR. One particular form of plasticity in the cerebellar cortex has received the most
attention, long-term depression of synapses from parallel
fibers to Purkinje cells, known as cerebellar LTD [30]. To
evaluate the potential contribution of this form of synaptic plasticity to learning, several studies have compared
the behavioral sensitivity of VOR adaptation to various
pharmacological manipulations with the sensitivity of
cerebellar LTD to similar pharmacological manipulations
in vitro. A number of parallels have been found. Systemic
administration of (6R)-5,6,7,8-tetrahydro-L-biopterin
(R-THBP) produces an increase in VOR gain and
occludes induction of further gain increases with behavioral training [31•]. R-THBP facilitates activation of
guanylate cyclase in the cerebellum, which, in turn, has
been implicated in the induction of cerebellar LTD in
vitro. Likewise, inhibition of protein kinase C (PKC) in
Purkinje cells blocks both cerebellar LTD in vitro and
VOR adaptation in vivo [32••].
These pharmacological similarities are consistent with a
contribution of cerebellar LTD to VOR adaptation.
However, in vivo recording results are difficult to reconcile with traditional hypotheses [22,30] about the role of
cerebellar LTD in VOR adaptation. Measured changes
in the vestibular sensitivity of Purkinje cells accompanying VOR adaptation are in the opposite direction from
that predicted by LTD [26,33,34]. In addition, the patterns of neural activity present in vivo during VOR
adaptation would, in some cases, fail to trigger cerebellar
LTD in the appropriate synapses to account for the
changes in behavior [35••,36••]. Further experiments
that combine in vivo recordings with pharmacological or
molecular manipulations are needed to evaluate whether
and how cerebellar LTD might contribute to VOR adaptation [37••,38••].
Future studies must also evaluate whether VOR adaptation involves plasticity in the cerebellar cortex at
synapses other than those from parallel fibers to Purkinje
cells. Synaptic plasticity has been observed between
cerebellar inhibitory interneurons and Purkinje cells,
and could contribute to learning in the VOR [39,40].
Finally, future research needs to characterize the plasticity mechanisms in the vestibular nucleus that
contribute to VOR adaptation [20,21 •,35 ••,41 •,42 •].
These studies will be facilitated by new experimental
preparations that have been introduced recently.
Promising new experimental preparations
Several new experimental preparations for studying the
oculomotor system have been introduced recently and
should prove extremely useful for linking the different
forms of oculomotor learning with their underlying cellular–molecular mechanisms.
First amongst the new experimental preparations is the study
of oculomotor learning in the rodent [32••,43••,44•,45•].
The mouse VOR undergoes a 50% increase in gain after
one hour of behavioral training [32••,43••]. Use of a mouse
model system makes it possible to apply the powerful
molecular–genetic manipulations that are becoming
increasingly available to the study of the molecular pathways involved in VOR adaptation. For example, transgenic
mice have been used to demonstrate the dependence of
VOR adaptation on intact PKC signaling pathways in
Purkinje cells [32••]. The mouse also exhibits a robust
optokinetic reflex [32••,43••,46], although learning has not
yet been examined in the optokinetic reflex of this species
(see Note added in proof).
A second new experimental preparation is the zebrafish,
another species that is amenable both to molecular–genetic approaches and to studying oculomotor learning. The
VOR, the optokinetic reflex and saccadic eye movements
of zebrafish are mature by 5 days post-fertilization
[47,48••]. Oculomotor learning has not yet been examined
in zebrafish, but it has been described in other species of
fish. Zebrafish have a rapid life cycle, and can be easily
raised in large numbers, making them well suited for mutagenesis screens. Because eye movements are readily
measured in zebrafish with video analysis, a screen for
mutants with a loss of oculomotor learning seems quite
feasible. In addition, the rapid development of the
zebrafish suggests that it could be used to relate gene
expression in a tissue- and stage-specific manner to the
development of oculomotor learning.
Finally, a guinea pig isolated whole brain in vitro preparation that may serve to bridge in vitro and in vivo studies has
recently been described [49,50•,51•]. This preparation
offers stable intracellular recordings while keeping neural
circuits intact, so that neurons with well-defined intrinsic
membrane properties can be placed in a functional context. An in vitro analog of the VOR can be elicited in this
preparation by electrically stimulating the vestibular nerve
and recording the neural responses in the abducens nerve.
Likewise, an in vitro analog of visually driven eye movements, such as the optokinetic reflex, can be elicited by
stimulating the optic tract and recording from the
abducens nerve. If an in vitro analog of learning could be
induced in these in vitro eye movement ‘behaviors’, it
could provide a powerful tool for linking systems-level and
cellular analyses of plasticity.
Learning in the oculomotor system: from molecules to behavior Raymond
General issues in learning can be studied in
the oculomotor system
Eye movements exhibit an extensive repertoire of learning. Saccades and the VOR exhibit learned changes in
direction and dynamics as well as learned changes in gain
(e.g. [3,8,52–54]), and both types of eye movement exhibit non-associative changes in gain (e.g. habituation or
central fatigue) in addition to the associative changes
described above [55–58]. Other types of eye movement —
such as smooth pursuit, ocular following, the optokinetic
reflex, and vergence — also exhibit learning (e.g. [59–62];
for more recent results, see [63–68]). Thus, different forms
of learning can be compared within a single type of eye
movement, and similar forms of learning can be compared
across different types of eye movement. Many of the fundamental issues in motor learning are experimentally
accessible in one or more eye movement systems. Below, I
discuss just a few such issues.
Are there multiple mechanisms for accomplishing the
same behavioral change?
A decrease in the gain of saccades or the VOR can be
induced through habituation/central fatigue, through associative training to decrease gain, or during the recovery
from increase gain training induced by a return to normal
viewing [6••,55,56,69]. Are these three forms of gain
decrease mediated by the same synaptic modifications? A
recent study of saccadic adaptation shows the time course
of decrease gain adaptation and recovery from increase
gain to be similar, consistent with the idea of a common
mechanism [6••]. In the VOR, lesion studies suggest that
habituation and adaptive decreases in gain are mediated by
different regions of the cerebellum and are therefore distinct processes (see e.g. [70]).
Similarly, increases in VOR gain induced with optical
manipulations can be compared with increases in VOR
gain during recovery from unilateral vestibular lesions, a
process known as vestibular compensation [71,72]. There
is evidence that these two forms of VOR gain increase
have different mechanisms [73], and comparison of the two
could provide insight into how the plasticity mechanism
used to achieve a particular behavioral change is selected
from a number of potentially available mechanisms.
What are the mechanisms for modification of behavior
on different time scales?
Several types of oculomotor learning can be induced either
within hours (using ‘short-term’ training paradigms within
a single experimental session) or over several days (using
‘long-term’ training paradigms in the home cage). To what
extent are the mechanisms of the short- and long-term
forms of learning similar? If they are different, what happens at the molecular level during the transition from
short-term to long-term learning?
Eye movements can also undergo immediate, on-line
modifications of gain. A recent study demonstrates that,
773
under appropriate conditions, monkeys can simultaneously store in long-term memory two motor sets that use
different oculomotor ranges, and they can rapidly switch
between them [74••]. On-line modifications also can be
produced by factors such as fixation distance, strategy, or
alertness (see e.g. [6••,75]). Do these on-line modifications
share sites of action or cellular–molecular effector systems
with short- and long-term learning?
What are the neural signals that drive learning?
Sensory stimuli that drive learning must be encoded in
patterns of neural activity at the site(s) of plasticity.
Recent studies of VOR adaptation have begun to evaluate
the features of these patterns that are necessary and sufficient to trigger plasticity by comparing the patterns
present during a broad range of stimuli that induce VOR
adaptation. This type of analysis can map the protocols
used to induce synaptic plasticity in vitro onto the induction of learning in vivo [35••,36••].
An important consideration in addressing this issue is the
timing of neural signals that trigger plasticity. For most
movements, error signals about the accuracy of a movement are delayed relative to the neural activity that drives
the movement. How is this delayed feedback used to modify the synapses responsible for the error? Biologically
constrained computational models of saccadic adaptation,
VOR adaptation, and predictive pursuit suggest that in the
cerebellum, error signals carried by climbing fibers may
interact with a delayed ‘eligibility trace’ related to previous
parallel fiber activity that contributed to the erroneous
movement [35••,76,77•].
Conclusions
The oculomotor system offers an excellent opportunity to
study neural plasticity in the context of a wide range of
well-defined learning paradigms. Studies of VOR adaptation have begun to combine molecular–genetic approaches
with more traditional, systems-level approaches, with the
goal of establishing causal links between synaptic plasticity and changes in behavior. Progress in identifying sites of
plasticity for saccadic adaptation and other forms of oculomotor learning will soon make it possible to apply similar
approaches to a wide range of eye movement behaviors.
The many experimental advantages of the oculomotor system enumerated in this review make it an outstanding
model for examining fundamental issues in motor learning,
from the level of molecules through to behavior.
Note added in proof
Recently, a form of learning that depends on an intact flocculus was demonstrated in the optokinetic reflex of the
mouse [78].
Acknowledgements
Thanks to A Churchland, M Churchland, M Kahlon, N Priebe, and
H Rambold for critical comments on the manuscript and to S Lisberger for
his support. The author’s work is funded by National Institutes of Health
grants DC03342 and EY03878.
774
Motor systems
References and recommended reading
Papers of particular interest, published within the annual period of review,
have been highlighted as:
• of special interest
•• of outstanding interest
1. Büttner-Ennever JA, Horn AK: Anatomical substrates of oculomotor
•
control. Curr Opin Neurobiol 1997, 7:872-879.
Review of the neural circuitry of the oculomotor system, with an emphasis on the extent to which circuits for different types of eye movement
remain segregated.
2. Ilg UJ: Slow eye movements. Prog Neurobiol 1997,
•
53:293-329.
Review of the behavioral, anatomical and physiological properties of the different types of slow eye movements.
3.
Optican LM, Robinson DA: Cerebellar-dependent adaptive
control of primate saccadic system. J Neurophysiol 1980,
44:1058-1076.
4.
McLaughlin SC: Parametric adjustment in saccadic eye
movements. Percept Psychophys 1967, 2:359-362.
5.
••
Scudder CA, Batouria EY, Tunder GS: Comparison of two methods
of producing adaptation of saccade size and implications for the
site of plasticity. J Neurophysiol 1998, 79:704-715.
Demonstrates that saccadic adaptation induced with a muscle-weakening
paradigm takes place at a similar rate to adaptation induced with intra-saccadic target steps if the two adaptation paradigms are matched for number
of visual targets. The results remove the rationale for hypothesizing that the
two paradigms employ different neural mechanisms, which had been proposed on the basis of the longer time course of adaptation that was generally observed with the muscle-weakening paradigm.
6.
••
Straube A, Fuchs AF, Usher S, Robinson FR: Characteristic of
saccadic gain adaptation in rhesus macaques. J Neurophysiol
1997, 77:874-895.
A thorough characterization of several features of saccadic adaptation, including time course, spatial specificity, and effects on the duration, velocity and
latency of saccades. These features constrain the neural loci of plasticity.
7.
Chaturvedi V, van Gisbergen JA: Specificity of saccadic adaptation
•
in three-dimensional space. Vision Res 1997, 37:1367-1382.
Demonstrates the ability of the saccadic system to adopt different gains for
saccades to different depth planes.
8.
Mack A, Fendrich R, Pleune J: Adaptation to an altered relation
between retinal image displacements and saccadic eye
movements. Vision Res 1978, 18:1321-1327.
9.
Miller JM, Anstis T, Templeton WB: Saccadic plasticity: parametric
adaptive control by retinal feedback. J Exp Psychol [Hum Percept
Perform] 1981, 7:356-366.
10. Frens MA, van Opstal AJ: Transfer of short-term adaptation in human
saccadic eye movements. Exp Brain Res 1994, 100:293-306.
11. Albano JE: Adaptive changes in saccade amplitude: oculocentric
or orbitocentric mapping? Vision Res 1996, 36:2087-2098.
12. Phillips JO, Fuchs AF, Ling L, Iwamoto Y, Votaw S: Gain adaptation
•• of eye and head movement components of simian gaze shifts.
J Neurophysiol 1997, 78:2817-2821.
Reports generalization of saccadic adaptation induced with the head fixed
to both the head- and eye-movement components of head-free gaze shifts.
The authors found that saccadic adaptation did not alter the way gaze saccades of a given size were parsed into separate eye- and head-movement
components. These results are consistent with the hypothesis that in monkeys, saccadic adaptation occurs upstream of the parsing of the commands
for gaze shifts into separate head- and eye-movement components.
13. Kröller J, Péllison D, Prablanc C: On the short-term adaptation of
eye saccades and its transfer to head movements. Exp Brain Res
1996, 111:477-482.
14. Goldberg ME, Musil SY, Fitzgibbon EJ, Smith M, Olson CR: The role
of the cerebellum in the control of saccadic eye movements. In
Role of the Cerebellum and Basal Ganglia in Voluntary Movements.
Edited by Mano N, Hamada I, DeLong MR. Amsterdam: Elsevier;
1993:203-211.
15. Frens MA, Van Opstal AJ: Monkey superior colliculus activity
•• during short-term saccadic adaptation. Brain Res Bull 1997,
43:473-483.
During saccadic adaptation, responses of saccade-related burst neurons in
the intermediate and deep layers of the superior colliculus remained appropriate for the saccade that was required to foveate the initial target, rather
than for the adapted saccade amplitude. These results suggest that pathways
upstream from the colliculus are not responsible for the altered saccadic gain.
16. Fitzgibbon EJ, Goldberg ME, Segraves MA: Short term saccadic
adaptation in the monkey. In Adaptive Processes in the Visual and
Oculomotor Systems. Edited by Keller E, Zee DS. Oxford, UK:
Pregammon; 1986:329-333.
17.
Melis BJM, Van Gisbergen JAM: Short-term adaptation of
electrically-induced saccades in monkey superior colliculus.
J Neurophysiol 1996, 76:1744-1758.
18. Robinson FR: Role of the cerebellum in movement control and
adaptation. Curr Opin Neurobiol 1995, 5:755-762.
19. Gonshor A, Melvill Jones G: Extreme vestibulo-ocular adaptation
induced by prolonged optical reversal of vision. J Physiol Lond
1976, 256:381-414.
20. Lisberger SG: Neural basis for motor learning in the
vestibuloocular reflex of primates. III. Computational and
behavioral analysis of the sites of learning. J Neurophysiol 1994,
72:974-998.
21. Highstein SM, Partsalis A, Arikan R: Role of the Y-group of the
•
vestibular nuclei and flocculus of the cerebellum in motor learning
of the vestibulo-ocular reflex. Prog Brain Res 1997, 114:383-397.
Reviews evidence for sites of plasticity associated with adaptation of the vertical VOR. Studies of the horizontal and vertical VOR have independently
suggested that VOR adaptation involves plasticity both in the vestibular
nuclei and in a pathway through the cerebellar cortex.
22. Ito M: Cerebellar control of the vestibulo-ocular reflex — around
the flocculus hypothesis. Annu Rev Neurosci 1982, 5:275-298.
23. Lisberger SG, Sejnowski TJ: Motor learning in a recurrent network
model based on the vestibulo-ocular reflex. Nature 1992,
360:159-161.
24. Pastor AM, de la Cruz RR, Baker R: Cerebellar role in adaptation of the
goldfish vestibuloocular reflex. J Neurophysiol 1994, 72:1383-1394.
25. McElligott JG, Beeton P, Polk J: Effect of cerebellar inactivation by
•
lidocaine microdialysis on the vestibuloocular reflex in goldfish.
J Neurophysiol 1998, 79:1286-1294.
Inactivation of the cerebellum with lidocaine blocked the induction and retention
of VOR adaptation during vestibular stimulation at 0.125 Hz. This study confirms
the dependence of adaptation of the low-frequency components of the VOR on
an intact cerebellum, which was previously reported using surgical lesions [24].
26. Lisberger SG, Pavelko TA, Bronte-Stewart HM, Stone LS: Neural
basis for motor learning in the vestibulo-ocular reflex of primates:
II. Changes in the responses of horizontal gaze velocity Purkinje
cells in the cerebellar flocculus and ventral paraflocculus.
J Neurophysiol 1994, 72:954-973.
27.
Gerrits NM, Voogd J: The topographical organization of climbing and
mossy fiber afferents in the flocculus and ventral paraflocculus in
the rabbit, cat, and monkey. Exp Brain Res 1989, 17(suppl):26-29.
28. Nagao S, Kitamura T, Nakamura N, Hiramatsu T, Yamada J: Location
•
of efferent terminals of the primate flocculus and ventral
paraflocculus revealed by anterograde axonal transport methods.
Neurosci Res 1997, 27:257-269.
See annotation [29•].
29. Nagao S, Kitamura T, Nakamura N, Hiramatsu T, Yamada J:
•
Differences in the primate flocculus and ventral paraflocculus in
the mossy and climbing fiber input organization. J Comp Neurol
1997, 382:480-498.
This paper (see also [27,28•]) is part of a current controversy about the relative
roles of the flocculus and ventral paraflocculus in the VOR. The authors report
overlapping but non-identical patterns of afferent and efferent projections of
these two cerebellar regions. They interpret the differences as support for the
hypothesis that the flocculus is mainly involved in the VOR whereas the ventral
paraflocculus is mainly involved in smooth pursuit eye movements. However,
other interpretations are possible. For example, the flocculus and ventral
paraflocculus could make identical contributions to the VOR that depend on the
anatomical connections they have in common, but they could have additional,
different functions that require different anatomical connections.
30. Ito M: Long-term depression. Annu Rev Neurosci 1989, 12:85-102.
31. Nagao S, Kitazawa H, Osanai R, Hiramatsu T: Acute effects of
•
tetrahydrobiopterin on the dynamic characteristics and
adaptability of vestibulo-ocular reflex in normal and flocculus
lesioned rabbits. Neurosci Lett 1997, 231:41-44.
Tetrahydrobiopterin produces an increase in VOR gain, which the authors
suggest might be mediated by effects on guanylate cyclase or on
monoamines in the cerebellum.
Learning in the oculomotor system: from molecules to behavior Raymond
32. De Zeeuw CI, Hansel C, Bian F, Koekkoek SK, van Alphen AM,
•• Linden DJ, Oberdick J: Expression of a protein kinase C inhibitor in
Purkinje cells blocks cerebellar LTD and adaptation of the
vestibulo-ocular reflex. Neuron 1998, 20:495-508.
This pioneering work uses transgenic mice to investigate the molecular pathways in Purkinje cells necessary for VOR adaptation. The pcp-2 (L7) gene
promoter was used to selectively express a pseudosubstrate PKC inhibitor
in Purkinje cells. Mice expressing the transgene exhibited no adaptation of
the VOR, and cerebellar slice cultures from these animals exhibited no LTD.
This work demonstrates the power and potential of applying
molecular–genetic approaches to the oculomotor system. In addition, it has
generated discussion about how molecular–genetic approaches might best
be combined with systems-level approaches to study the neural mechanisms
of learning [37••,38••].
33. Miles FA, Braitman DJ, Dow BM: Long-term adaptive changes in
primate vestibulo-ocular reflex. IV. Electrophysiological
observations in the flocculus of adapted monkeys. J Neurophysiol
1980, 43:1477-1493.
34. Arikan R, Highstein SM: Modulation changes of individual Purkinje
cells in the flocculus vertical zone during adaptation of the
vestibulo ocular reflex. Soc Neurosci Abstr 1997, 23:749.
35. Raymond JL, Lisberger SG: Neural learning rules for the
•• vestibulo-ocular reflex. J Neurosci 1998, 18:9112-9129.
This work compares the induction of learning in vivo to the requirements for
induction of synaptic plasticity in vitro. It examines the patterns of neural
activity present in the cerebellar cortex and vestibular nuclei during a range
of stimuli that induce VOR adaptation. Patterns of neural activity that are
consistently present during learning are identified as candidates for the in
vivo trigger of synaptic plasticity.
36. Raymond JL, Lisberger SG: Multiple subclasses of Purkinje cells in
•• the primate floccular complex provide similar signals to guide
learning in the vestibulo-ocular reflex. Learn Mem 1997,
3:503-518.
This study demonstrates that different subclasses of Purkinje cells in the
floccular complex all receive similar information about the change in VOR
gain required to improve image stability on the retina. These findings extend
the results about the neural signals available to guide learning obtained for
horizontal gaze velocity Purkinje cells [35••].
37. Lisberger SG: Cerebellar LTD: a molecular mechanism of
•• behavioral learning? Cell 1998, 92:701-704.
Cautions that a loss of VOR adaptation in mutant mice can only be interpreted by analyzing the detailed workings of the neural circuit for the VOR
in vivo, in both wild-type and mutant mice. A similar caution applies to any
attempt to use genetically manipulated animals to establish molecular mechanisms of learning.
38. Mauk MD, Garcia KS, Medina JF, Steele PM: Does cerebellar LTD
•• mediate motor learning? Toward a resolution without a smoking
gun. Neuron 1998, 20:359-362.
Proposes five criteria that could be used to evaluate whether cerebellar LTD
is causally related to cerebellum-mediated motor learning. The same criteria
could be used, in general, to evaluate the contribution of any given form of
synaptic plasticity to a particular form of learning.
39. Llano I, Leresche N, Marty A: Calcium entry increases the sensitivity
of cerebellar Purkinje cells to applied GABA and decreases
inhibitory synaptic currents. Neuron 1991, 6:565-574.
40. Kano M, Rexhausen U, Dressen J, Konnerth A: Synaptic excitation
produces a long-lasting rebound potentiation of inhibitory
synaptic signals in cerebellar Purkinje cells. Nature 1992,
356:601-604.
41. Takahashi Y, Kubo T: Excitatory synaptic transmission in the rat
•
medial vestibular nucleus. Acta Otolaryngol (Stockh) 1997,
528(suppl):56-58.
Reviews results from in vitro brainstem slice preparations on the neuropharmacology of excitatory synaptic transmission in the medial vestibular nucleus.
42. Sastry BR, Morishita W, Yip S, Shew T: GABA-ergic transmission in
•
deep cerebellar nuclei. Prog Neurobiol 1997, 53:259-271.
Review of GABAergic transmission and its plasticity in the deep cerebellar
nuclei. Plasticity mechanisms in the vestibular nuclei may be similar to those
in the deep cerebellar nuclei, because the vestibular nuclei function as the
deep cerebellar nuclei for the floccular complex.
43. Koekkoek SKE, Alphen AMV, Galjart N, Burg JVD, Grosveld F,
•• De Zeeuw CI: Gain adaptation and phase dynamics of
compensatory eye movements in mice. Genes Function 1997,
1:175-190.
This is the first study of VOR adaptation in the mouse. It compares wild-type
mice with lurcher mice, a mutant that lacks Purkinje cells, and with mice with
surgical lesions of the flocculus. The authors demonstrate that in mice, as in
other species, VOR adaptation depends on an intact cerebellum.
775
44. Cransac H, Peyrin L, Farhat F, Cottet-Emard JM, Pequignot JM,
•
Reber A: Brain monoamines and optokinetic performances in
pigmented and albino rats. Comp Biochem Physiol [A] 1997,
116:341-349.
Reveals differences in the VOR, optokinetic reflex, and patterns of
monoamine distribution in the medial vestibular nuclei of pigmented DA-HAN
versus albino Sprague-Dawley rats.
45. Quinn KJ, Rude SA, Brettler SC, Baker JF: Chronic recording of the
•
vestibulo-ocular reflex in the restrained rat using a permanently
implanted scleral search coil. J Neurosci Methods 1998, 80:201-208.
Describes an implant that allows eye movements to be monitored in individual rats over several weeks.
46. Mitchiner JC, Pinto LH, Vanable JW Jr: Visually evoked eye
movements in the mouse (Mus musculus). Vision Res 1976,
16:1169-1171.
47.
Easter SS Jr, Nicola GN: The development of vision in the
zebrafish (Danio rerio). Dev Biol 1996, 180:646-663.
48. Easter SS Jr, Nicola GN: The development of eye movements in
•• the zebrafish (Danio rerio). Dev Psychobiol 1997, 31:267-276.
Describes the rapid development of the VOR, the optokinetic reflex, and saccades in normal and dark-reared zebrafish. Reveals the potential utility of the
zebrafish for studying the oculomotor system.
49. Muhlethaler M, de Curtis M, Walton K, Llinas R: The isolated and
perfused brain of the guinea-pig in vitro. Eur J Neurosci 1993,
5:915-926.
50. Babalian A, Vibert N, Assie G, Serafin M, Muhlethaler M, Vidal PP:
•
Central vestibular networks in the guinea pig: functional
characterization in the isolated whole brain in vitro. Neuroscience
1997, 81:405-426.
Demonstrates the utility of the guinea pig whole brain in vitro preparation for
studying the physiological and pharmacological properties of neurons in the
VOR circuit.
51. Vibert N, De Waele C, Serafin M, Babalian A, Muhlethaler M, Vidal PP:
•
The vestibular system as a model of sensorimotor
transformations. A combined in vivo and in vitro approach to
study the cellular mechanisms of gaze and posture stabilization
in mammals. Prog Neurobiol 1997, 51:243-286.
Reviews the use of combined in vivo/in vitro approaches to study vestibular
circuits in the guinea pig.
52. Schultheis LW, Robinson DA: Directional plasticity of the
vestibuloocular reflex in the cat. Ann NY Acad Sci 1981,
374:504-512.
53. Angelaki DE, Hess BJM: Visually induced adaptation in threedimensional organization of primate vestibuloocular reflex.
J Neurophysiol 1998, 79:791-807.
54. Raymond JL, Lisberger SG: Behavioral analysis of signals that
guide learned changes in the amplitude and dynamics of the
vestibulo-ocular reflex. J Neurosci 1996, 16:7791-7802.
55. Straube A, Robinson FR, Fuchs AF: Decrease in saccadic
performance after many visually guided saccadic eye movements
in monkeys. Invest Opthal Vis Sci 1997, 38:2810-2816.
56. Dodge R: Habituation to rotation. J Exp Psychol 1923, 6:1-35.
57.
Dow EE, Anastasio TJ: Analysis and neural network modeling of
the nonlinear correlates of habituation in the vestibulo-ocular
reflex. J Comput Neurosci 1998, 5:171-190.
58. Torte MP, Clément G, Courjon J-H, Magenes G: Absence of
vestibular habituation of the vestibulo-ocular reflex in the vertical
plane in the cat. Exp Brain Res 1997, 116:73-82.
59. Optican LM, Zee DS, Chu FC: Adaptive responses to ocular
muscle weakness in human pursuit and saccadic eye
movements. J Neurophysiol 1985, 54:110-122.
60. Miles FA, Kawano K: Short-latency ocular following responses of
monkey. III. Plasticity. J Neurophysiol 1986, 56:1381-1396.
61. Collewijn H, Grootendorst AF: Adaptation of optokinetic and
vestibulo-ocular reflexes to modified visual input in the rabbit. In
Progress in Brain Research. Reflex Control of Posture and Movement.
Edited by Granir R, Pompeiano O. Amsterdam: Elsevier; 1979:772-781.
62. Oohira A, Zee DS: Disconjugate ocular motor adaptation in rhesus
monkey. Vision Res 1992, 32:489-497.
63. Ogawa T, Fujita M: Adaptive modifications of human post-saccadic
pursuit eye movements induced by a step-ramp-ramp paradigm.
Exp Brain Res 1997, 116:83-96.
776
Motor systems
64. Jardon BL, Bonaventure N: Involvement of NMDA in a plasticity
phenomenon observed in the adult frog. Vision Res 1997,
37:1511-1524.
65. Marsh E, Baker R: Normal and adapted visuoocular reflexes in
goldfish. J Neurophysiol 1997, 77:1099-1118.
66. Bucci MP, Kapoula Z, Eggert T, Garraud L: Deficiency of adaptive
control of the binocular coordination of saccades in strabismus.
Vision Res 1997, 37:2767-2777.
67.
Rosenfield M: Tonic vergence and vergence adaptation. Optom
Vis Sci 1997, 74:303-328.
74. Crawford JD, Guitton D: Primate head-free saccade generator
•• implements a desired (post-VOR) eye position command by
anticipating intended head motion. J Neurophysiol 1997,
78:2811-2816.
Demonstrates that the relative contribution of eye- and head-movement
components of a gaze shift can be altered by appropriate training.
Monkeys were trained to make head-free gaze shifts to visual targets
while wearing opaque spectacles with a small aperture. After several
weeks of training, monkeys had learned a new motor ‘set’ with a reduced
eye movement contribution to gaze shifts, and they could rapidly switch
between saccades using the normal and learned motor set.
68. Kahlon M, Lisberger SG: Neuronal correlate of pursuit learning in
the floccular complex of the cerebellum. Soc Neurosci Abstr 1997,
23:1298.
75. Telford L, Seidman SH, Paige GD: Canal-otolith interactions in the
squirrel monkey vestibulo-ocular reflex and the influence of
fixation distance. Exp Brain Res 1998, 118:115-125.
69. Miles FA, Eighmy BB: Long-term adaptive changes in primate
vestibuloocular reflex. I. Behavioral observations. J Neurophysiol
1980, 43:1406-1425.
76. Schweighofer N, Arbib MA, Dominey PF: A model of the cerebellum
in adaptive control of saccadic gain. I. The model and its
biological substrate. Biol Cybern 1996, 75:19-28.
70. Cohen H, Cohen B, Raphan T, Waespe W: Habituation and adaptation
of the vestibuloocular reflex: a model of differential control by the
vestibulocerebellum. Exp Brain Res 1992, 90:526-538.
77.
•
71. Dieringer N: ‘Vestibular compensation’: neural plasticity and its
relations to functional recovery after labyrinthine lesions in frogs
and other vertebrates. Prog Neurobiol 1995, 46:97-129.
72. Curthoys IS, Halmagyi GM: Vestibular compensation; a review of
the oculomotor, neural, and clinical consequences of unilateral
vestibular loss. J Vestib Res 1995, 5:67-107.
73. Broussard DM, Hong JA, Bhatia JK, Butt AR: Effects of motor learning
on the response of the vestibulo-ocular reflex to current pulses and
high-frequency rotation in cats. Soc Neurosci Abstr 1996, 22:1094.
Kettner RE, Mahamud S, Leung HC, Sitkoff N, Houk JC,
Peterson BW, Barto AG: Prediction of complex two-dimensional
trajectories by a cerebellar model of smooth pursuit eye
movement. J Neurophysiol 1997, 77:2115-2130.
Simulates predictive pursuit of complex, two-dimensional target trajectories and eye-movement responses to unexpected perturbations in target
trajectory using a model based on the anatomy and physiology of the cerebellum.
78. Katoh A, Kitazawa H, Itohara S, Nagao S: Dynamic characteristics
and adaptability of mouse vestibulo-ocular and optokinetic
response eye movements and the role of the flocculo-olivary
system revealed by chemical lesions. Proc Natl Acad Sci USA
1998, 95:7705-7710.