Download NEUROTRANSMITTER SYSTEMS IN THE VISUAL CORTEX OF

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

Sensory cue wikipedia , lookup

Apical dendrite wikipedia , lookup

Nonsynaptic plasticity wikipedia , lookup

NMDA receptor wikipedia , lookup

Cognitive neuroscience of music wikipedia , lookup

Connectome wikipedia , lookup

Biology of depression wikipedia , lookup

Optogenetics wikipedia , lookup

Stimulus (physiology) wikipedia , lookup

Visual search wikipedia , lookup

Metastability in the brain wikipedia , lookup

Endocannabinoid system wikipedia , lookup

Neurotransmitter wikipedia , lookup

Visual selective attention in dementia wikipedia , lookup

Binding problem wikipedia , lookup

Premovement neuronal activity wikipedia , lookup

Human brain wikipedia , lookup

Eyeblink conditioning wikipedia , lookup

Visual extinction wikipedia , lookup

Time perception wikipedia , lookup

Neuroeconomics wikipedia , lookup

Visual servoing wikipedia , lookup

Visual memory wikipedia , lookup

Environmental enrichment wikipedia , lookup

Spike-and-wave wikipedia , lookup

Synaptic gating wikipedia , lookup

Cortical cooling wikipedia , lookup

Aging brain wikipedia , lookup

Neuroplasticity wikipedia , lookup

Molecular neuroscience wikipedia , lookup

Activity-dependent plasticity wikipedia , lookup

Neuroesthetics wikipedia , lookup

Neural correlates of consciousness wikipedia , lookup

C1 and P1 (neuroscience) wikipedia , lookup

Clinical neurochemistry wikipedia , lookup

P200 wikipedia , lookup

Neuropsychopharmacology wikipedia , lookup

Feature detection (nervous system) wikipedia , lookup

Inferior temporal gyrus wikipedia , lookup

Cerebral cortex wikipedia , lookup

Transcript
ACTA NEUROBIOL.
EIXP. 1988,
48: 335-370
NEUROTRANSMITTER SYSTEMS IN THE VISUAL CORTEX
OF THE CAT: POSSIBLE INVOLVEMENT
IN PLASTIC PHENOMENA
Jolanta SKANGIEL-KRAMSKA
Department of Neurophysiology, Neacki Institute of Experimental Biology
Pasteura 3, 02-093 Warsaw, Poland
Key words: visual cortex, neurotransmitter systems, neuronal plasticity, cat
Abstract. The aim of this paper is to review some of the investigations on neurotransmitter systems suggesting their possible role in visual information processing and their putative involvement in the plastic phenomena observed in the primary visual cortex of the cat. The
neurotransmitters discussed include excitatory amino acids, y-aminobutyric acid, acetylcholine, noradrenaline and serotonin. The following
problems are discussed: (i) the occurrence and localization of the various
neurotransmitter system components, (ii) the developmental changes of
the components of a given neurotransmitter system, particularly in the
critical period, (iii) the effects of manipulating the visual input on neurotransmitter system markers. It seems that especially during the critical period there exists a peculiar pattern of interactions between numerous neurotransmitters and neuromodulators. This may create unique conditions, which enable the visual cortical neurons to change their
properties as a results of alterations of the visual input.
INTRODUCTION
The understanding of the molecular mechanisms underlying the plastic changes in the brain is still fragmentary and does not permit a consistent hypothesis. However, the increasing number of reports that neu-
.
rons may profoundly change neurotransmitter expression and metabolism during development and maturity led Black et al. (15) to suggest
that "neurotransmitter mutability may constitute a unique mechanism
underlying plasticity in the nervous system". One of the manifestations
s f the plastic properties of the brain is the possibility to provoke major
alterations in the functional connectivity of the primary visual cortex
a s a result of abnormal vision. Especially, the binocular and monocular
Normal
6nocularly
deprived
Monocularly
deprived
1
1 2 3 4 5 6 7
Ocular dominance group
VU
Fig. 1. The acular dominance histograms showing the effect of different rearing
condition. Each histogram illustrates the number d visual cortical neurons seen
i n each of seven ocular dominance groups, where 1 represents cells exclusively
driven by the contralateral eye; 7, represents cells exclusively driven by the i w lateral eye; 4, represents cells equally driven by both eyes. Group 3 and 5 represent cells slightly more strongly dominated by the contralateral and ipsilateral eyes
respectively; group 2 and 6 represent neurons very strongly dominated by the contralateral and ipsilateral eyes respectively. W, the number of neurons visually
unresponsive. (Adapted from Ref. 16).
deprivation methods give a powerful tool for manipulating the sensory
input, and offer experimental paradigms for investigating the plasticity
of the visual cortex. The physiological response to these two methods of
deprivation is different. Binocular deprivation (BD) results in the loss or
severe attenuation of responsiveness of the neurons in the visual cortex,
whereas after monocular deprivation (MD) the cortical cells change their
preferred. ocularity to the nondeprived eye (Fig. 1). The modification
of binocularity and also other receptive field properties of cortical neurons is confined to a period in the early postnatal life, the so-called
"critical period" (58, 164). The critical period in kittens has been found
to span from the third week to the fourth month after birth.
The mechanisms underlying the two types of deprivation are different. 'MD during the critical period causes a redistribution of the thalamocortical afferents from the two eyes. In normal adult cats the afferemts
. from each eye are segregated and form alternating domains, termed
ocular dominance (O.D.) columns (122, 123). MD alters this pattern and
the O.D. columns representing the functioning eye extend their boundaries and widen at the expense of the adjacent columns, representing
the deprived eye. This is a manifestation of the competition of the afferents from the two eyes for synaptic space. The amatomical substrate for
O.D. columns has been demonstrated in cats (72, 81 123) and Old World
monkeys (50, 82) by using different neuroanatomical techniques (Fig. 2).
Fig. 2. Scheme of the organizat~ionof the primate striate cortex. ODC, ocular dominance columns revealed by unit recordings and neuroanatomical autoradiographic techniques. Shaded regions correspond to cytochrome oxidase staining pattern. I-VI, cortical layers.
Moreover, a surface antigen that identifies O.D. columns has been described (56). BD from birth postpones the onset of the critical period for
O.D. plasticity (30, 90). Therefore BD seems to prevent some maturation
processes, and the binocularly deprived visual cortex can be regarded as
an underdeveloped structure. Thus it can be concluded that visual experience during the critical period is a necessary factor in establishing
fully functioning and stabilized connections within the primary visual
cortex (91).
Both BD and MD produce changes in the appearance of the nerve
terminals. The cortical cells in BD cats show both a low density bf synaptic vesicles (43) and a reduction of the number of synapses associated with each neuron (27), which may be understood as a result of the
arrest of neuronal maturation due to the lack of some trophic i'nfluence
of the visual afferents on immature neurons (76, 161). In MD cats the
synaptic terminals of deprived afferents are also abnormal morphologically and fewer in number (157).
The studies of the neurochemical correlates of visual deprivation
provide a profitable basis for the search for unified principles, which
may help to explain visual cortical plasticity. The present article includes a review of some of the investigations on neurotransmitter systems,
suggesting their possible role in the visual information processing and
their probable involvement in the plastic phenomena observed in the
primary visual cortex. The neurotransmitters discussed include classical
mediators: excitatory amino aciddglutamic (Glu) and aspartic acids
(Asp)/, y-aminobutyric acid (GABA), acetylcholine (ACh), noradrenaline
(NA), serotonin (5-HT). The evidence concerning the involvement of some
processes relating to neurotransmission, such as protein phosphorylation,
is also considered.
The following problems are discussed (i) the occurrence and localiza- .
tion of the various neurotransmitter system components (such as the
neurotransmitter itself, the neurotransmitter metabolising enzymes and
the neurotransmitter receptors) in the primary visual cortex of a normal adult cat, (ii) the developmental changes of the components of' a gi:
ven neurotransmitter system, particularly in the critical period, (iii) the
effects of manipulating the visual input on neurotransmitter system
markers.
GLUTAMATE AND/OR ASPARTATE
There is considerable experimental support for the view that glutamate (Glu) and/or aspartate (Asp) are transmitters in the cerebral
cortex. Endogenous Glu is released from the rat cerebral cortex following subcortical electrical stimulation (61) or surface administration of
K+ (22). Using the method of transmitter-specific retrograde labeling
of neurons, Baughman and Gilbert (6) found that, in cats, neurons in
layer VI of area 17 projecting to the lateral geniculate nucleus (LGN)
use Asp and/or Glu as their transmitters. The results of biochemical studies of the effect of visual cortex ablation particularly on high affini-
ty uptake of [3H]Glu or D-[3H}Asp, also provide evidence for considering GluIAsp as neurotransmitters in the efferents to the LGN, the
pulvinar and the superior colliculus in rats (40, 77, 78) and cats (41).
Moreover, it appears also that GluIAsp are synaptic transmitters of the
cortico-cortical visual pathways in the cat (6, 41, 55). Despite some
efforts, earlier reports did not find evidence of the neurotransmitter
action of Glu/Asp in the geniculocortical afferents (6, 54).
The excitatory amino acids act through several receptors. Four classes of receptors have so far been distinguished. Three are defined by
the agonists that activate them: the NMDA, kainate and quisqualate receptors. The fourth class is defined by the antagonist action of L-2-aminophosphonobutyric acid (L-AH). The introduction of potent antagonists of different GluIAsp receptors permits us to reinvestigate the pro-
Fig. 3. Simplified diagram of probable GluIAsp geniculocortical projection to area
17 of the cat. LGN, lateral geniculate nucleus; I-IV, cortical layers (Ref. 160 modif ied).
blem of operaking sites of Glu/Asp in the cortical neuronal circuitry.
Recent studies by Tsumoto et al. (160) show that some antagonists of
Glu/Asp receptors effectively bloch visually induced and Glu/Asp induced excitation of the cortical cells of the cat. Among the drugs
examined kynurenic acid (KYNA) was the most potent. This endogenously occurring L-tryptophan metabolite blocked almost completely the excitatory actions of NMDA and kainate and to a lesser extent
the excitation induced by quisqualate. The effectiveness of KYNA in
suppressing visual response was related to the types of receptive fields
of the cells and their laminar location. A great majority of simple cells
are completely blocked by this antagonist (76O/o) whereas only a small
percentage of complex cells react to KYNA with a complete block of
their visual responses. The highest proportion of cells completely blocked by KYNA was found in layer IV ab, IV c and the upper part of
layer VI whereas most of the cells in the other layers were incompletely suppressed or not suppressed at all. Since the primary geniculocortical afferents project mainly to these layers (80), the above results
suggest that visual responses of cortical cells are mediated at least partially by GluIAsp in the cat's visual cortex (Fig. 3). Thus it is reasonable to suppose that alteration in the normal visual input should affect
first of all these excitatory amino acid systems.
There is accumulating evidence of the role of excitatory amino acid
receptors in the plastic phenomena in the brain and special attentiori
is focused on NMDA receptors channel which is permeable to Ca++
ions only if the postsynaptic membrane is sufficiently depolarized. Thus
the NMDA receptor permits current flow only if there is coincident
pre- and postsynaptic activation (26). Therefore it is not surprising
that the involvement of NMDA receptors in visual cortical plasticity
has been considered.
Electrophysiological studies show that the proportion of cells supressed by iontophoretic application of D-2-amino-5-phosphonovalerate
(APV) - a potent selective antagonist of NMDA receptors - is much
higher in the visual cortex of young kittens than in adult cats. These
results suggest that NMDA receptors in the visual cortex during the
critical p e r i d are more effective than those of mature cats (159). Furthermore, developmental studies indicate that the APV-sensitive [3H]Glu
binding sites are especially numerous during the critical period (17). It
can be conjectured that the enhanced effectiveness of NMDA receptors
on first order visual cortical neurons is a necessary requirement for plasticity. Interestingly, the BD of cats does not affect the ontogeny of
NMDA receptor sites in the visual cortex (17). More information concerning the participation of NMDA receptors in visual cortical plasticity
derives from the study of MD animals. In the visual cortex of the kitten the O.D. shift normally occurring in response to monocular vision
can be prevented by chronic administration of APV. The effect of APV
seems to be dose-dependent (69, 70, 141). Similarly, ketamine-xylazine
anaesthesia prevents the O.D. shift in kittens which receive monocular
exposure (109). Probably this effect is due to the blockade of NMDA
receptors by ketamine (156). Evidence concerning the role of GluIAsp
transmission involving NMDA receptors comes also from the model experiments on the amphibian optic system. Antagonist-AVP appears to
desegregate reversibly the O.D. columns when applied to the tecta of
tadpoles with a grafted supranumerary eye, while chronic application
of agonist-NMDA appears to potentiate the formation of O.D. stripes
(23). To sum up, these results seems to indicate that NMDA receptors
play a crucial role in O.D. plasticity expression. They also point to the
participation of Ca+t ions in the plastic phenomena since NMDA receptor activation raises Ca+t conductance into postsynaptic cells. The
resulting Ca+ fluxes may trigger off synaptic modifications. Apart from
the investigations concerning the role of NMDA receptors in visual cortical plasticity, there have been studies, conducted by Shaw and Cynader (125), of the possible involvement of glutamate in the O.D. changes.
They found that intracortical chronic infusion of L-Glu during the period of monocular vision in young kittens largely prevents the O.D.
shift which normally takes place under these conditions. Glutamate probably disrupts normal cortical activity, either by direct excitatory action
on cortical neurons or as a consequence of neurotoxic effects, and thus
blocks plasticity. Therefore, normal postsynaptic activity seems to be an
important factor of cortical plasticity.
GABA
There is evidence from morphological, biochemical and electrophysiological stuhes that GABA is a transmitter in the visual cortex. By
using immunocytochemical methods and high affinity uptake of exogenous [3H]GABA it has been possible to provide the morphological
characteristics of stained neurons and information about their connections. It appears that GABAergic interneurons and terminals are present in all layers of the striate cortex. They form a heterogeneous population with regard to size and synaptic input. Besides a powerful
asymmetric, presumably excitatory synaptic input along their dendrites
some GABAergic neurons receive synaptic input from other GABAergic
neurons not only on their cell bodies but also on their dendrites. Therefore GABAergic neurons are likely to be involved in many different
ways in visual information processing (150).
Light and electron miroscopic examination of the organization of
[3H]GABA transporting neurons in the monkey's cerebral cortex (Macaca) shows that GABAergic neurons form a bidirectional system of
connections which join together cells in the superficial and deep layers
of the cortical columns (probably representing functional cortical units)
(35). A vertical organization of GABAergic neurons has also been reported in the visual cortex of the rhesus monkey (149). GABAergic
cells seem to constitute 8-20°/a of all the neurons present in layer IV
of the cat visual cortex (42, 148).
Glutamic acid decarboxylase (GAD), the enzyme synthezising GABA,
is another marker used to localize GABAergic neurons and termi-
nals, because it is confined almost entirely to neurons. Immunocytochemica1 staining with an antiserum to GAD appears to be rather uniformly distributed throughout the cortical layers of the rat (112). In the
monkey visual cortex, however, distinct laminar variations in GAD immunoreactivity are present, with the highest densities of GAD positive terminals in and around layer IV (51). Moreover, in the Macaca, within
laminae I1 and I11 of area 17 some periodic variations in the density
of GAD containing. cell bodies and terminals can be observed. This pattern of periodic dots (puffs) is identical with that shown for cytochrom
oxidase and for 2-deoxyglucose (2DG) labeling. This suggests that the
rows of GAD positive neurons in supragranular layers are preferentially related to each eye (50). The distribution of GABAergic neurons
seems to correlate with the "puff" regions of elevated cytochrom oxidase activity also in the striate cortex of the squirrel monkey (19). In the
kitten striate cortex GAD immunostaining associated with cells was found
in all layers and was uniformly distributed in layers I1 to IV. In contrast, axon terminals positive for GAD immunostaining' showed laminar
variations and formed a distinct band in layer IV. However, no evidence
of dots in supragranular layers was reported (11).
Important information about the possible site of GABA action derived from the examination of the distribution of GABA receptors. Our
own studies (144) as well as the data of other authors (94, 128) show
a distinct laminar pattern of [3H]muscimol - GABA agonist - binding
in the visual cortex of young kittens and adult cats. The highest concentration of [3H]muscimol binding occurs in cortical layer IV. No evidence was found for periodicity in the pattern of distribution of binding sites within the laminar organization (92, 144), as was described
by Hendrickson et al. (51) for GAD immunoreactivity. The GABA level
itself varied 'gmong the cortical layers in the occipital cortex of, thq
rat's brain, the highest being detected in layer IV (59).
The pattern of distribution of GABAergic markers showing the
highest concentration in layer IV suggests a strong influence of GABA
on the terminals of thalamocortical afferents in the striate cortex and its
passible role in shaping the properties of visual cortical neurons.
The earliest evidence that GABA is the major inhibitory transmitter in the visual cortex from the results of Iversen, et al. (60), showing
that GABA is released during inhibition in the cat's visual cortex in
a calcium dependent way. Subsequently, Sillito and his coworkers demonstrated that GABA has a strong inhibitory effect on visual cortical
neurons, which can be blocked by the GABA antagonist, bicuculline
(133, 134). Moreover, GABA-mediated inhibitory mechanisms contribute to the visual response properties of neurons in the cat's visual cortex
because the iontophoretic application of bicuculline reduces directional
and orientation selectivities and produces other modifications of the receptive field properies of cells (135). It has also been found that GABAmediated inhibition plays some role in determining O.D. (140). Electrophysiological studies revealed powerful GABAergic mechanisms operating in the visual cortex, which determine the receptive field properties and may dictate the actual responsiveness of cells to the visual input. These data together with the results concerning the presence and
distribution of GABAergic markers in the striate cortex suggest a possible involvement of GABAergic transmission in the plastic changes of
the properties of cortical neurons observed after manipulation of visual
input during the critical period.
Some inchcations concerning the possible participation of the
GABAergic mechanism in cortical plasticity has been provided by ontogenetic studies. Despite the fact that GABAergic inhibition operates
already in a group of neurons in the visual cortex of very young kittens (118, 167), intracortical GABAergic inhibition has not been fully
established in the visual cortex until the end of the critical period for
cortical modifiability. It appears that GAD activity develops relatively
late and in the rat visual cortex remains very low at the time of eye
opening (78, 85). The absolute and relative numbers of inhibitory synapses using GABA as a transmitter are much lower in the postnatal than
in the mature cortex (166, 168). In the visual cortex of the cat the peak
of [3H]muscimol binding density occurred at 3 months postnatally. In
3 days old kittens binding density was about 40°/o of the maximal value
wkereas in adult cats it represented about 70°/c of the peak of binhng
density (128). Moreover, the number of GABA binding sites is higher
in cats after a long period of dark rearing than in normal cats (126).
One of the most fully investigated phenomena of neuronal plasticity is the change of O.D. pattern following monocular deprivation. The
mechanism underlying the effect of MD is still not clear. It has been
suggested that MD results in the reduction of the efficiency of excitatory input in deprived pathways (139). Another explanation focusses on
selective suppression on the deprived eye input by the GABA mediated
inhibition exerted by connections from the open eye. The possibility of
the involvement of inhibitory mechanisms in the O.D. shift derives
from the following observations: (i) in MD kittens enucleation of the
open eye restored responsiveness to the deprived eye in many cortical
neurons. This effect was age related and, being most pronounced in 4-5
weeks old kittens, it still persisted in adult cats. Thus it appeared that
the responses of the deprived eye were somehow being suppressed by
the tonic activity of the normal eye (74); (ii) iontophoretically applied
bicuculline restored binocularity in MD kittens (37). Therefore, when the
inhibitory inputs are blocked by bicuculline, excitatory inputs are unma- .
6
- Acta
Neurobiol Exp. 8/88
sked and the responses of neurons to the deprived eye stimulation are
restored. In this case GABAergic inhibition seems to play an active role in the effects of MD. However, the effects of bicuculline could be given an explanation which speaks in favour of the mechanism involving
the r.eduction of excitatory input in the deprived pathw'ay. If MD reduces the efficiency of the excitatory input with the inhibitory inputs
unaffected, bicuculline makes it possible to reveal the weak cortical responsiveness to deprived pathway stimulation. Therefore in this case
GABAergic inhibition seems to play only a passive role in the effect of
MD (11).
If the shift of O.D. towards the open eye is associated with an active
role of GABAergic inhibition, one might expect to see differences in the
level of the GABAergic markers and/or their distribution between normal and MD animals. Some indications concerning the alterations of
the GABA system in the visual cortex as a result of abnormal visual
experience were found in our studies of the effects of short monocular
vision on GABA receptor binding activity in 5 weeks old kittens (145).
Using [3H]muscimol as a ligand, we were able to show that 3 days of monocular vision resulted in an increase of GABA receptor binding in the
striate cortex of both normally reared and binocularly deprived animals
(by nearly 50°/a). BD since birth did not affect the GABA receptor binding.
It should be stressed that neither 3 days of binocular vision in deprived
kittens nor 3 days of binocular deprivation in normally reared ones produced significant changes in GABA receptor binding (145). The enhancement of [3H]muscimol binding observed after 3 days of monocular vision seems to be related to the increase of receptor affinity rather than
to the change of the total number of receptor sites (146). Comparison of
autoradiographic images generated from normal and MD kittens showed
an increase of the [3H]muscimol label in the visual cortex area 17 in
MD kittens confirming the biochemical results. However, no alterations
in the laminar distribution of labeling was detected. We suggest that
changes in the activation of the GABAergic system are specifically due
to the asymmetric visual input and are probably connected with the mechanisms of O.D. shift. Surprisingly, Mower et al. (92) reported that
long MD, lasting 8-12 months, did not affect the total number, affinity
and regional distribution of the GABA receptor in the visual cortex a s
revealed by the [3H]muscimol binding. Similarly, unilateral eyelid closure in the rat from the 11th to the 25th day did not affect the developmental pattern of GABA a receptor in the visual cortex (121). On the
other hand, recent results of Shaw and Cynader (126) demonstrate that
hlD from the time of eye opening results in an increase of the GABA
receptor number. This tends to support our suggesticxn about the role of
GABA mediation in the mechanism of O.D. shift.
From the results showing the increase of the [3H]muscimol binding
activity after MD as well as from the results of bicuculline administration it seems reasonable to conclude that the GABA influence on the
O.D. shift involved GABA, receptor sites. However, other GABA receptors can also be affected by MD. For instance GABA, receptor
sites, bicuculline insensitive, are present in the striate cortex of the cat,
because the iontophoretic application of baclofen inhibited the spontanous and visually evoked responses of neurons independently of the bicuculline action (7). Appart from the alteration of the receptor sites other
changes in the GABA system as a result of MD are also possible. These
include changes in the GABA containing neurons and terminals. In fact,
Hendry and Jones (52) found that eye removal or MD reduce the number of immunocytochemically stained GABA and GAD somata and terminals in deprived-eye dominance columns in area 17 of the monkey.
They interpreted this reduction a s being due to a decline in the GABA
and GAD concentrations within the individual neurons and terminals
to the point where the substances are no longer detectable by immunocytochemistry, since no reduction in the total neuronal density in area
17 was detected. The decline of GABA levels in cortical intrinsic neurons and axon terminals should produce changes in the physiological
activity of their target cells. These changes could account for the functional expansion of the intact-eye dominated columns. These data seem
to show a regulation of cortical transmitter levels by sensory experience. Interestingly, this effect was observed in adult monkeys 2 weeks after eye removal or 11 weeks after suturing the lid of one eye. In contrast, Bear and his coworkers (ll), using GAD immunocytochemistry as
well as biochemical GAD assay, were unable to find changes in the distribution and activity of this enzyme in the striate costex in the monocularly enucleated or MD kittens. Both enucleation and MD have no
consistent effects on either the regional or the laminar distribution of
GAD in the striate cortex. The band of layer IV puncta remained uniform even though the periods of MD examined had been sufficient to
cause a physiological O.D. shift in the striate cortex. These results led
the authors to the conclusions that the numerical density of GABAergic
synapses in the visual cortex is not regulated directly by thalamic activity and that changes in the density of GABAergic synapses do not
account for the O.D. shift observed in the kitten striate cortex after MD.
The data discussed imply that the results of experiments concerning
the involvemznt of GABAergic system in visual cortical plasticity are
not univocal. The postulated key role of GABAergic mechanisms in
adaptive changes (113) and in input selection and restriction (38) in the
nervous system imply, however, the involvement of GABA meaated
processes in visual cortical plasticity.
ACETYLCHOLINE
Neuroanatomical studies show that in the striate cortex of the cat all
cholinergic innervation arises entirely from extrinsic sources (8). Similarly, no evidence exists for the involvement of cholinergic transmission
in the primary optic system of the rat (14, 163). Thus it seems that basal .forebrain nuclei are the sole source of cholinergic projection to the
striate cortex and probably also to the subcortical visual centres. The
studies in question employed acetylcholinesterase (AChE) histochemistry
to investigate cholinergic innervation. Recently, using monoclonal antibodies directed against choline acetyltransferase (ChAT), Stichel et al.
(155) have strenghened the earlier results showing that in the visual
cortex (area 17) of the cat, despite dense cholinergic innervation only
a few ChAT positive neurons can be detected. ChAT positive fibers and
varicosities have been found t a be present in all layers of the Visual
cortex. The distribution pattern of cholinergic fibers is compatible with
the notion that ACh modulates cortical activity (152, 153). Although the
direct cholinergic innervation of visual centers is lacking, iontophoretically applied ACh strongly affects the activity of visual neurons. In the
striate cortex the responses of most neurons are modified byl ACh. ACh
exerts a dual action on the specific responses of cells in area 17 of the
cat. The responses of most cells are enhanced by ACh, Cells facilitated
by ACh have been found in all cortical laminae, while those inhibited by
ACh were found in laminae I11 and IV (138). Recent electrophysiological
investigations also suggest a dual action of ACh in cat LGN : ACh blocked intrageniculate inhibition and exerted direct facilitation of relay
cells (36). These observations speak in favour of the view that ACh acts
as a modulator in visual information processing.
Biochemical investigations of late 60ties and early 70ties indicated
that keeping the animals in a modified visual environment can affect
the level of AChE, a degradative enzyme for ACh (13, 21, 84). These
observations may suggest the involvement of ACh transmission in modulating visual functions. Our biochemical results concerning the changes of AChE as well as ChAT activities in the visual areas of the cat's
brain during development suggest the participation of cholinergic mechanisms in visual information processing (106). We have found that the
ChAT activity rises from the cat's birth to reach the maximal level by
the 3rd month of its life and then drops to the adult level. In contrast,
the developmental profile of AChE is quite different. The enzyme activity, reaching 90°/o of the adult value at 1 week of age, drops at the onset of the critical period and remains low until the third postnatal
month. These results can be interpreted as an expression of enhanced
availability of ACh during the critical period. Furthermore, BD from
the 8th day of life, (the time of eye opening) apparently increases the
activity of AChE in the cat's visual cortex and in LGN, but t h s ChAT
activity remains unchanged in comparison with that observed in cats
with normal binocular vision. Therefore it is plausible to maintain that
the level of ACh is reduced in BD animals.
Histochemical studtes have shown the pattern of developmental
changes in AChE positive fibres and AChE positive cell bodies in the
visual cortex of the cat. Namely, Bear et al. (9) observed that the development of AChE positive axons in the striate cortex is not fully
complete until the age of at least 3 months, and is characterized by
a number of distinct events. The most interesting phenomenon seems to
be the appearance and position of AChE positive cell bodies. Stained
cells first appear in the white matter subjacent to layer VI shortly after
birth. After 2 weeks most cells in layer VI are also AChE positive. The
staining of these cells gradually disappears over the next 2 months until, at the age of 3 months, there are no AChE positive cells in the cat's
visual cortex. Some stained neurons are detected in layer V by the age
of 1 year and persist throughout adulthood. Lately, reports showed that
there were no indications of the striate cortex of the receiving an especially dense cholinergic input during the critical period (154). Comparison of our biochemical data with those found by using histochemistry
clearly shows discrepancies between the results obtained. One of the
possible explanations proposed by Stichel and Singer (154) is that biochemical investigations consider all forms of AChE whereas histochemical methods exclude the soluble forms of the enzyme. Laminar distribution of AChE activity shows some unexpected features. In adult mammals layer IV is poor in AChE activity (34). However, at least in the
visual cortex of the rat, a transient AChE activity in layer IV appears
during the period in which geniculocortical axon terminals are establishing functicmal connections with the postsynaptic sites in the cortex
(115, 116). These data suggest that AChE may play a role in the development of neuronal connections, as has been suggested by Robertson (114).
Thus the changes in AChE activity registered during development can
be related to processes not exclusively connected with cholinergic mechanisms (48).
Unilateral enucleation performed on the monkey c5anges the pattern
of distribution of AChE activity in the striate cortex Stripes of high
AChE activity alternate regularly with zones of low AChE activity In
layer IVc. Stripes of dark staining for AChE coincided with stripes of
low cytochrom oxidase activity, clearly showing a relation to the O.D..
The reason of this response is unknown but this study shows that AChE
activity in the visual cortex is dependent on the activity in the thalamo-cortical pathways (47, 50).
i
The function of ACh in the cerebral cortex appears to be largely mediated by muscarinic receptors (MChRs) (75). Biochemical and autoradiographic investigations have shown the presence of muscarinic binding
sites in the visual cortex (24, 25, 117, 129). Our autoradiographic studies (144) have demonstrated the laminar variations in [3H]quinuclidinyl
benzilate ([3H]QNB) binding in visual cortical areas in young kittens
(5 weeks old). The weakest binding occurred in layer IV. Supragranular
layers and layer V were strongly labelled. A similar pattern of distribution of muscarinic binding sites was found also in adult cats by using
. [ ~ H ] Q N Bor N-[3H]methylscopolamine ([SHINMS) as a ligand (29).
Among muscarinic binding sites both subclass M1 and subclass M2 are
present. A comparison of the inhibition of the [3H]NMS binding by
100 pM carbachol and 300 nM pirenzepine showed that subcortical visual structures, e.g. LGN and superior colliculus, contain predominantly
M2 sites. However, M1 sites constitute the main population of MChRs
in the primary and accessory visual cortical areas but layer IV seems
to be enriched in Mp sites. The predominance of Mz sites in the subcortical visual structures and the relatively higher proportion of these sites
in cortical layer IV are compatible with the view that mainly M2 sites
are involved in the modulation of synaptic activity at the primary stage
of the visual information processing (29). Detailed ontogenetic studies of
Shaw et al. (129) revealed striking changes in the distribution of
MChRs in the striate cortex. In 3 days old kittens they found the highest [3B]QNB binding in layer IV. During development the pattern of
binding reversed and by 3 months layer IV was the least densely labelled. Moreover, the biochemical data show that, especially during the critical period, changes in number and affinity of MChRs take place. The
highest receptor sensitivity (expressed as the ratio Rs = Bmax: 2Kd)
was observed after 6 weeks postnatally, then it decreased and achieved
the adult levels in 3 months old kittens. The changes both in number
and affinity may suggest a mechanism of controlling or enhancing the
sensitivity of the postsynaptic response during development. Furthermore, the observation that the peak of receptor sensitivity occurs during the, critical period may suggest a possible functional role for this
phenomenon in the plasticity process (130). The results of earlier stu-
.
dies performed on rats also suggest the involvement of MChRs in modulating visual functions. An elevation of [3H]QNB binding level was
found in dark reared rats within 3 hours of the onset of light exposure:
however, the response was transient and by 24 h had disappeared (117).
MD and dark rearing resulted in a reduction of [3H]QNB binding in the
visual cortex of both hemispheres but an elevation in both superior colliculi in 25 days old rats. However, these effects disappeared completely in adult rats (119). Since the first visual stimulation and abnormal
visual environment affect MChRs, it is reasonable to conclude that cholinergic muscarinic action may influence visual functions.
Apart from muscarinic action, the possibility of cholinergic nicotinic
transmission in modulating neuronal activity in the cat's visual areas
should be taken carefully into consideration. Recently, the presence of
[3H]nicotinic binding sites has been reported (100). These sites were
located preferentially in layer IV of area 17. The conclusive results of
Prusky et al. (107, 108) clearly show that nicotine receptors are located
presynaptically on LGN terminals in the primary visual cortex of the
cat. Thus nicotinic receptors-may play a specific function in regulating
the activity of the primary visual thalamic input. To sum up, both muscarinic and nicotinic cholinergic receptors are present in the visual
system. The different distribution of these receptor sites suggests that
the action of ACh may depend on the precise cortical location at which
the transmitter is released. The striking developmental characteristics
of both the muscarinic and the nicotinic ACh receptors associated with
the critical period for visual cortex plasticity suggest an important role
for cholinergic systems in the alterations of the cortical function by the
changed visual input (107). New information concerning the involvement
of ACh in modulating visual cortical plasticity emerges from the results
of Bear and Singer (12). They show that the combined destruction of the
cortical noradrenergic and cholinergic innervations reduces the physiological response to MD although lesions of either system alone are ineffective. These data point to the possibility that ACh and NA facilitate
synaptic modifications in the visual cortex by a common molecular mechanism (12, 140). It has been suggested that the removal of a sufficient
amount of facilitatory extrageniculate input could lower cortical excitability below the threshold of synaptic modification and therefore block
plasticity (12). The authors, however, did not exclude the possibility that
the actions of NA nad ACh can regulate some events more specifically
related to the control of synaptic plasticity, for example second messenger-dependent phosphorylation of proteins which are involved in the
modification of synaptic properties.
The changes in cholinergic markers during development and after
manipulation with visual input, and the combined effect of ACh and
NA depletions on the O.D. plasticity give good reasons to believe that
ACh is involved in the plastic modification of the properties of visual
cortical neurons. Moreover, the new results of Metherate et al. (86),
showing that iontophoretically administered ACh permits enhancement
of neuronal responsiveness in the cat's primary somatosensory cortex,
support the belief in the cholinergic influence on plastic phenomena.
NORADRENALINE
Neurons containing noradrenaline which are located in the locus coei-uleus project widely and monosynaptically upon the cerebral cortex
(71, 89, 158). The major noradrenergic fibres are oriented longitudinally
through the grey matter and branch widely; thus NA can modulate neuronal activity synchronously throughout a vast area of the neocortex.
Noradrenergic innervation of area 17 in primates exibits the most highly
differentiated laminar pattern of all the neocortical areas (88). Layers
V and VI appear to receive dense innervation whereas layer IV receives
only a poor projection. The pattern of NA innervation and the fact that
monoaminergic fibres form also conventional synapses in the cerebral
cortex (98) suggest that NA action is directed at a specific set of visual
cortical neurons, and therefore it seems that the action of NA is much
more specific than has been suggested (158). On the other hand, the fact
that NA fibres are very few in layer IVc (87, 88) seems to indicate that
NA does not play a significant role in the modulation of the activity
of geniculocortical terminals. The early ontogenetic appearance of the
NA innervations and their tangential distribution pattern suggest a neurotrophic role (158).
In 1976 Kasamatsu and Pettigrew (65) reported that 6-hydroxydopamine treatment prevents the O.D. shift produced by MD, and suggested that catecholamines may play an important role in the regulation
of neuronal plasticity in the kitten visual cortex. The results of their
subsequent studies, especially their discovery that microperfusion with
NA resulted in restoration of cortical plasticity (66), seemed to support
this suggestion and led to the hypothesis that NA input to the visual cortex is necessary to maintain plasticity (evidence reviewed by Kasamatsu, 64). But the results of other authors have failed to demonstrate the
relationship between the cortical NA level and the ability of visual neurons to change their functional properties (1, 10, 32, 33). Nevertheless,
there is general agreement that microperfusion of the visual cortex
with 6-hydroxydopamine blocks plasticity during the critical period.
Few data are available concerning the cortical distribution of NA
receptors. In rats P-adrenergic receptors are present in all layers but
predominate in layers I-IV (96, 97), whereas a-receptors are uniformly
distributed throughout the cortex in low concentration (169). In the visual cortex of the cat the density of P-adrenergic receptors has been
found to be highest in the supragranular layers and lowest in layer IV
(3). Also in the ferret visual cortex laminar variations in the distribution of al and PI-adrenoreceptors have been reported (45). The density
of &-receptors has been found to be high over layers 1-11 and very low
over layer IV, and intermediate over deep layers. In contrast, the areceptors are seen to be diffusely distributed, but preferentialy concentrated in layer IV and upper layers. The distribution of these receptor
types may suggest two different ways for NA influence. The al-receptors may be preferentially associated with the enhancement of the excitatory input to layer IV, while the P1-receptors can be linked to the
enhancement of inhibitory response outside layer IV. The high concentration of a, adrenoreceptors in layer IV, which coincides with the distribution of afferent terminals of the LGN in cats (80), and a periodic pattern of a,-receptors seen in cortical layer I11 imply that these
receptors may be involved in the O.D. columns. The distribution of the
13-receptors suggest that these receptor sites act mainly in the supragranular layers and therefore may play a role in regulating secondary intracortical processing.
To test the possible role of (3-adrenergic receptors in determining the
state of developmental plasticity, the ontogeny of binding capacity in
the visual cortex of the cat has been examined. Wilkinson et al. (165)
found that in the visual cortex of the cat the binding of P-adrenergic
ligand [3H] dihydroalprenolol ([SHIDHA) increases quickly from birth
to 4 weeks, then rises slowly and reaches the maximum at 12 weeks.
The adult pattern of the distribution of 0-adrenergic receptor sites is
already formed at the beginning of the critical period (3). These re4
sults show that fl-adrenergic bindng site density increase coincides with
the onset of the physiologically defined critical period and correlates with
developmental profile of NA levels (62). Jonssbn and Kasamatsu, however, registered a different developmental pattern of P-adrenergic receptor sites (62). They observed a clear peak of [3H]DHA binding at the
age of 7-9 weeks; then followed a decrease and at 11 weeks postnatally
the adult level was reached. Such a maturation curve may indicate
a correlation between the (I-adrenergic sites and the duration of the critical period. Lack of the effect of dark rearing and monocular eyelid
suture on the developmental pattern in the visual cortex of kittens (3,
165) and rats (120) inhcates that the receptor sites are probably not
involved directly in determining the state of plasticity that is seen du-
ring the critical period. But using pharmacological approach Shirokawa
and Kasamatsu (132) were able to show the concentration dependent
suppression by fl-adrenergic antagonists of the shift on O.D. following
MD. These results were interpreted as suggesting that there is a positive
correlation between the number of activated fl-adrenergic receptors within the visual cortex and the extent of changes in O.D. following MD
(67, 68, 132). On the other hand, the comparison of the responses of visual cortical neurons during iontophoresis of NA showed no differences in the effects of NA observed between adult cats and kittens (162).
NA involvement in visual'cortical plasticity is still disputed (compare
162) but, as mentioned in the section concerning ACh, it seems that the
presence of both ACh and NA inputs is necessary for the plasticity of
t h e visual cortex (12).
SEROTONIN
Neurons containing serotonin (5-HT) which are located in midbrain
raphe nuclei project widely and monosynaptically upon the cerebral
cortex (71). The anatomical studies on primates have demonstrated that
5-HT axons are differentially distributed within the visual cortex and
show a distinct laminar pattern of innervation (71, 88). Although some
species differences exist, it has been found that layer IV receives very
dense serotonergic projections. Since serotonergic innervation is directed
to definite cortical layers, it seems that the action of this transmitter has
a restricted role in cortical circuitry. Interestingly, in the squirrel monkey (New World monkey) complementary 5-HT and NA laminar innervation was observed, but in the macaque (Old World monkey) 5-HT and
NA axons show a considerable overlap (87, 89). Nonetheless, the 5-HT
projection has a markedly higher density in layer IV in both species as
compared with other layers (71). Moreover, it seems that in the visual
cortex 5-HT levels are higher than those of NA (46).
The relationship between serotonergic terminals and the columiar
organization of the primary visual cortex remains obscure. Hendrickson
(50) reported that 5-HT is preferentially localized in the interdots zones
of the macaque's visual cortex area 17, which correspond t a the regions
of low'cytochrome oxidase activity. This result suggests some link between serotonergic innervation and O.D. columns. Other authors, however, have been unable to detect any periodicity in the tangential distribution of 5-HT fibre density (71).
5-HT was found to inhibit the visually evoked activity of the majority of cortical neurons. This effect is prolonged (110, 111). On the
grounds of these data it has been postulated that serotonin has a modulatory role in the visual cortex.
The biochemical data show a higher concentration of endogenous
5-HT in the visual cortex of the monkey compared with the more anterior cortical areas (18, 111). However, such a pattern was not found
in the cerebral cortex of the cat, where the frontal cortices showed much
higher 5-HT concentration (44). The developmental studies performed
by Jonsson and Kasamatsu (62) show that the 5-HT level in the visual
cortex of kittens is relatively high at birth. It increases dramatically,
peaking at 3-5 weeks with a value similar to the adult level, and then
decreases between 5 and 7 weeks. Between 7 and 13 weeks the serotonin level is about 50-60°/o of the adult value.
The distribution of 5-HT receptors was investigated by using receptor binding autoradiography. Both high affinity (5-HT1) and low affinit y (5-HT,) binding sites are present in the visual cortex (101-104). Some
laminar variations in 5-HT1 receptor densities were found in the visual
cortex of the rat, the internal laminae showing higher binding than the
external ones (102). In yoqng kittens [3H]5-HT labelled intensively the
layer 1/11 and V (143). In the human brain, in Brodman area 17, the
highest 5-HT binding was found in layer IVcP (103). It was found
that there are species differences in the distribution of 5-HT receptors.
The presence of a significant concentration of 5-HT receptor sites in the
visual cortex may suggest the involvement of these receptors in the
control of visual information processing. The developmental pattern
of 5-HT receptor binding sites also shows a distinct peak at the age of
4 weeks (62). At this time the value of the binding activity is more than
three times the adult value. Afterwards the binding decreases gradually. The authors argue that their data imply that the 5-HT system itself
is not directly involved in the regulation of the critical period for cortical plasticity, but it is possible that it induces or triggers off the onset
of sensitivity. The highest level of endogenous 5-HT and the highest
receptor binding may be a manifestation of some processes related to
the completion of neuronal differentiation regulated by 5-HT (49).
Some evidence concerning the involvement of 5-HT in visual functions derives from the study of the effect of visual deprivation and visual stimulation upon the level of serotonergic system markers. Enucleation has no significant effect on the concentration of 5-HT in the
subcortical visual centres of the rat but, it decreases markedly the concentration of 5-HT in the occipital cortex. These results show that the
cortical 5-HT system is more susceptible to enucleation than LGN and
the superior colliculus. Monocular lid suture affects all the visual structure and increases the concentration of 5-HT in the occipital cortex in
both hemispheres (79). No changes in 5-HT receptors were found in
dark reared rats, suggesting that the development of the 5-HT innervation to the cortex is not criticaly dependent on the visual input. But,
if dark reared rats were exposed twlight for 3h, a significant increase
in 5-HT binding was observed in the visual cortex and the motor cortex. This effect was transient and there was a return to normal level
after 7 days. Since the increase of binding was found also in the motor cortex, it seems that the response of 5-HT receptors reflects a more
general effect, probably related to the overall arousal of the animal (93).
Nevertheless, when kittens deprived binocularly of pattern vision until
the 28-th day of their life had their first monocular visual experience,
changes of serotonin levels were observed in the visual cortex. Stimulation for 3h resulted in an increase of the 5-HT level, whereas after
14 h it produced a decrease. The effect was transient, no longer detectable after 75 h of stimulation (73). The observed fluctuations in 5-HT
levels may reflect changes in the state of cortical metabolism. These data seem to indicate that there is no simple relationship between the
first visual experience and the activation of the serotonergic system, but
rather that 5-HT is involved in many comdex, dynamic processes occurring during the first hours of stimulation in the visual cortex. The
involvement appears to be transient, possibly subsiding after the first
plastic changes in connectivity have taken place.
FINAL REMARKS
In this paper an attempt has been made to review our present-day
knowledge of the possible involvement of different neurotransmitter
systems in the plastic phenomena occurring in the visual cortex during
the critical period. It has been shown that numerous transmitter systems
are present in the visual cortex, and are probably affecting the visual
functions. These neurotransmitter systems reveal a specific laminar pattern of innervation and also a clearly defined laminar distribution of
their receptors (Table I). This may reflect the pattern of cortical circuitry connected with the visual information processing. Even dopamine,
not included in the above results and often neglected because of its low
concentration in the striate cortex, shows a very specific distribution1 of
innervation mainly addressed to laminae VI (105).
In investigating the problem of neurotransmitter involvement in visual cortical plasticity two main experimental approaches are followed.
The first one attempts to examine whether experimentally induced
changes in neurotransmitter systems affect the visual cortical plasticity.
The second one attempts to find which properties in the neurotransmitter system components are confined to the critical period in the hope
of answering the question what makes the visual cortex susceptible to
modification of its connectivity.
.
TABLEI
Laminar distribution of difierent neurotransmitter receptor sites in striate cortex of adult cat.
APV, D-2-amino-5-phosphonovalerate; [3H] AMPA, [3H] amino-3-hydroxy-5-methyl-isoxazole-4-propionic acid; [3Hl NMS, N-[3H]methyIscopolamine; [3H] QNB, [3H] quinuclidinyl benzilate;
I3H] 5-HT, [3H] 5-hydroxytryptamine; ['251] ICYP, [lZ5I]iodocyanopindolol; [3H] DHA, [3H
dihydroalprenolol; [3H] CHA, [3H] cyclohexyladenosine; [3H] NECA, [3H] 5-N-ethylcarboxamidoadenosine
-
-
Receptor
Ligand
GABAA
NMDA
Quisqualate
Muscarinic
MI
M,
3H muscimol
3H glutamatef APV
3H AMPA
3H NMS or 3H QNB
pirenzepine
carbachol
Nicotinic
5-HTI
0-adrenergic
3H nicotine
3H 5-HT
l Z 5 I ICYP
3H DHA
3H CHA
3H NECA
adenosine Al
+
+
Laminae with the highest
denisity of receptor sites
IV
I, 11, 111
I, 11, VI
I, 11, I11
In all laminar proportion
of MI sites is higher
except layer IV
IV
I, 11, V
I, 11, 111 (lowest in IV)
I, 11, 111, VI
I, 11, 111 (absent in layer IV)
I, I1
References
94, 143
17
131
29,
129, 143
100, 107
143
3
2, 127
127
--
Using the first of the above experimental approaches, Kasamatsu
and Pettigrew (66) constructed the attractive hypothesis that NA plays
an essential role in visual cortical plasticity, This initiated a series of
investigations designed to verify the above notion. However, their results are contradictory and the hypothesis has been strongly criticized
(33). Recently, Bear and Singer (12) have reported that both NA and
ACh inputs are necessary for plasticity. A critical overview of the results concerning the pharmacology of visual cortical plasticity was presented by Sillito (138), who conjectured that any procedure that severely
disturbs the normal functioning of the cortex may block plasticity. Nevertheless, it seems that these studies provide some valuable information,
especially if they are considered together with the results of the experiments employing the second experimental approach. This approach is
directed towards investigating the developmental pattern of particular
neurotransmitter system components in normally reared and visually
deprived animals, special attention being paid to the critical period.
Although the results of these experiments do not prove causality between the registered developmental changes and the plastic properties
of the cortex, they strongly suggest that neurotransmitter receptors may
be the sites of modifications through which the alterations of synaptic
TABLE11
Neurotransmitter receptors and visual cortical plasticity in the cat
Receptor
NMDA
B-adrenergic
Muscarinic
5-HTI
A, adenosine
~odificationof
properties during
critical period
Effect of manipulation with
visual input
Effect of application of
receptors blockkers
ketamine preVents O.D. shift
after M D (109)
APV prevents
O.D. shift after
M D (69,70,141)
bicuculline resPeak of sensitivity 3 days of asymat 9 weeks (128,
metrical visual in- tores binocula130)
put rises receptor rity in M D
binding level ducats (37)
ring the critical
period (145)
MD and D R rises
receptor densities
(126)
M D has no effect
on receptor densities (92)
Peak of binding
D R has no effect propranolol and
at 7-9 weeks (62) on developmental sotalol supress
profile (165) or on O.D. shift after
Progressive increa- laminar distribuM D (67,68,132)
se of binding up
tion (3, 165)
to 12 weeks; then
the level remains
constant (1 65)
Peak of sensitivity Transient increase
in binding after
at 6 weeks (129,
130)
D R ; at 30 days
D R has no effect
on distribution
and number of
receptors (29,31)
Peak of binding
at 4 weeks (62)
Adult binding le- D R has no effect
vel at 30 days
on binding pattern
Peak of binding
D R has no effect
at 4 weeks (17)
on developnlental
Peak of effective- profile (17)
ness during the
critical period (159)
DR, dark rearing; MD, monocular deprivation; O.D., ocular dominance
Involvement
in plastic
phenomena
postulated
strongly
suggested
disputed
not directly
involved
not directly
involved
not directly
involved
connectivity can be achieved, as has been suggested by Changeux and
Dachin (20) (see Table 11). Therefore it is not surprising that in mechanism underlying visual cortical plasticity the involvement of receptor alterations has been postulated (31, 124, 130, 131, 142). Cynader and
Shaw (31) proposed a hypothesis showing how receptor alterations might
be involved in binocular competition. They claimed that the asymmetry of the visual input as a result of MD causes the migration of the
receptor sites o n the surface of the visual neuron towards the input from
the open eye. A similar process has been suggested for the receptor sites
of other neurotransmitters and neuromodulators, which may also influence
the way in which the postsynaptic receptors of the primary visual afferents are reorganized. In consequence of this redistribution the nerve terminals from the deprived affereats lose the functional contact with the
postsynaptic neuron. This model, however, does not explain why the adult
cortex does not exhibit changes in cortical connectivity in response to
MD. It seems therefore that the reorganization of connectivity achieved
by the suggested receptor redistribution as a result of the changed visual
input requires some additional conditions to produce permanent alterations.
Ontogenetic studies suggest that NMDA receptolrs can act as med~ator in use dependent changes of neuronal response properties since
the peak of their density in the visual cortex occurs during the critical
period. Moreover, a block of thtse receptors prevents the O.D. shift
normally seen after a period of MD (Table 11). NMDA receptors could
also be involved in the consolidation of plastic changes as indicated b y
experiments with ketamine (109). The observation that intact cholinergic
and noradrenergic inputs are necessary to provoke the shift of O.D.
after MD during the critical period (12) may suggest that the facilitatory
action of these neuroltransmitters: on the visual cortical neurons can cause the rise of the response to an excitatory stimulus from the open eye
above some critical threshold necessary to activate the NMDA receptors,
and m consequence to produce permanent changes in the properties of
the visual neurons (Fig. 4).
The results reviewed imply, however, that not cmly the excitatory
input but also the inhibitory input is somehow involved in O.D. plasticity. Our results show that the GABA receptor sites are preferentially localized in layer IV (Table I). They respond to the asymmetry of
the visual input with an increase of binding activity. Together wlth
providing the data showing that bicuculline administration restores binocularity in MD cats, they might indicate the involvement of these
receptor sites in O.D. plasticity. Thus it seems reasonable to assume
that not only the imbalance of the excitatory input from the two eyes
may be responsible for loss of binocularity but also the changed activity of the inhibitory inputs may play a role. Recently, Artola and Singer
(5) have found that the induction of long-term potentiation (LTP) in the
visual cortex of the rat requires both the activation of the NMDA receptors and the concorninant reduction of GABA inhibition. These data may indicate that in the visual cortex NMDA mediated processes are
controled by GABAergic mechanisms. Interestingly, recent morphological studies of Einstein et al. (39) have demonstrated that part of the geExcltato~g
Input
\Release of Glu/Asp
from geniculocortical
nerve endings
\
A cti vatton of postsyn~ptic
non NMDA recepfors
Depo/o~izutionof
visual neupons
NA and ACh
inputs
-------t
Facilitatwy
action
Activation o f NMDA
receptors
I
i
Plastic changes
Fig. 4. Visual input evokes the release of excitatory neurotransmitter, presumably
Glu/Asp from geniculocortical nerve endings, and activates non NMDA glutamatergic receptors. This causes depolarization of pastsynaptic membrane. With coactivation of excitatory input and NA and ACh modulatory inputs the postsynaptic
membrane depolarization reaches a threshold for the activation of the NMDA receptors. The NMDA channel becomes permeable to calcium ions and long lasting
changes in the efficacy of excitatory synapses can be induced.
niculate synapses in layer IV appear to be syrhmetrical, suggesting that
they might be GABAergic. The authors assumed that both direct excitatory and inhibitory terminals. are important contributors to the receptive field properties of the visual cortex.
'
There are some suggestions that it is the intracortical synapses and
not geniculocortical synapses that are the first to be functionally altered
by the asymmetry of the visual input, since the changes in activation
in the striate cortex are first visible outside layer IV as revealed by 2DG studies (72). At the moment, however, it is not possible to decide
if the intracortical activity plays a primary role in O.D. plasticity.
In recent years a good deal of information about the presence of
neuropeptides in the visual cortex and their coexistence with classical
neurotransmitters has become available (83, 89, 99, 151). The regulation of long lasting events by neuropeptides may be of general importance in plastic phenomena occurring in the visual cortex during the
critical period. Especially the prolonged action of neuropeptides can be
a possible basis for permanent changes of the properties of cortical neurons after an abnormal visual input.
Long term modifications of the visual cortical neuron properties can
be accomplished by the processes mediated by second messengers. There
is more and more evidence that the activation of protein kinases is
associated with a prolonged modification of neuronal excitability (63,
95). It seems therefore reasonable to expect some changes in the phosphorylation processes after manipulation with the visual input. Immunocytochemical studies of Hendry and Kennedy (53) show that MD in
monkeys causes an increase in the concentration of calcium/calmodulin
dependent kinase I1 within the deprived eye columns. It has been found
that the rise in this kinase immunoreactivity does not simply reflect
a generalized increase in the synthesis of regulatory proteins inr the deprived neurons, because staining for synapsin I, a substrate for this
kinase, has been unaffected. These results seems to indicate that changing levels of activity in visual cortical neurons can alter their regulatory machinery. Aoki and Siekevitz (4) have investigated the developmental changes in cAMP stimulated phosphorylation of the cat's visual
cortex proteins. They report that dark rearing (DR) causes in vivo an
increase in cAMP dependent phosphorylation of microtubule-associated
protein (MAP 2) whereas exposure to light of DR animals results i n
a large decrease in phosphorylation of MAP 2. Since MAP 2 phosphorylation appears to play a crucial role in the &ssociation of cytoskeletal structure, the authors suggested, that following DR, the cytoskeletal
structure is in a relatively uncross-linked flexible state. After visual
stimulation, MAP 2 is present in vlsual neurons mostly in a deplhosphorylated state, and Aoki and Siekevitz (4) suggest that a cytoskeletal
structure of the scaffolding for the dendrites may be forming. It seems
that cAMP dependent phosphorylation and dephosphorylation of MAP-2
may be an important factor inducing plasticity during the critical period
7
- Acta Neurobiol. Exp.
6/88
through its action on the dendritic cytoskeletal organization involving
tubulin and actin.
An increase in tubulin synthesis in .the visual cortex of DR rats after
exposure to light during the critical period was reported by CronllyDillon and Perry (28). Then the present authors found (142, 147) that
short monocular stimulation (for several hours or 3 days) of normal or
binocularly deprived kittens produces an elevation of the tubulin level
in the visual cortex. Since the effect is disrupted by later binocular visual stimulation, we suggested that it may be related to a shift of O.D.
due to an asymmetric visual input. Thus it seems that the changed activity in the visual pathways can affect the neuronal cytoskeleton.
Neurotransmitters activating specific receptors affect, directly or indirectly via second messenger systems, the metabolism of neurons, especially protein phosphorylation. In this way long-lasting changes in synaptic connectivity can be geneqated. It seems that during the critical
period there exists a peculiar pattern of interactions between numerous
neurotransmitters and neuromodulators. This, together with the presumable action of some trophic factors, may create the unique conditions,
which enable the visual cortical neurons to change their properties a s
a result of alterations of the visual input.
REFERENCES
1. ALLEN, E. E., BLAKEMORE, L. J., TROMBLEY, P. Q. and GORDON, B.
1987. Effect of desmethylimipramine cm norepinepherine content and plasticity of kitten visual cortex. Brain Res. 401: 397-400.
2. AOKI, C. 1985. Development of the A 1 adenwine receptors in the visual
cortex of cats, dark reared and normally reared. Dev. Brain Res. 22: 125133.
3. AOKI, C., KAUFMAN, D. and RAINBOW, T. C. 1986. The ontogeny of the
4.
5.
6.
7.
8.
laminar distribution of (3-adrenergic receptors in the visual cortex of cats
normally reared and dark-reared. Dev. Brain Res. 27: 109-116.
AOKI, C. and SIEKEVITZ, P. 1985. Ontogenetic changes in the cyclic 3',5'-monophosphate-stimutable phosphorylation of cat visual cortex proteins, particularly of microtubule-associated protein 2 (MAP 2); Effects of normal
and dark rearing and of the exposure to light. J. Neurosci. 5: 2465-2483.
ARTOLA, A. and SINGER, W. 1987. Long-term potentiation and NMDA receptors in rat visual cortex. Nature 330: 649-652.
BAUGHMAN, R. W. and GILBERT, C. D. 1981. Aspartate and glutamate a s
possible neurotransmitters in the visual cortex. J. Neurosci. 1: 427-439.
BAUMFALK, U. and ALBUS, K. 1987. Baclofen inhibits the spontaneous and
visually evoked responses of neurmes in the striate cortex of the cat.
Neurosci. Lett. 75: 187-192.
BEAR, M. F., CARNES, K. M. and EBNER, F. F. 1985. An investigation of
ch?linergic circuitry in cat striate cortex using acetylcholinesterase histochemistry. J. Comp. Neurol. 234: 411-413.
-
9. BEAR, M. F., CARNES, K. M. and EBNER, F. F. 1985. Postnatal changes i n
the distribution of acetylcholinesterase in kitten ,striate cortex. J. Comp.
Neurol. 237: 519-532.
10. BEAR, M. F. and DANIELS, J. D. 1983. The plastic response to monocular
deprivation persists in kitten visual cortex after chronic depletion of noradrenaline. J. Neurosci. 3: 407-416.
11. BEAR, M. F., SCHMECHEL, D. E. and EBNER, F. F. 1985. Glutamic acid
decarboxylase i n the striate cortex of normal and monocularly deprived
kittens. J. Neurosci. 5: 1262-11275.
12. BEAR, M. F. and SINGER, W. 1986. Modulation of visual cortical plasticity
by acetylcholine and noradrenaline. Nature 320: 172-176.
13. BENNETT, E. L., ROZENZWEIG, M. R., DIAMOND, M. C., MORIMOTO,
H. and HERBERT, M. 1974. Effects of successive environmentis on brain
measures. Physiol. Behav. 12: 621-631.
14. BIGL, V. and SCHOBER, W. 1977. Cholinergic transmission in subcortical and
cortical visual centers of rats: No evidence for the involvement of primary optic system. Exp. Brain Res. 27: 211-219.
15. BLACK, J. B., ADLAR, J. E., DREYFUS, C. F., JONAKAIT, G. M., KATZ,
D. M., LaGAMMA, E. F. and MARKEY, K. M. 1984. Neurotransmitter plasticity a t the molecular level. Science 225: 1266-1270.
16. BLAKEMORE, C. and Van SLYTERS, R. C. 1974. Reversal of the physiological effects of monocular deprivation in kittens: further evidence for
a sensitive period. J. Physiol. 237: 195-216.
17. BODE-GREUEL, K. and SINGER, W. 1987. Developmental changes of APVsensitive [SHIglutamate binding sites in the visual cortex of kitten~sand
adult cats. Neuroscience (Suppl.) 22: S 229
18. BROWN, R. M., CRANE, A. M. and GOLDMAN, P. S. 1979. Regional distribution of monoamines in the cerebral cortex and subcortical structures of
the rhesus monkey: Concentrations and in vivo synthesis rates. Brain Res.
168: 133-150.
19. CARROLL, E. W. and WONG-RILEY, M. 1985. Correlation between cytochrom
oxidase staining and the uptake and laminar distribution of tritiate aspartate, glutamate, y-aminobutyrate and glycine in the striate cortex
of squirrel monkey. Neuroscience 15: 959-976.
20. CHANGEUX, J. P. and DACHIN, A. 1976. Selective stabilization of develo~ i n e ,synapses a s a mechanism for the specifitation of neuronal network.
Nature 264: 2697-2704.
21. CHARKRABARTI, T., DIAS, P. D., ROYCHOWDHURY, D. and DAGINAWALE H. F. 1974. Effect of unilateral visual deprivation on the activities
of acetylcholinesterase, cholinesterase and carbonic anhydrase of the optic
lobe of the pigeon. J. Neurochem. 22: 865-867.
22. CLARK, R. M. and COLLINS, G. G. S. 1976. The release of endogenous amino acids from the rat visual cortex. J. Physiol. (Lond.) 262: 383-400.
23. CLINE, H. T., DEBSKI, E. and CONSTANTINE-PATON, M. 1987. NMDA
receptor antagonist desegregates ocular dominance columns. Neuroscience
(Suppl.) 22: S 216.
24. C O R m S , R. and PALACIOS J. M. 1986. Muscarinic cholinergic receptor subtypes i n the rat brain. Quantitative autoradiographic studies. Brain Res.
362: 227-238.
25: CORTfCS, R., PROBST, A. and PALACIOS, J. M. 1987. Quantitative light mic-
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
roscopic autoradiographic localization of cholinergic muscarinic receptors
,
in the human brain: forebrain. Neuroscience 20: 65-107.
COTMAN, C. W. and IVERSEN, L. L. 1987. Excitatory amino a,cids i n the
brain - focus on NMDA receptosrs. Trends Neurosci. 10: 263-265.
CRAGG, B. G. 1975. The development of synapses i n the kitten visual cortex during visual deprivation. Exp. Neurol. 46: 445-451.
CRONLY-DILLON, J. and PERRY, G. W,. 1979. Effect of visual experience
on tubulin synthesis during a critical period of visual cortex development
in the hooded rat. J. Physiol. (Lond.) 293: 469-484.
CYMERMAN, U., PALACIOS, J. M., C O R m S , R. and SKANGIEL-KRAMSKA, J. 1987. Autorad~iographi'clocalization of muscarinic cholinergic receptors in visual areas of cat brain: variations i n sensitivity of N [3Hjmethylscopolamine biind4ing sites to carbachol and pirenzepine. Neurosci. Lett.
81: 13-18.
CYNADER, M. and MITCHELL, D. E. 1980. Period of susceptibility of kitten visual cortex to the effect of monocular deprivatio'n extends beyond
six months of age. Brain Res. 91. 545-550.
CYNADER, M. and SHAW, C,. 1986. Mechanisms underlying developmental
alterations of cortical ocular dominance. I n E. L. Keller and D. Zee (ed.),
Adaptive processes i n visual and oculomotor systems. Pergamon Press,
Oxford, p. 53-61.
DAW, N. W., VIDEEN, T. O,., PARKINSON, D. and RADER, R. K. 1985.
DSP-4(N-(2-chlorwthyl)-N-ethyl-2-bromobenzylamine)depletes noradrenaline i n kitten visual cortex without altering the effects of monocular deprivation. J. Neurosci. 5: 1925-1933.
DAW, N. W., VIDEEN, T. O., ROBERTSON, T. W,. and RADER, R. K. 1985.
An evaluation of the hypothesis that noradrenaliine affects plasticity in
the developing visual cortex. In A. Fein and J. S. Levine (ed.), The visual
system. Liss, New York, p. 133-144.
DEAN,' A. F., BUNCH, S. T., TOLHURST, D. J. and LEWIS, P. R. 1982. The
distribution of acetylcholinesterase i n the lateral geniculate nucleus of the
cat and monkey. Brain Res. 244: 123-134.
DeFELIPE, J. and JONES, E. G. 1985. Vertical organization of y-aminobutyric acid-accumulating intrinsic neuronal systems in monkey cerebral cortex. J. Neurosci. 5: 3246-3260.
De LIMA, A. D., MONTERO, IV. M. and SINGER, W. 1985. The cholinergic
innervation of the visual thalamus: a n EM immunocytochemical study.
Exp. Brain Res. 59: 206-212.
DUFFY, F,. H., SNODGRASS, S. R. and BURCHFIEL, J. L. 1976. Bicuculline
reversal of deprivation amblyopia i n the cat. Nature 260: 256-257.
DYKES, R. W., LANDRY, P., METHERATE, R. and HICKS, T. P. 1984.
Functional role of GABA i n cat primary somatosensory cortex: shaping
receptive fields of cortical neurons. J. Neurophysiol. 52: 1066-1093.
EINSTEIN, G., DAVIS, T. L. and STERLING, P. 1987. Ultrastructure of synapse from t h e A-laminae of the lateral geniculate nucleus in layer IV of
the cat striate cortex. J. Comp. Neurol. 260: 63-75.
FOSSE, V. M. and FONNUM, F. 1987. Biochemical evidence for glutamate
and/or aspartate as neurotransmitters i n fibers from the visual cortex
to the lateral posterior thalamic nucleus (pulvinar) in rats. Brain Res. 400:
219-224.
FOSSE, V. M., HEGGELUND, P., IVERSEN, E. and FONNUM, F. 1984. Effects
of area 17 ablation in neurotransmitter parameters in efferents to area
18, the lateral geniculate body, pulvinar and superior colliculus in the cat.
Neurosci. Lett. 52: 323-328.
GABBOTT, P. L. A. and SOMOGYI, I. 1986. Quantitative distribution of
GABA-immunoreactive neurons i n the cat visual cortex (area 17) of t h e
cat. Exp. Brain Res. 61: 311-323.
GAREY, L. J. and PETTIGREW, D. J. 1974. Ultrastructural changes in kitten
visual cortex after environmental modification. Brain Res. 66: 165-172.
GAUDIN-CHAZAL, G., DASZUTA, A., FAUDON, M. and TERNAUX, J . P.
1979. 5-HT concentration in cats brain. Brain Res. 160: 281-293.
GOFFINET, A. and ROCKLAND, K. S. 1985. Laminar distribution of al-and
8,-adrenoceptors in ferret visual cortex. Brain Res. 333: 11-17.
GOLDMAN-RAKIC, P. S. and BROWN, R. M. 1982. Postnatal development of
monoamine eontent and synthesis in the cerebral cortex of Rhesus monkey. Dev. B r a ~ nRes. 4: 339-349.
GRAYBIEL, A. M. and RAGSDALE Jr., C. W. 1982. Pseudocholinesterase
staining in primary visual pathway of the macaque monkey. Nature 299:
439-442.
GREENFIELD, S. 1984. Acetylcholinesterase may have novel functions in the
braln. Trends Neurosci. 7: 364-3M.
HAYDON, P. G., McCOBB, D. P. and KATER, S. B. 1987. The regulation
of neurite outgrowth, growth cone motility, and electrical synaptogenesis
by seroton~n.J. Neurobiol. 18: 197-215.
HENDRICKSON, A. 1985. Dots, stripes and columns in monkey visual cortex.
Trends Neurosci. 8: 406-410.
HENDRICKSON, A. E., HUNT, S. P. and WU, J. Y. 1981. Immunocytochemical localization of glutamic acid decarboxylase in monkey striate cortex.
Nature 292: 605-607.
HENDRY, S. H. C. and JONES, E. G. 1986. Reduction i n number of immunostained GABAergic neurones in deprived-eye dominance colums of monkey area-17. Nature 320: 750-753.
HENDRY, S. H. C. and KENNEDY, M. B. 1986. Immunoreactivity for a calmodulin-dependent protein kinase is selectively increased i n macaque striate cortex after monocular deprivation. Proc. Natl. Acad. Sci. USA 83:
1536-1540.
HESS, R. and MURATA, K. 1974. Effects of glutamate and GABA on specific
response of neurones i n the visual cortex. Exp. Brain Res. 21: 285-297.
HICKS, T. P., RUWE, W. D., VEALE, W. L. and VEENHUIZEN, J. 1987. Aspartate and glutamate as synaptic transmitters of parallel visual cortical
pathways. Exp. Brain Res. 58: 421-425.
HOCKFIELD, S., McKAY, R. D., HENDRY, S. H. C. and JONES, E. G. 1983.
A surface antigen that identifies mular dominance columns in the visual
cortex and laminar features of the lateral geniculate nucleus. Cold Spring
Harbor Symposia on quantitative Biology, Vol. XLVIII; Molecular Neurobiology, Cold Spring Harbor Laboratory, p. 877-889.
HOUSER, C. R., VAUGHM, J. E., HENDRY, S. H. C., JONES, E. G. and
PETERS, A. 1984. GABA neurons in the cerebral cortex. In E. G. Jones
and A. Peters (ed.), Cerebral cortex. Plenum Press, New York, Vol. 2,
I
58. HUBEL, D. H. and WIESEL, T. N. 1980. The period of susceptibility to the
physiological effects of unilateral eye closure in kittens. J. Physiol. (Lond.)
206: 419-436.
59. ISHIKAWA, K., WATABE, S. and GOTO, N. 1983. Lamina1 distribution of
y-aminobutyric acid (GABA) in the occipital cortex of rats: evidence as
a neurotransmitter. Brain Res. 277: 361-364.
60. IVERSEN, L. L., MITCHELL, J. F. and SRINIVASAN, V. 1971. The release
of y-amino-butyric acid during inhibition in the cat visual cortex. J. Physiol. (Lond.) 212: 519-534.
61. JASPER, H. H. and KOYAMA, I. 1969. Rate of release of amino acids from
the cerebral cortex is affected by brainstem-thalamic stimulation. C a n
J. Physiol. Pharmacol. 47: 889-905.
62. JONSSON, G. and KASAMATSU, T. 1983. Maturation of monoamine neurotransmitters and receptors in cat occipital cortex during postnatal critical
period. Exp. Brain Res. 50: 449-458.
63. KACZMAREK, L. K., STRONG, J. A, and KAUER, J . A. 1986. The role of
protein kinases in the control of prolonged changes in neuronal excitability. Prog. Brain Res. 69: 77-106.
64. KASAMATSU, T. 1983. Neuronal plasticity maintained by the central norepinephrine system in the cat visual cortex. Prog. Psychobiol. Physiol.
Psychol. 10: 1-111.
65. KASAMATSU, T. and PETTIGREW, J. D. 1976. Depletion of brain catecholarnines: failure of ocular dominance shift after monocular occlusion i n
kittens. Science 194: 206-209.
66. KASAMATSU, T., PETTIGREW, J. and ARY, M. 1979. Restoration of visual
cortical plasticity by local microperfusion of norepinephrine. J. Comp.
Neurol. 185: 163-182.
67. KASAMATSU, T. and SHIROKAWA, T. 1985. Involvement of 0-adrenorecep-8
tors in the shift of ocular dominance after monocular deprivation. Exp.
Brain Res. 59: 507-514.
68. KASAMATSU, T, and SHIROKAWA, T. 1985. Are Beta adrenoreceptors involved in vlsuocortical plasticity? In B. E. Will, P. Schmitt and J. C. Dalrymple-Alford (ed.), Brain plasticity, learning, and memory. Plenum FUblishing Co., p. 79-84.
69. KLEINSCHMIDT, A., BEAR, M. F. and SINGER, W. 1986. Effects of thei
NMDA receptor antagonist APV on visual cortical plasticity in monocularly deprived kittens. Neurosci. Lett. Suppl. 26: S58.
70. KLEINSCHMIDT, A., BEAR, M. F. and SINGER, W. 1987. Blockade of
"NMDA" receptors disrupts experience-dependent plasticity of kitten striate cortex. Science 238: 355-358.
71. KOSOFSKY, B. E., MOLLIVER, M. E., MORRISON, J. H. and FOOTE, S. L.
1984. The serotonin and norepinephrine innervation of primary visual cortex in the cynomolgus monkey (Macaca fascicullaris). J. Comp. Neurol.
230: 168-178.
72. KOSSUT, M., THOMPSON, I. D. and BLAKEMORE, C. 1983. Ocular dominance columns in cat striate cortex and effects of monocular deprivation:
A 2-deoxy-glucose study. Acta Neurobiol. Exp. 43: 273-282.
73. KOSSUT, M., WOJCIK, M. and SKANGIEL-KRAMSKA, J. 1981. Dynamic
changes of serotonin levels during the first visual experience in binocularly deprived kittens. J. Neurochem. 37: 1077-1080.
74. KRATZ, K. E.7 SPEAR, D. and SMITH, D. C. 1976. Postcritical period reversal of effects of mOn0~ular deprivation on striate cortex cells i n the cat.
J . Neurophysiol. 39: 501-511.
75. KRNJEVIC, K. 1984. Neurotransmitter i n cerebral cortex a general account.
In E. J. Jones and A. Peters (ed.), Cerebral cortex. Vol. 2. Pergamon Press,
New York, p. 39-61.
76. KUPFER, C. and PALMER, P. 1984. Lateral geniculate nucleus, histological
and cytochemical changes following afferent denervation and visual deprivation. Exp. Neurol. 9: 400-409.
77. KVALE, I. and FONNUM, F. 1983. The effects of unilateral neonatal of visual
cortex on transmitter parameters in the adult superior colliculus and lateral geniculate body. Dev. Brain Res. 11: 261-266.
78. KVALE, I., FOSSE, V. M. and FONNUM, F. 1983. Development of neurotransmitter parameters i n lateral geniculate body, superior colliculus and visual cortex of t h e albino rat. Dev. Brain Res. 7: 137-145.
79. LAI, H., MAKOUS,'w. L., QUOCK, R. H. and HORITA, A. 1978. Visual deprivation affects serotonin levels in the visual system. J. Neurochem. 30:
1187-1189.
80. Le VAY, S. and GILBERT, C. D. 1976. Laminar patterns of geniculo-cortical
projection in the cat. Brain Res. 113: 1-19.
81. Le VAY, S., STRYKER, M. P. and SHATZ, C. J. 1978. Ocular dominance column and their development i n layer IV of the cat's visual cortex: a quantitative study. J. Comp. Neurol. 179: 223-244.
,
82. Le VAY, S., HUBEL, D. H. and WIESEL, ?]. N. 1975. The pattern of ocular
dominance columns in macaque visual cortex revealed by reduced silver
stain. J. Comp. Neurol. 159: 559-576.
83. LIN, C. S,, LU, S. M. and SCHMECHEL, D. E. 1986. Glutamic acid decarboxylase and somatostatin immunorad~oactivityin rat visual cortex. J. Comp.
Neurol. 244: 369-383.
84. MALETTA, G. J. and TIMIRAS, P. S. 1967. Acetylcholinesterase activity i n
optic structures after complete light deprivation from birth. Exp. Neurol.
19: 513-518.
83. McDONALD, J. K., SPECIALE, S. G. and PARNAVELAS, J. G. 1981. The
development of glutamate decarboxylase i n the visual cortex and the
dorsal lateral geniculate nucleus of the rat. Brain Res. 217: 364-367.
86. METHERATE, R., TREMBLAY, N. and DYKES, R. W. 1987. Acetylcholine
permits long-term enhancement of neuronal responsiveness in cat primary
somatosensory cortex. Neuroscience 22: 75-81.
87. MORRISON, J. H. and FOOTE, S. L. 1986. Noradrenergic and serotonergic innervation of cortical, thalamic and tectal visual structures in Old and New
World monkeys. J. Comp. Neurol. 243: 117-138.
8& MORRISON, J. H., FOOTE, S. L., MOLLIVER, M. E., BLOOM, F. E,. and
LIDOV, H. G. W. 1982. Noradrenergic and serotonergic fibers innervate
complementary layers in monkey primary visual cortex: An immunohistochemical study. Proc. Natl. Acad. Sci. USA 79: 2401-2405.
89. MORRISON, J. H. and MAGISTRETTI, P. J . 1983. ~ o n o a m i n e sand peptldes in cerebral cortex-contrasting principles of cortical organization. rends
Neurosci 6: 146-151.
90. MOWER, G. D., BERRY, D., BURCHF'IELD, J. L. and DUFFY, F. H. 1981.
Comparison of the effects of dark-rearing and binocular suture on development and plasticity of cat visual cortex. Brain Res. 220: 255-267.
91. MOWER, G. D., CHRISTEN, W. G. and CAPLAN, C. J. 1983. Very brief visual
experience eliminates plasticity in the cat visual cortex. Science 221: 178180.
92. MOWER, G. D., WHITE, W. F. and RUSTAD, R,. 1986. [3H] ~ u s c i m o lMnding
of GABA receptors in the visual cortex of normal and monocularly deprived cats. Brain Res. 380: 253-260.
93. MURPHY, S., UZBEKOV, M. G. and ROSE, S. P. R. 1980. Changes i n serotonin receptors in different regions after light exposure of dark-reared rats.
Neurosci. Lett. 17: 317-321.
94. NEEDLER, M. C., SHAW, C. and CYNADER, M. 1984. Characteristics and distribution of muscimol binding sites in c a t visual cortex. Brain Res. 308:
347-353.
95. NESTLER, E. J., WALAAS, S. I. a n d GREENGARD, P. 1984. Neuronal phosphoproteins: physiological and clinical implications. Science 225: 1337-1364.
96. PALACIOS, J. M. a n d KUHAR, M. J. 1980. Beta-adrenergic receptor localization by light microscopic autoradiography. Science 208: 1378-1380.
97. PALACIOS, J. M. and KUHAR, M. J. 1982. Beta-adrenergic receptor localization i n r a t brain by light microscopic autoradiography. Neurochem. Int.
4: 473-490.
98. PAPADOPOULOS, G. C., PARNAVELAS, J . G. and BUIJS, R. 1987. Monoaminergic fibers form conventional synapses in the cerebral cortex. Neurosci. Lett. 76: 275-279.
99. PAPADOPOULOS, G. C., PARNAVELAS, J. G. and COVANAGH, M. E.
1987. Extensive co-existence of neuropeptides i n t h e r a t visual cortex. Brain
Res. 420: 95-100.
100. PARKINSON, D. and DAW, N. W. 1986. [3H]-Nicotine binding in cat visual
cortex and receptor autoradiography. Soc. Neurosci. (Abstr.) 12: 582.
101. PAZOS, A., C O R T S , R. and PALACIOS, J . M. 1985. Quantitative autoradiographic mapping of serotonin receptors in the r a t brain. 11. Serotonin-2
receptors. Brain Res. 346: 231-249.
102. PAZOS, A. and PALACIOS, J . M. 1985. Quantitative autoradiographic mapping of serotonin receptors in the r a t brain. I. Serotonin-1 receptors. Brain
Res. 346: 205-230.
103. PAZOS, A., PROBST, A. and PALACIQS, J. M. 1987. Serotonin receptors
in the human brain - 111. Autoradiographic mapping of serotonin-1 receptors. Neuroscience 21: 97-122.
104. PAZOS, A,, PROBST, A. and PALACIOS, J. M. 1987. Serotonin receptors i n
the human brain - IV. Autoradiographic mapping of serotonin-2 receptors. Neuroscience 21: 123-139.
105. PHILIPSON, 0. T., KILPATRICK, I. C. and JONES, M. W. 1987. Dopaminergic innervation of the primary visual cortex in the rat, a n d some correlations with human cortex. Brain Res. Bull. 18: 621-633.
106. POTEMPSKA, A., SKANGIEL-KRAMSKA, J. a n d KOSSUT, M. 1979. Development of cholinergic enzymes and adenosine-triphosphate activity of
optic system of cats in normal and restricted visual input conditions. Dev.
107. PRUSKY, G. and CYNADER, M. 1986. The distributions of nicotinic a n d muscarinic acetylcholine binding sites in the developing cat visual cortex. Soc.
Neurosci. (Abstr.) 12: 1372.
108. PRUSKY, G. T., SHAW, C. and CYNADER, M. S. 1987. Nicotine receptors
are located a n lateral geniculate nucleus terminals i n cat visual cortex.
Brain Res. 412: 131-138.
RAUSCHECKER, J. P. and HAHN, S. 1987. Ketamine-xylazine anaesthesia
blocks consolidation of ocular dominance changes i n kitten visual cortex.
Nature 326: 183-185.
READER, T. A. 1978. The effect of dopamine, noradrenaline and serotonin
in the visual cortex of the cat. Experientia 34: 1586-1587.
READER, 'T. A., FERRON, A,, DESCARRIES, L. and JASPER, H. H. 1979.
Modulatory role for biogenic amines in the cerebral cortex. Micro-iontophoretic studies. Brain Res. 160: 217-229.
.
RIBAK, C. E. 1978. Aspinous and sparsly-spinlous stellate neurons in the visual cortex of rats contain glutamic acid decarboxylase. J. Neurocytol. 7:
461-478.
ROBERTS, J. 1986. What do GABA neurons really do? They make possible
variability generation in relation to demand. Exp. Neurol. 93: 279-290.
ROBERTSON, R. T. 1987. A morphogenic role for transiently expressed acetylcholinesterase i n developing thalamocortical systems? Neurosci. Lett.
75: 259-264.
ROBERTSON, R. T., FOGOLIN, R. P., TIJERINA, A. A. and YU, J. 1987.
Effects of neonatal monocular and binocular enucleation on transient acetylcholinesterase activity in developing rat visual cortex. Dev. Brain Res.
33: 185-187.
ROBERTSON, R. T., TIJERINA, A. A. and GALLIVAN, M. E. 1985. Transient patterns of acetylcholinesterase activity in visual cortex of the rat:
normal development and the effects of neonatal enucleation. Dev. Brain
Res. 21: 203-214.
ROSE, S. P. R. and STEWART, M. G. 1978. Transient increase in muscarinic
acetylcholine receptor and acetylcholinesterase in visual cortex on first
exposure of dark-reared rats to light. Nature 271: 169-170.
SATO, H. and TSUMOTO, T. 1984. GABAergic inhibition already operates
o n a grsoup of neurons i n the kitten visual cortex a t the time of eye opening. Dev. Brain Res. 12: 311-315.
SCHLIEBS, R., BIGL, V. and BIESOLD, D. 1982. Development of muscarinic
cholinergic receptor binding in the visual system of monocularly deprived dark reared rats. Neurochem. Res. 7: 1181-1197.
SCHLIEBS, R., BURGOYNE, R. D. and BIGL, V. 1982. The effect of$ visual
deprivation on 6-adrenergic receptors in the visual centres of the rat.
J. Neurochem. 38: 1038-1043.
SCHLIEBS, R. W. and ROTHE, T. 1987. Development of GABA A receptors
i n the central visual structures of rat brain. Effect of visual pattern deprivation. Neuroscience (Suppl.) 22: S 235.
SHATZ;C. J., LINDSTROM, S., and WIESEL, T. N. 1977. The distribution
of afferents representing the right and left eyes in the cat's visual cortex. Brain Res. 39: 41-48.
SHATZ, C. J., and STRYKER, M. P. 1978. Ocular dominance in layer I V of
the cat's visual cortex and the effect of monocular deprivation. J. Physiol.
(Lond.) 281: 267-283.
SHAW, C., AOKI, C., NEEDLER, M. C., WILKINSON, M. and CYNADER,
M. 1983. Ontogenesis of 6-adrenergic, GABA and benzodiazepine receptor
binding sites i n cat visual colltex and the effects of dark rearing, In Abstracts of 3rd Capo Boi Conference on Neuroscience. The cell biology of
-
neuranal plasticity. Capo Boi Hotel Villasimius, Sardinia, Italy, p. 283-284.
125. SHAW, C. and CYNADER, M. 1984. Disruption of cortical activity prevents
ocular dominance changes in monocularly deprived kittens. Nature 308:
731-734.
126. SHAW, C. and CYNADER, M. 1987. Increase i n GABAA receptor number
in cat visual cortex after early visual deprivabon. Neuroscience (Suppl.)
22: S 289.
127. SHAW, C., HALL, S. and CYNADER, M. 1986. Characterization, distribu-
tion and ontogenesis of adenosine binding sites in cat visual cortex.
J. Neurosni. 6: 3218-3228.
128. SHAW, C., NEEDUR, M. C. and CYNADER, M. 1984. Ontogenesis of muscimol binding sites in cat visual cortex. Brain Res. Bull. 13: 331-334.
129. SHAW, C., NEEDLER, M. C. and CYNADER, M. 1984. Ontogenesis of muscarinic acetylcholine binding sites in cat v,isual cortex: Reversal of specific laminar distribution duning critical period. Dev. Brain Res. 14: 285299.
130. SHAW, C., NEEDLER, M. C., WILKINSON, M., AOKI, C. and CYNADER, M.
1985. Modification of neurotransmitter receptor sensitivity in cat visual
cortex during the critical period. Dev. Brain Res. 22: 67-73.
131. SHAW, C., WILKINSON, M., CYNADER, M., NEEDLER, M. C., AOKI, C. and
HALL, E E. 1986. The laminar distribution and postnatal development
of neurotransmitter and neuromodulator receptors in cat yisual cortex.
Brain Res. Bull. 16: 661-671.
132. SHIROKAWA, T. and KASAMATSU, T. 1986. Concentration-dependent supression by 0-adrenergic antagonists of the shift in ocular dominance following monocular deprivation in kitten visual cortex. Neuroscience 18:
1035-1046.
133. SILLITO, A. M. 1975. The contribution of inhibitory mechanism to the recep-
tive field properties of neurones in the striate cortex of the cat. J. Physiol.
(Lond.) 250: 305-329.
134. SILLITO, E. 1975. The effectiveness of bicuculline as an antagonist of GABA
and visually evoked inhibition in the cat's striate cortex. J. Physiol. (Lond.)
250: 287-304.
135. SILLITO, A. M. 1984. Functional considerations of the operation of GABAergic inhibitory processes in the visual cortex. In E. G. Jones and A.
Peters (ed.), Cerebral cortex. Vol. 2. Plenum Press, New York, p. 91-117.
136. SILLITO, A. M. 1986. Cholinergic input and plasticity. Nature 320: 109-110.
137. SILLITO, A. M. 1986. Conflicts in the pharmacology of visual cortical plasticity. Trends Neurosci. 9: 301-303.
138. SILLITO, A. M. and KEMP, S. A. 1983. Cholinergic modulation of the functional organization of the cat visual cortex. Brain Res. 289: 143-155.
139. SILLITO, A. M., KEMP, J. A. and BLAKEMORE, C. 1981. The role of
GABAergic inhibition in the cortical effects of monocular deprivation.
Nature 291: 218-320.
140. SILLITO, A. M., KEMP, J. A. and PATEL, H. 1980. Inhibitory interactions
contributing to the ocular dominance of monocularly dominated cells in the
normal cat striate cortex. Exp. Brain Res. 41: 1-10.
141. SINGER, W., KLEINSCHMIDT, A. and BEAR, M. F. 1986. ~nfusion of a n
NMDA receptor antagonist disrupts ocular dominance plasticity in kitten
striate cortex. Soc. Neurosci. (Abstr.) 12: 786.
SKANGIEL-KRAMSKA, J. 1980. Biochemical approach to the studies on plasticity of cerebral cortex of visually deprived cats. In M. Brzin, D. Sket
and H. Bachelard (ed.), Synaptic constituents in health a n d disease. Mladinska Knjiga-Pergamon Press, Ljubljana, Oxford p. 353-378.
SKANGIEL-KRAMSKA, J., CYMERMAN, U. and KOSSUT, M. 1985. Distribution of GABA, serotonin and muscarinic receptors in visual structures
of kitten brain. Neurosci. Lett. (Suppl.) 22: S 384.
SKANGIEL-KRAMSKA, J., CYMERMAN, U. and KOSSUT, M. 1986. Autoradiografic localization of GABAergic and muscarinic cholinergic receptor
sites in t h e visual system of the kitten. Acta Neurobiol. Exp. 46: 119-130.
SKANGIEL-KRAMSKA, J. and KOSSUT, M. 1984. Increase of GABA receptor binding activity after short lasting monocular deprivation in kittens.
Acta Neurobiol. Exp. 44: 33-39.
SKANGIEL-KRAMSKA, J. and KOSSUT, M. 1985. Monocular deprivation
affects GABA receptor i n the visual cortex of kittens. Physiol. Bohemoslov. 34: 145-147.
SKANGIEL-KRAMSKA, J., KOSSUT, M. and POTEMPSKA, A. 1983. Monocular stimulation induces tubulin response in the kittens visual cortex.
In Abstracts of 3rd Capo Boi Conference on Neuroscience. The cell biology of neuronal plasticity. Capo Boi Hotel Villasimius, Sardinia, Italy, p. 285.
SOLNICK, B., DAVIS, T. L. and STERLING, P. 1984. Numbers of specific
types of neurons in layer IV a b of cat striate cortex. Proc. Natl. Acad.
Sci. USA 81: 3898-3900.
SOMOGYI, P.; COWEY, A,, HALASZ, N. and FREUND, T. F. 1981. Vertical
organization of neurones accumulating 3H-GABA in visual cortex of rhesus
monkey. Nature 294: 761-763.
SOMOGYI, P., FREUND, T. W., WU, J. Y. and SMITH, A. D. 1983. The
section-Golgi impregnation procedure. 2. Immunocytochemical demonstration of glutamate deca~boxylase in Golgi-impregnated neurons and their
afferent synaptic boutms in the visual cortex of the cat. Neuroscience 9:
475-490.
SOMOGYI, P., HODGSON, A. J., SMITH, A. D., NUNZI, M. G., GORIO, A.
and WU, J. Y. 1984. Different populations of GABA-ergic neurons in the
visual cortex and hippocampus olf cat contain somatostatin- or cholecystokinin-immunoreactive material. J. Neurosci. 4: 2590-2603;
STICHEL, C. C., De LIMA, A. D. and SINGER, W. 1987. A search for choline acetyltransferase-like immunoreactivity in neurons of the cat striate
cortex. Brain Res. 405: 395-399.
STICHEL, C. C. and SINGER, W. 1985. Organization and m~orphologicalcharacteristics of choli~neacetyltransferase-containing fibres i n the visual thalamus and striate cortex of the cat. Neurosci. Lett. 53: 155-160.
STICHEL, C. C. and SINGER, W. 1987. Quantitative analysis of the choline
acetyltransferase-immunoreactive axonal network in the cat primary visual
cortex: I. Adult cats. J. Com~p.Neurol. 258: 91-98.
STICHEL, C. C. and SINGER, W. 1987. Quantitative analysis of the choline
acetyltransferase-immunoreactive axonal network in the cat primary visual
cortex: 11. Pre- and postnatal development. J. Comp. Neurol. 258: 99-111.
THOMSON, A. M., WEST, D. C. and LODGE, D. 1985. An N-methylawpartate
157.
158.
159.
160.
161.
162.
163.
164.
165.
166.
167.
168.
169.
receptor mediated synapse in r a t cerebral cortex: a site of action of ketamine? Nature 313: 479-481.
TIEMAN, S. B. 1985. The anatomy of geniculocortical connections in monocularly deprived cats. Cell. Mol. Neurobiol. 5: 35-45.
TIGGES, J. and TIGGES, M. 1985. Subcortical sources of direct projections
to visual cortex. In A. Peters and E. G. Jones (ed.), Cerebral cortex. Plenum Press, New York, Vol. 3, p. 351-378.
TSUMOTO, T., HAGIHARA, K., SATO, H. and HATA, Y. 1987. NMDA receptors i n the visual cortex of young kittens are more effective than those
of adult cats. Nature 327: 513-514.
TSUMOTO, T., MASUI, H, and SATO, H. 1986. Excitatory amino acid transmitters in neuronal circuits of t h e cat visual cortex. J. Neurophysiol.
55: 469-483.
TURLEJSKI, K. and KOSSUT, M. 1985. Decrease in the number of synapses
formed by subcortical inputs t o the striate cortex of binocularly deprived
cats. Brain Res. 331: 115-125.
VIDEEN, T. O.,DAW, N. W. and RADER, R. K. 1984. The effect of norepinephrine on visual cortical neurons in kittens and adult cats. J . Neurosci. 4:
1607-1617.
WENK, H., BIGL, V. and MEYER, U. 1981. Ch'olinergic projections from magnocellular nuclei of the basal forebrain to cortical areas in rats. Brain
Res. Rev. 2: 211-219.
WIESEL, T. N. and HUBEL, D. H. 1963. Single-cell responses in the striate
cortex of kittens deprived of vision in one eye. J. Neurophysiol. 26: 10031017.
WILKINSON, M., SHAW, C., KHAN, I. and CYNADER, M. 1983. Ontogenesis
of 8-adrenergic binding sites in kitten visual cortex and t h e effects of
visual deprivation. Dev. Brain Res. 7: 349-352.
WINFIELD, D. A. 1983. The postnatal develo@ent of synapses in the different laminae of the visual cortex in the normal kitten and in kitten with
eyelid closure. Dev. Brain Res. 9: 155-169.
WOLF, W., HICKS, T. P. and ALBUS, K. 1986. The contribution of GABAmediated inhibitory mechanisms to visual response properties of neurons
in the kitten's striate cortex. J . Neurosci. 6: 2779-2795.
WOLFF, J. R., BOTTCHER, H., ZETZSCHE, T., OERTEL, W. H, and CHRONWALL, B. M. 1984. Development of GABAergic neurons in r a t visual cartex a s identified by glutamate decarboxylase-like immunoreactivity. Neurosci. Lett. 47: 207-212.
YOUNG, W. S. and KUHAR, M. J. 1980. Noradrenergic alpha-1 a n d alpha-2
receptors: light microscopic localization. Proc. Natl. Acad. Sci. USA 77:
1696-1700.
Accepted 26 May 1988