You are on page 1of 12
Journal of Clinical Nearophysolory TNgis20 41 Raven Press, Ld New York (© 998 Amendan Eleceoencephilopaphe Society Basal Ganglia-Thalamocortical Circuits: Their Role in Control of Movements Garrett E. Alexander Department of Neurology, Emory University School of Medicine, Atlanta, Georgia, U.S.A. ‘Summary: The motor circuitry of the basal ganglia is organized in parallel with basal ganglia circuits involved in oculomotor, associative, and limbic functions. ‘These circuits comprise reentrant pathways that originate in various neocortical domains and pass through the basal ganglia before returning, by way of the thala- ‘mus, to select portions of the frontal lobe, Recent progress in characterizing the anatomy and physiology of these pathways has provided new insights into the pathophysiology of basal ganglia-rclated movement disorders and enhanced ‘our understanding of the normal role of the basal ganglia in movement control and adaptive motor behavior. Key Words: Basal ganglia—Thalamus—Motor behavior FUNCTIONAL ORGANIZATION OF BASAL GANGLIA The basal ganglia comprise the striatum (including, the neostriatum, composed of the putamen and cau- date nucleus, and the ventral striatum), the globus pallidus, the substantia nigra, and the subthalamic nucleus, all of which are functionally subdivided into skeletomotor, oculomotor, associative, and limbic territories based on their physiological properties and on their interconnections with cortical and thalamic territories of the same functionalities (Alexander et al., 1990). As indicated in Fig. 1, the large-scale orga- nization of the basal ganglia can be viewed asa family of reentrant loops that are organized in parallel, each taking its origin from a particular set of functionally related cortical fields (skeletomotor, oculomotor, etc.), passing through the functionally corresponding portions of the basal ganglia, and returning to parts of those same cortical fields by way of specific basal ganglia-recipient zones in the dorsal thalamus. Due to maintained segregation along each of these ‘Address corespondence and reprint requests to Dr. GE. Alex. ander at Department of Neurology, Emory University School of Medicine, WMB-6000, 1639 Pieree Dr, P.O. Drawer V, Atlanta, GA 30323, USA. 420 cortico-basal ganglia-thalamocortical circuits, there is little direct communication among the separate functional domains. While virtually the entire cortical mantle is mapped topographically onto the striatum, often considered the “input” portion of the basal gan- alia, the cortically directed signals from the basal gan- glia output nuclei (internal pallidum and substantia a pars reticulata) are returned exclusively to foci within the frontal lobe (after first passing through the corresponding portions of the thalamus.) Because of their parallel organization, it is generally suspected that the operations performed at corresponding sta- tions along each of these loops are quite similar. In this review, the focus will be on the organization and physiology of the skeletomotor loop. MOTOR CIRCUITRY OF BASAL GANGLIA ‘The skeletomotor circuitry of the basal ganglia in- cludes several noteworthy features, some of which may be important in shaping the specific contribu- tions that these structures make to the control of movement. The broad outlines of basal ganglia motor circuitry are depicted in Fig. 2. Like the cerebellum, the basal ganglia receive topo- FIG. 1. Parallel, functionally scaresated, basal ‘ganglis-thalamocortical loops Information Nov is irom narrow to wide end ofeach pathway graphic inputs from most of the sensorimotor territo- ries of the cerebral cortex, including primary and sec ondary somatosensory areas, primary motor cortex, and a variety of premotor areas including the supple- mentary motor area, the dorsal and ventral premotor areas, and the cingulate motor areas (Kiinzle, 1975, 1977, 1978; Selemon and Goldman-Rakic, 1985; Ca- vada and Goldman-Rakic, 1991; Dum and Strick, 1991; Flaherty and Graybiel, 1991, 19936; Hoover and Strick, 1993; Yeterian and Pandya, 1993). Di- rected toward the sensorimotor portion of the neostri- atum, these coordinated corticostriatal inputs impose a well defined somatotopic organization upon their target nucleus, which, in primates, coincides roughly with the putamen (Crutcher and DeLong, 1984a; Al- exander and DeLong, 1985a: Crutcher and Alexan- der, 1990). In addition to their respective corticostria- tal neurons, most of the putamen-projecting sensori- motor areas also contain separate populations of neurons that project directly to the spinal cord (Hum- ‘melsheim et al., 1986; Hutchins et al., 1988; Dum and Strick, 1991). Some evidence suggests that the cortical message conveyed to the striatum may differ from that projected directly to the spinal cord (Bauswein et al., 1989). If'so, the prevailing view that the basal gan- lia receive cortical signals that are faithful replicas of corticospinal commands, i.e., simple corollary dis- charges, may need to be revised. The majority of the putamen-projecting sensorimo- tor areas also send topographic projections to the sub- thalamic nucleus (Hartmann-von Monakow et al GLIA-THALAMOCORTICAL CIRCUITS IN MOVE: 1978). Both the corticostriatal and the corticosubtha- lamic projections are excitatory and probably gluta- matergic (Wilson, 1986; Cherubini et al., 1988; Na- kanishi et al., 1988; Kawaguchi et al., 1989) [as are most of the corticospinal projections (Valtschanoff et al, 1993)], but the conduction velocities of these two pathways may differ significantly. Direct comparisons in the rat suggest that the corticostriatal projection ‘may be substantially slower than the corticosubtha- lamic (Kitai and Deniau, 1981), which could have a significant influence on the way information is pro- cessed within the basal ganglia. Thus far, direct com- parisons of these two pathways have not been made in the monkey, but it does seem clear from primate stud- ies that conduction along the corticostriatal pathway is substantially slower than conduction along the pathways leading from cortex to brainstem and spinal levels (Bauswein et al., 1989) Nevertheless, there are significant species differ- ences in basal ganglia connectivity, so extrapolations must be made with caution. In rats, for example, the ‘earliest cortically induced EPSPS in striatum appear to come from axon collaterals of large-diameter, brainstem-projecting fibers (Donoghue and Kitai, 1981), but in primates this collateral projection seems to be lacking, with the massive corticostriatal pro- jection arising almost exclusively from a separate pop- ulation of small-axoned, direct-projecting corticostri- atal neurons (Jones et al., 1977). In the putamen, as in the rest of the striatum, the large majority of neurons (>75% in primates, up to J.Clin Neurophysiol, Vol 1, No.4, 1998 422 G, E, ALEXANDER 95% in rodents) are of the medium spiny type (MSN) (Wilson and Groves, 1980). MSNs are y-aminobut- yrate-(GABA)-ergic projection neurons (Kita and Ki- tai, 1988), and in primates they comprise two distinet populations, one projecting to the external segment of the globus pallidus (GPe) and the other projecting cither to the internal segment of the globus pallidus (GPi) or to the substantia nigra pars reticulata (SNr) (Selemon and Goldman-Rakic, 1990; Flaherty and Graybiel, 1993a), At rest, putamen projection neu- rons are nearly silent (Alexander and DeLong, 19855; Kimura ct al., 1990), but most have well defined sen~ sorimotor fields and discharge selectively in relation to specific parameters of movement or specific aspects of the preparation for movement (Liles, 1985; Alex- ander and Crutcher, 1990a,b; Crutcher and Alexan- der, 1990; Kimura et al., 1992). ‘The other type of putamen neuron that has been well studied is the large, aspiny interneuron (Wilson et al., 1990; Kawaguchi, 1992). Converging evidence from anatomical studies as well as from extracellular and intracellular recording studies indicates that these neurons have the following properties. Unlike MSNs, the large interneurons are spontaneously active (with discharge rates of 3~7 Hz), and they do not discharge in relation to specific parameters of movement prepa- ration or execution. Rather, they discharge briefly (only one or two spikes) and synchronously following the presentation of a conditioned sensory stimulus 1.Clin, Neurophysiol, Vol 1, No.4, 1994 FIG. 2. Basal ganglia skeletomotor circ ‘Major motor pathways involving the basal sla, Excitatory pathways are indicated by light Shading. inhibitory by dark shading. Pathway Srigins and terminations are indicated by nar- ow and wide ends, respectively. Connections tetween thalamus and cortex are reciprocal [Not shown are connections to and from SNr, ‘which parallel those of GPi. GPe, extemal es. tent of globus pallidus; GPi, internal segment fof globus pallidus, NRT, reticular nucleus ofthe thalamus. PPN, pedunculopontine nucleus: ‘SNe, substantia nigra pars compacta; STN, sub- thalamic nucleus; V-thalamus, basal ganeli e- ‘Gpient zone in ventrolateral thalamus. To sim- Diy the illustration, the thalamostriate proj- ction from the centromedian nucleus is shown arising from VLthalamus. ‘that signifies the imminent delivery of a reward (Ki- mura, 1986). These spontancously active interneu- rons appear to be cholinergic (Bolam et al., 1984), but their synaptic influence on neighboring MSN has not een directly determined. Electron microscopic stud- jes have shown that cholinergic synapses on the So- mata and dendritic shafts and spines of MSNs are of the symmetrical type (Phelpset al., 1985), which sug gests that the aspiny interneurons may have an inhib- itory effect on MSNs. “Those portions of GPi and SNr that receive input from the putamen constitute the sensorimotor output nuclei of the basal ganglia, sending GABAergic pro- jections to their respective targets in both the ventro- Jateral thalamus (n, ventralis lateralis and n. ventralis anterior) and the centromedian nucleus of the thala~ ‘mus (Carpenter et al,, 1976; Kim et al., 1976; Ueki et al., 1977; Uno et al., 1978; Penney and Young, 1981; De Vito and Anderson, 1982; Yamamoto et al., 1985). Anatomically GPi and SNr share many com- ‘mon features, and from a functional standpoint they ‘can be viewed as two subdivisions of a single basal ganglia output nucleus (Nauta, 1979; Parent, 1986). In GPi/SNr, large projection neurons are by far the ‘most abundant cell type, but a minute population of ‘small cells that may represent interneurons has also been described (Fox and Rafols, 1976; Difiglia and Rafols, 1988). Neurons in the sensorimotor territories of GPi/SNr have movement-related receptive fields BASAL GANGLIA-THALAMOCORTICAL CIRCUITS IN MOVEMENT 423 FIG.3, Direct and indirect pathways through basal ganglia. Conventions same asin Fig. 2. that are similar to those of the putamen’s MSNs, but GPi/SNr neurons differ from MSNs in having rela- tively high spontaneous discharge rates (60-80 Hz) that are modulated both upward and downward de- pending on their particular sensorimotor response fields (DeLong et al., 1983; Mitchell et al., 1987; Ha- mada et al., 1990), Neurons in the sensorimotor terri- tory of GPe have discharge properties and sensorimo- tor fields that are comparable with those in GPi/SNr (Mitchell et al., 1987). The basal ganglia output nuclei (GPi/SNr) receive not only direct inputs from the striatum via the GPi- and SNr-projecting MSN, but also an important con- verging source of information via what has come to be termed the “indirect” pathway through the basal ganglia (Alexander and Crutcher, 1990c). The direct and indirect pathways are illustrated in Fig. 3. The in- direct pathway takes its origin from the GPe-project- ing MSNs. In turn, the GABAergic neurons of GPe project mainly to the subthalamic nucleus (Carpenter and Jayaraman, 1990), whose excitatory, gluta- matergic neurons send feed-forward connections to GPi/SNr, completing one arm of the indirect pathway (Kita and Kitai, 1987; Smith and Parent, 1988; Ro- bledo and Feger, 1990). Subthalamic neurons also send glutamatergic feedback connections to GPe and to the putamen (Nakano et al., 1990; Parent, 1990; Robledo and Feger, 1990). Recently a second arm of the indirect pathway has been revealed: GPe projects not only to the subtha- lamic nucleus, but also to the output nuclei, GPi/SNr (Hazrati et al., 1990; Bolam and Smith, 1992). A re- markable consequence of this is that activation of MSN associated with either arm of the indirect path- way should increase basal ganglia output by increas- ing neuronal activity at the level of GPi/SNr: in one case, by disinhibiting the subthalamic nucleus with its excitatory projections to GPi/SNr: in the other, by di- rectly disinhibiting GPi/SNr. In contrast, activation ‘of MSNs associated with the direct pathway should decrease basal ganglia output by directly suppressing activity at the level of GPi/SNr. It has long been recognized that the large-scale or- ganization of the basal ganglia constitutes a reentrant system, with cortical influences passing through the basal ganglia and being returned to cortical levels through specific portions of the thalamus (DeLong, and Georgopoulos, 1981). Of some interest recently has been the question of how the direct and indirect pathways may differentially influence thalamocorti- cal operations. Current evidence indicates that these ‘two pathways have opposing effects on the output nu- clei. Activation of the direct pathway should result in a corresponding suppression of GABAergic GPi/SNr output, thereby disinhibiting the basal ganglia-receiv- ing portions of the thalamus, while activation of either arm of the indirect pathway should lead to an en- hancement of GPi/SNr-mediated thalamic inhibition J.Clin Neurophysiol. Vol. 1. No.4, 1904 424 G. E. ALEXANDER (Alexander and Crutcher, 1990¢). Given the reentrant nature of basal ganglia-thalamocortical connections, cortically initiated activation of the direct pathway might therefore be expected to result in positive feed~ back at cortical levels, with corresponding activation of the indirect pathway giving rise, conversely, to neg- ative feedback. A detailed understanding of how the basal ganglia’s motor circuitry operates is complicated by a variety of pathways that feed backward or forward across multiple nuclei. Two of the feedback pathways in- clude excitatory projections returned to the putamen: Already mentioned was the projection from the sub- thalamic nucleus (Parent, 1990); also noteworthy are the thalamostriatal projections that originate in the centromedian nucleus (Parent, 1983; Francois et al., 1991), Both of these pathways appear to provide pos- itive feedback to the putamen. GPe sends an inhibitory, feed-forward projection to the nucleus reticularis of the thalamus (NRT) (Haz- rati and Parent, 19915), which in turn imposes a ro- bust, GABAergic modulation upon the basal ganglia recipient nuclei of the ventrolateral thalamus (as well as upon other thalamic nuclei) (Jones, 1985). There appears to be a functional consistency among the var- ious GPe projections, in that cortically induced acti- vations of GPe-projecting MSNs should have the net effect of producing negative feedback at cortical lev- els, based on the functional effects of each of the known GPe projections—including both arms of the indirect pathway and the GPe projections to the NRT. On the other hand, the role of the NRT in basal ganglia operations may be much more complicated. It isgenerally accepted that the NRT plays an important role in gating thalamocortical transmission from the dorsal thalamus and that it does so in a roughly topo- graphic manner that links specific portions of the NRT with specific thalamic nuclei and their cortical target zones (Jones, 1985). The NRT receives collat- eral, excitatory inputs from corticothalamic as well as thalamocortical fibers that must pass through this nu- cleus en route to their respective termination zones in thalamus and cortex. In addition, the NRT may also receive collateral inputs (Jones, 1975) from both the inhibitory pallidothalamic pathway (Uno and Yos- hida, 1975; Uno et al., 1978) and the excitatory tha- lamostriate projections (Wilson et al., 1983) While the pedunculopontine nucleus (PPN) is not generally considered to be part of the basal ganglia, this structure has strong reciprocal connections with the basal ganglia output nuclei and with the subtha- J.Clin Newophysol, Vol. 11. No.4, 1994 lamic nucleus, and it also sends unreciprocated pro- Jjections to the dopaminergic neurons of the substantia nigra pars compacta (SNe; see later) as well as to the putamen and caudate nucleus (Kim et al., 1976; De- Vito et al., 1980; DeVito and Anderson, 1982; Har- nois and Filion, 1982; Carpenter and Jayaraman, 1990; Nakano et al., 1990; Smith et al., 1990; Hazrati and Parent, 199 1a). In primates the majority of palli al inputs to the PPN are apparently collaterals of pal- lidothalamic projections (Hamois and Filion, 1982; Parent, 1990; Hazrati and Parent, 19914). The PPN contains two populations of neurons, one cholinergic and the other noncholinergic (Lee et al., 1988). Itis mainly the noncholinergic neurons that project back to GPi/SNr and the subthalamic nucleus (Rye et al., 1987; Lee et al., 1988), while the cholinergic neurons project primarily to the thalamus (Hallanger et al., 1987; Hallanger and Wainer, 1988). The noncholin- ergic neurons have excitatory effects on their basal ganglia targets (Gonya-Magee and Anderson, 1983; Hammond et al., 1983; Scarnati et al., 1987), and there is some evidence that they may use glutamate as their transmitter (Scarnati etal, 1986). As with GPi/SNr, there is a convergence of inputs from both the direct and the indirect pathways at the level of the PPN. In this case, however, activation of either pathway at the level of the MSNs leads to the same polarity of response within the PPN (that is, ac- tivation of PPN by disinhibited subthalamic inputs or direct disinhibition of PPN by inputs from GPe or from GPi/SNr). Until recently, descending projections of the basal ganglia output nuclei have received relatively litte at- tention, it having been generally assumed that directly descending basal ganglia influences on the skeletomo- tor system extended no further caudally than the PPN. Recent studies in rats, however, have also re- vealed substantial PPN outflow to the reticulospinal system (Nakamura et al., 1989, 1990). These findings have not yet been extended to primates. One of the remarkable features of basal ganglia cir- cuitry is that functionally discrete channels of infor- mation processing are maintained throughout the various cortico-basal ganglia-thalamocortical path~ ways (Nambu et al, 1988, 1990, 1991) in the face of layer-to-layer connectivity that is highly convergent (Wilson and Groves, 1980; Yelnick et al., 1984). In the case of the skeletomotor pathways, for example, basal ganglia neurons have been shown to have highly refined sensorimotor fields that are comparable in their somatotopic and behavioral specificity to the sensorimotor fields of neurons at cortical levels (Geor- BASAL GANGLIA-THALAMOCORTICAL CIRCUITS IN MOVEMENT 425 gopoulos et al., 1983; Crutcher and DeLong, 1984b; DeLonget al., 1985; Alexander and Crutcher, 1990). Recent anatomical studies using transneuronal tracers have revealed at least three separate channels of arm representation that extend continuously throughout the basal ganglia-thalamocortical path- ways (Hoover and Strick, 1993). Each channel origi- nates in a separate cortical motor area (motor cortex, ventral premotor area, or supplementary motor area) and remains segregated from the others throughout the entire cortico-basal ganglia-thalamocortical loop. On the other hand, at the cellular level there is con- siderable convergence of inputs along these same pathways, Based on electron microscopic analyses, it has been estimated that the convergence ratio along the corticostriatal pathways, which provide approxi- mately half of the excitatory input to the neostriatum, ‘may be on the order of $,000:1. Cell counts in the hu- ‘man striatum and globus pallidus (Schroder et al. 1975; Thorner et al., 1975) suggest minimum con- vergence ratios on the order of 300:1 and 100:1, re- spectively, for the striatal projections to GPe and GP ‘The mechanisms that underlie the development and maintenance of functional specificity in basal ganglia networks (despite the abundant anatomical convergence) have yet to be identified. Much of this organization may be genetically determined, but ex- perience-dependent synaptic modification may also play a role. If synaptic plasticity in basal ganglia net- works operates according to Hebb-like principles (Brown et al., 1990), and there is some evidence that this may be the case (see following), the correlative nature of Hebbian learning might well be expected to strengthen preferentially those connections that link neurons with similar functional properties (Linsker, 1990). ROLE OF DOPAMINE Another important feature of basal ganglia organi- zation is the pervasive role of dopamine. The func- tionality of dopamine in basal ganglia operations ap- pears to be very complex. Dopaminergic input to the putamen consists of nigrostriatal projections that originate in the SNc (Parent, 1990), and the cortical motor and premotor areas receive separate dopamin- ergic projections from the ventral tegmental area (Gaspar et al., 1989; DeKeyser et al., 1990; Smiley et al., 1992), At the network level, dopamine appears to have differential effects on the direct and indirect pathways, tending to activate striatal MSNs that pro- ject directly to GPi while suppressing those that pro- ject to GPe (Alexander and Crutcher, 1990¢). Given the fact that there appear to be reciprocal reentrant effects associated with differential activation of the di- rect versus indirect pathways (i.e., positive feedback to cortex via the direct pathway and negative feedback via the indirect; see previously), the differential effects of dopamine on these two pathways suggest that its overall impact on basal ganglia operations may in- clude the enhancement of positive feedback and sup- pression of negative feedback, returned to the various cortical areas that receive basal ganglia influences. Dopamine has also been shown to have a role in synaptic plasticity within the striatum, being impli- cated in both long-term potentiation (LTP) and long- term depression (LTD) (Calabresi et al., 199246; Walsh, 1993). The nigrostriatal pathway provides an extraordinarily dense dopaminergic input to each MSN, which is comparable in magnitude to the 5,000 oso corticostriatal synapses that individual MSNs re- ceive (Doucet et al., 1986). The behavioral correlates of dopamine neurons have been studied in depth by Schultz. and colleagues (Romo and Schultz, 1990; Schultz and Romo, 1990). Unlike nearly all the other basal ganglia motor circuit neurons, except perhaps the large, aspiny interneurons of the striatum (Ki- mura, 1986; Kimura et al., 1990), dopamine neurons do not show activity changes in relation to movement. per se, but discharge instead in relation to conditions involving the probability and imminence of behav- ioral reinforcement (Romo and Schultz, 1990; Schultz and Romo, 1990). ‘The newly discovered facts that dopamine may play a crucial role in the cellular mechanisms of LTD and LTP within the striatum and that dopamine neurons discharge sclectively in relation to specific reward con- tingencies combine to suggest the intriguing hypothe- sis that dopamine neurons may play an important role in determining when striatal synapses should be strengthened or weakened. In this respect, dopamine neurons might be seen as playing a role in striatal in- formation processing analogous to that of an “adap- tive critic” in connectionist networks. It is noteworthy, however, that the striatum is not the only basal ganglia structure for which there is evi- dence of adaptive synapses. With their combined volt- age dependency and ligand specificity, N-methyl-D- aspartate (NMDA) receptors are widely assumed to play a role in at least one form of activity-dependent synaptic plasticity, viz., LTP (Rauschecker, 1991). Evidence for the existence of NMDA receptors has been found in a number of basal ganglia nuclei that are believed to receive significant glutamatergic input, J.Clin, Newophysiol, Vol 1. No.4, 198 426 G. E, ALEXANDER including the striatum (putamen and caudate nu- cleus) (Cherubini et al., 1988; Young et al., 1990; Ka- ‘waguchi, 1992), the subthalamic nucleus (Nakanishi etal., 1988; Young etal., 1990), and the SNe (Mercuri etal., 1992). NMDA receptors have also been demon- strated at cortical levels (Young et al., 1990; Arm- strong et al., 1993; Kobayashi et al., 1993), where the phenomenon of LTP has been documented in many regions including some of the sensorimotor fields (Baranyi and Szente, 1987; Bindman et al., 1988; Iriki et al., 1989, 1991), Taken together, these findings in- dicate that adaptive synapses may be distributed at multiple junctures along the basal ganglia-thalamo- cortical pathways. LIMITS OF OUR CURRENT UNDERSTANDING OF BASAL GANGLIA FUNCTION Disinhibition plays an important role in many cur- rent models of basal ganglia operation (Chevalier and Deniau, 1990). In the case of the basal ganglia’s ocu- lomotor pathways, for example, there is compelling evidence that striatally induced, phasic inhibition of SNr neurons leads to the generation of saccadic eye movements through the release of command neurons in the superior colliculus from the tonic, GABAergic inhibition they receive via the nigrotectal pathway (Hikosaka and Wurtz, 1983, 1985; Hikosaka et al., 1989). The elegant simplicity of this model has helped to motivate similar models of how the basal ganglia's skeletomotor circuitry may contribute to movement control, the chief difference being the emphasis in most skeletomotor models on disinhibition at the level of the thalamus rather than brainstem (Alexan- der and Crutcher, 19900). These models of the basal ganglia’s skeletomotor circuitry assume that voluntary movements are facili- tated in the context of focused disinhibition of the basal ganglia-recipient portions of the ventrolateral thalamus. This focused thalamic disinhibition is thought to be generated either by phasic enhancement of transmission through the direct pathway, by phasic suppression of transmission through the indirect pathway, or by a combination of these two processes. Conversely, according to this same scheme, decreased transmission through the direct pathway, or increased transmission through the indirect, would have the effect of suppressing voluntary movements. This relatively simplistic model raises questions of its own, however. It is not known, for example, whether a given corticostriatal neuron (or functional J.Clin.Newrophysol. Vo. 11. No.4, 1994 group of such neurons) engages both the direct and the indirect pathways in a balanced manner. Nor is it known whether the convergence of inputs from the direct and indirect pathways onto individual GPi/SNr neurons (Bolam and Smith, 1992) results in a func tional interaction between the two pathways that is antagonistic or complementary. A possibility at one extreme would be a functionally antagonistic, push/ pull system that could be used to scale and/or brake the intended movement (e.g., with flexion-activated inputs from the two pathways being superimposed and thereby tending to cancel one another out). At the other extreme would be a functionally complemen- tary, center-surround system that might serve to facil- itate the intended movement pattern while suppress- ing potentially conflicting ones. Whether cither of these alternative schemes applies, or whether instead there may be a purely random convergence of direct and indirect projections onto GPi/SNr neurons, has important implications for any functional model of basal ganglia organization Studies of various primate models of basal ganglia— induced movement disorders have tended to confirm many of the predictions of the thalamic disinhibition models (DeLong, 1990). These models predict that excessive basal ganglia outflow should be associated with hypokinetic states such as are seen in the various akinetic/rigid disorders of movement control. Be- cause of dopamine’s reciprocal effects on the direct and indirect pathways, experimental striatal dopa- mine deficiency should and does result in excessive discharge of neurons in the subthalamic nucleus as well as in GPi/SNr. In primates with parkinsonian akinesia induced experimentally by treatment with MPTP, the discharge rates of GPi neurons as well as of neurons in the subthalamic nucleus have been found to be abnormally high, consistent with predic- tions of the thalamic disinhibition models (Filion et al., 1988; Miller and DeLong, 1988). Voluntary movements can be restored in these akinetic animals by selective lesion of the subthalamic nucleus, which results in the return of GPi neurons to more normal rates of discharge (Bergman et al., 1990). Also consis- tent with thalamic disinhibition modelsis the fact that akinesia in humans with idiopathic Parkinson's dis- ease can be relieved by selective stereotaxic lesions of GPi (especially when the lesions are centered in that nucleus’ sensorimotor territory) (Laitinen et al., 1992) Such models are also able to account for hyperki- netic disorders (e.g., involuntary ballistic or chorei- form movements) on the basis of reduced basal gan- BASAL GANGLIA-THALAMOCORTICAL CIRC! glia outflow (DeLong, 1990). Such reduced outflow is ‘thought to result in excessive disinhibition of the basal ganglia-receiving territories within the thalamus, with consequent release of involuntary movements, This explanation is consistent with various clinical and experimental data, including the fact that lesions of the subthalamic nucleus result in dyskinesias (Car- penteretal., 1950) and that such lesionsare associated with reduced levels of spontaneous discharge in GPi (Bergman et al., 1990), ‘There are, however, a number of observations that are difficult to reconcile with simple models of basal ganglia function based primarily on the concept of thalamic disinhibition. One of the most serious such conflicts involves the fact that lesions of the basal gan- alia output nuclei do not result in dyskinesias (Horak and Anderson, 1984; Mink and Thach, 1991) as the models predict they should, although dyskinesias have been reported in association with muscimol duced suppression of GPi output (Kato and Kimura, 1992). A possible explanation for this discrepancy may be that the voltage-ependent, bistable properties of thalamic neurons (Steriade and Linas, 1988) may ‘cause them to respond to downward modulations of inhibitory basal ganglia input in a nonmonotonic manner. On the one hand, moderate reductions of GABAergic input from the basal ganglia, as would be expected with subthalamic lesions, may partially, but inappropriately, depolarize some thalamic relay neu- rons to the point that they enter a nonrelay burst ‘mode when they are supposed to be maximally hyper- polarized, leading to bursts of inappropriate thalamo- cortical activity and resultant dyskinesias. On the other, severe reductions in GABAergic input, as would be expected with GPi lesions, may actually re- store some stability by increasing the level of depolar- ization of all basal ganglia-recipient thalamic relay cells to the point where they enter into a more or less permanent, nonbursting, relay mode. Or perhaps the reason that complete removal of inhibitory basal gan- glia input to the thalamus is not associated with dys- kinesias is that the emergence of dyskinesias may re- quire that reduced tonic levels of thalamic inhibition be combined with the preservation of at least some degree of phasic modulation of these same inhibitory inputs (see Steriade and Llinas, 1988). A related difficulty for thalamic disinhibition theo- riesis that lesions of the ventrolateral thalamus do not result in akinesia, and yet according to such theories parkinsonian akinesia is generally attributed to in- creased basal ganglia outflow that results in excessive inhibition at the level of the ventrolateral thalamus 5 IN MOVEMENT 47 (DeLong, 1990). One possible explanation for this ap- parent discrepancy is that we may have underesti- ‘mated the functional significance of descending basal ganglia outflow to the PPN (and from there to the re- ticulospinal system), in which case lesions of the thal- amus that block only the reentrant (cortically pro- jected) influences of the basal ganglia may leave intact those descending influences that are conveyed more directly to the segmental motor apparatus. Further inconsistencies and unanswered questions continue to surface, making it difficult to accept any of the current theories of basal ganglia function as de- finitive. For example, why should striatal dopamine deficiency lead to severe akinesia whereas structural lesions of the striatum fail to do so? And if basal gan- glia dysfunction can lead to akinesia at one extreme and involuntary movements at the other, which sug- gests that the basal ganglia may play some role in the generation of movement, why is it that movement related neuronal activity in the basal ganglia begins relatively late with respect to movement onset, well after the beginnings of movement-related activity in cortical motor areas or in the cerebellum (Thach, 1978; Anderson and Horak, 1985; Crutcher and Al- exander, 1990)? And, perhaps most disconcerting of all, what sort of role can the basal ganglia be playing, in normal motor control, and how significant might that role be, if lesions of the output nuclei result only in minor disturbances of motor output (Horak and Anderson, 1984; Mink and Thach, 1991)? SUMMARY AND CONCLUSIONS Enormous progress has been made recently in char- acterizing the structure and the functional organiza- tion of the basal ganglia. We now know which parts of the basal ganglia are involved in movement control, how movement parameters are encoded in the dis- charge patterns of basal ganglia neurons, how the basal ganglia communicate with cortical and other subcortical structures involved in movement control, how abnormal activity within the basal ganglia can lead to a variety of movement disorders, and how therapeutic manipulations of these nuclei can restore motor function in certain types of basal ganglia disor- ders. ‘At the same time, much has yet to be learned. Al- though the anatomy and physiology—and to some extent even the pathophysiology—of these structures are now understood in some detail, we still do not un- derstand precisely what functions the basal ganglia subserve in the course of their normal operations. In J.C, Newophysial, Vol. 11. No.4 1994 428 G. particular, we have yet to learn what specific contri- butions the basal ganglia may make to the generation and coordination of voluntary movements. (Of course, it should also be noted that the same can be said of virtually every other brain structure now known to be involved in skeletomotor control.) Based on the impressive pace of recent discoveries concern- ing the functional organization of the basal ganglia, there is reason to hope that new insights into the basal ganglia's role in movement control may soon be forthcoming. Acknowledgment: This work was supported by grants NS- 17678 from the National Institutes of Health and NOOO 4- 92-J-11132 from the Office of Naval Research, REFERENCES Alexander GE, Crutcher MD. Preparation for movement: neural representations of intended direction in three motor areas of the monkey. J Newophvsiol 19904:64:133-50, [Alexander GE, Crutcher MD. Neural representations of the target (oa) of visually guided arm movements in three motor areas ofthe monkey. Neurophysiol 19906; 64:164-78 Alexander GE, Cruichet MD. Functional architecture of basal gan- lia circuits: neural substrates of parallel processing. Trends Neurosei 1990e, 13:266-71 Alexander GE, Crutcher MD, DeLong MR. Basel ganglia-thalamo. ‘cortical circuits: parallel substrates for motor, oculomotor, “prefrontal” and “limb” functions. Prog Brain Res 1980;85: 119-46, ‘Alexander GE, DeLong MR. Microstimlation of the primate neo- striatum. Il Somatotopic omanization of striatal microexcita- ble zones and their relation to neuronal response properties. J Neurophysiol 1985a:53:1417-30 Alexander GE, DeLong MR. Microstimulation ofthe primate neo- ‘striatum J, Physiological properties of striatal microexctable zones. J Newophysiol 985b;53:1401-16, [Anderson ME, Horak FB, lluence of the plobus pallidus on arm Tovements ia monkeys I, Timing of movement-relatedin- formation, J Neuroph)sio!1985;54:433-48, Armstrong JM, Welker E, Callahan CA. The contribution of NMDA and non-NMDA receptors to fist and slow transmis sion of sensory information in the rat SI barrel cortex. J New rose 1993; 13:2149-60 Baranyi A, Szente MB. Long-lasting potentiation of synaptic trans- ‘mission requires postsynaptic modifications in the neocortex. Brain Res 987;423378-84 Bauswein E, Fromm C, Preuss A. Cortcosriatal cells in compat “on with pyramidal tract neurons: contrasting properties in the behaving monkey. Brain Res 1989;493:198-203, Bergman H. Wichmann T, DeLong MR, Reversal of experimental parkinsonism by lesions of the subthalamic nucleus. Science 1990:249:1936-8, Bindman LJ, Murphy KPSI, Pockett S. Postsynaptic control of the induction of long-term changes in efficacy Of transmission at reocortical synapses in slices of rat brain. J’ Newophysiol 1988; 1053:1065. Bolam JP, Smith Y. Thestriatum and the globus pallidus send con- ‘vergent synaptic inputs onto single cellsin the entopeduncular nucleus ofthe rat: a double anterograde labelling study com bined with postembedding immunocytechemisiry for GABA. FComp Neurol 1992;321:456-76, Bolam JP, Wainer BH, Smith AD. Characterization of cholinergic 2. lin, Nevrophysol, Vol, No.4, 1904 ALEXANDE: rneutons in rt striatum. A combination ofcholine aetylrans- ferase immunocytochemistry, Golgi impregnation and clec- tron microscopy: Neuroscience 1984; 12:71 1-8. Brown TH, Kainss EW, Keenan CL. Hebbian synapses: biophysi- ‘al mechanisms and algorithms, Annu Rev Neurasci 1990;13: 475-311 CCalabresiP, Maj R, Pisani A, Mercuri NB, Bernardi G. Long-term synaptic depression inthe striatum: physiological and pharma- ological characterization J Newosci (992a,12:4224-33, Calabres P, Pisani A, Mercuri NB, Bernardi G. Long-term potenti- ‘ation in the striatum is unmasked by removing the voltage- ‘dependent magnesium block of NMDA receptor channels. Eur J Neurosei 1992.4:929-35, Carpenter MB, Jayaraman A. Subthalamic nucleus ofthe monkey: “connections and immunocylochemical features of afferents, J Hirnforsch 1990;31°653-68, Carpenter MB, Nakano K, Kim R. Nigrothalamic projections in the monkey demonstrated by autoradiographic technics. J Comp Neurol 1976;165:401-16 Carpenter MB, Whittier JR, Metter FA. Analysis of choreoid hy: perkinesain the rhesus monkey: surgical and pharmacological Stnalysis of hyperkinesia resulting from lesions in the subtha- Jamie nucleus of Luys. J Comp Neural 1950;92:293-332. ‘Cavada C, Goldan-Rakie PS. Topographic segregation ofcortico- “riatal projections from posterior parietal subdivisions in the macaque monkey. Neuroscience 1991;42:683-96, Cherubini E, Herring PL, Lanfomey L, Stanzione P. Excitatory ‘amino acids in symapiic excitation of rat striatal neurones in vitro. J Physiol (Lond) 1988;400:677-90. Chevalier G, Deniau JM, Disinhibition as a basic process in the expression of striatal functions. Trends Neurosci 1990; 13:277= 50. Crutcher MD, Alexander GE. Movement-elated neuronal activity ‘selectively coding either direction or muscle pattern in three motor areas of the monkey. J Neurophysiol 1990;64:151-63, ‘Crutcher MD, DeLong MR. Single cell studies ofthe primate putae ‘men. 1. Functional organization, Exp Brain Res 1984a:53: 233-83 Crutcher MD, DeLong MR. Single cell studies of the primate puta- ‘en. I, Relations to direction of movement and pattem of muscular activity, Exp Brain Res 19840; 53,244-58. Dekeyser J, Herregodks P, Ebinger G. The mesoneocortcal dopa- inine neuron system. Newroogy 1990:40:1660-2. DeLong MR. Primate models of movement disorders of basal gan- lis origin, Trends Neurosci 1990; 13:281-5. DeLong MR, Crutcher MD, Georgopoulos AP. Relations between ‘movernent and single cell discharge inthe substantia nigra of the behaving monkey. J Newrose! 1983:3:1599-606. DeLong MR, Crutcher MD, Georgopoulos AP. Primate globus pal- Tidusand subthalamic nucleus: functional organization. J New rophysiol 1985;53:530—43. DeLong MR, Georgopoulos AP. Motor functions ofthe basal gan tla, In: Brookhart JM, Mountcasle VB, Brooks VB, Geiger SR, eds. The nervous s)sterr: motor contro. Bethesda Ameri- ‘can Physiological Society, 1981:1017-61. (Handbook of phys- iology: sect 1, vol 2, pt2) DeVito JL, Anderson ME. An autoradiographic study of efferent connections of the globus pallidus. Exp Brain Res 1982:46: 107-17 DeVito JL, Anderson ME, Walsh KE, A horseradish peroxidase ‘ody of afferent connections ofthe globus pallidus in Macaca ‘mulasa. Exp Brain Res \980;38:65-73 Difilia M, Rafols JA. Synaptic organization of the globus pallidus, “J Eleciron Microsc Tech 1988; 10247-63. Donoghve JP, Kitai ST. A collateral pathway to the neostriatam from corticafugal neurons ofthe rat sensory-motor cortex: an intracellular HRP study. J Comp Neurol 1981;20:1~13. Doucet G, Descartes L, Garcia S. Quantification of the dopamine BASAL GANGLIA-THALAMOCORTICAL CIRCUITS IN MOVEMENT 429 mnervation in adult rat neostiatum, Newroscience 1986; 19: a27-45, Dum RP, Sirick PL. The origin of corticospinal projections from remotorareasin the frontal lobe. J Newrasci 1991; 1 667-89. Filion M, Tremblay L, Bedard PJ. Abnormal influences of passive lim movement on the activity of globus pallidus neurons in parkinsonian monkeys, Brain Res 1988: 444:165-76, Faherty AW, Graybiel AM, Corticostiatal transformations in the primate somatosensory system. Projections from physiologi- tally mapped body-part representations, J Neurophysiol 1991;66:1249-63, Fisherty AW, Graybiel AM, Output architecture ofthe primate pu amen. J Newroset1993a;13°3222-37, Flaherty AW, Graybiel AM. Two input systems for body represen- tations inthe primate striatal matrix: experimental evidence in the squirrel monkey. J Neurosci 19936; 13:1 120-27 Fox CA, Rafols JA. The stata effernts inthe globus pallidus and inthe subsiantia nigra, I: Yahr MD, ed. The basal ganglia, New York: Raven Press, 1976:37-55. Francois , Percheron G, Parent A, Sadikot AF, Fenelon G, Yelnik J. Topography ofthe projection from the central complex of the thalamus to the sensorimotor striatal leritary in monkeys Comp Neurol 1991;305:17-34, Gaspar P, Berger B, Febvtet A, Vigay A, Henry JP. Catecholamine innervation of the human cerebral cortex as revealed by com- parative immunohistochemistry of tyrosine hydroxylase and dopamine-beta-hydroxylase. J Comp Neural 1989;279:249- 1 Georgopoulos AP, DeLong MR, Crutcher MD. Relations between parameters of step-tracking, movements and single cell dis- harge inthe globus pallidus and subthalamic nucleus of the behaving monkey. J Neurosci 1983,3:1586-98. ‘Gonya-Magee T, Anderson ME. An clecrophysiologicalcharacter- ization of projections from the pedunculopontine area to en= topeduncular nucleus and globus palidusin the cat. Exp fain Res 1983:49:269-79, Hallanger AE, Levey Al, Lee HJ, Rye DB, Wainer BH. The origins of cholinergic and other “nonspecific” afferents to the thala- ‘usin the rat. J Comp Neurol 1987;262:105-24, Hallanger AE, Wainer BH. Ascending projections from the pedun- ‘culopontine tegmental nucleus and the adjacent mesopontine tegmentum in the rat. J Comp Newol 1988;274:483-S18, Hamada I, DeLong MR, Mano N-I. Activity of identified wrist- related pallidal neurons during step and ramp wrist move- ‘mentsin the monkey. J Neurophysiol 1990;64:1892-906, Hammond C, Rouzaire-Dubois B, oper J Jackson A, Crossman A. ‘Anatomical and cletrophysiologcal studies on the reciprocal projections between the subthalamic nucleus and nucleus te ‘menti pendunculopontinus in the rat. Neurascience 1983: $1232, ‘Harnois , Filion M. Palidofugal projections to thalamus and mid- brain: a quantitative antidromic activation study in monkeys and eats. Exp Brain Res 1982;47277-85. ‘Harimann-von Monakow K, Akert K, Kinzle H. Projections of the precentral motor cortex and other cortical area ofthe frontal lobe to the subthalamic nucleusin the monkey. Exp Brain Res 1978;33:395-203, Hazrati L-N, Parent A. Contralateral pallidothalamic and palido- tegmental projections in primates: an anterograde and retro- ‘rade labeling study. Brain Res 19914; 567:212-23 ‘Hazrati LN, Parent A. Projection from the external pallidum tothe reticular thalamic nucleus inthe squirrel monkey. Brain Res 1991); 580:142-6, Hazrati LIN, Parent A, Mitchell 8, Haber SN. Evidence for inter= ‘connections between the two Sepments ofthe globus pallidus in primates: a PHA-L anterograde tracing study. Brain Res 1990;333:171-5. Hikosaka 0, Sakamoto M, Usui 8. Functional properties of mon- key caudate neurons. I. Activities related to saccadic eye move: ‘ments. J Neurophysiol 1989;61:780-98, Hikosaka 0, Wurtz RH. Visual and oculomotor functions of mon- ‘key substantia nigra pars reticulata. IV. Relation of substantia nigra to superior colliculus. J Newrophysiol 1983,49:1285- Sor Hikosaka 0, Wurtz RH. Modification of saccadic eye movements bby GABA-telated substances. [, Effet of muscimol and bicu- callin in monkey superior coliculus.J Neurophysiol 1985;53 266-91 Hoover JE, Strick PL. Multiple output channelsin the basal ganglia, Sctence 1993;259:819-21 Horak FB, Anderson ME. Influence of globus pallidus on arm "movements in monkeys |. Effects of kainic-induced lesions. J Neurophysiol 984;52:290-304, Hummelsheim H, Wiesendanger M, Bianchetti M, Wiesendanger R, Macpherson J. Further investigations ofthe eflerent linkage of the supplementary motor area (SMA) with the spinal cord fof the monkey. Exp Brain Res 1986;65:75-82, Hutchins KD, Martino AM, Strick PL. Corticospinal projections rom the’ medial wall of the hemisphere. Exp Brain Res 1988,715:667-72, Wiki A, Pavlides C, Keller A, Asanuma H. Long-term potentiation in the motor cortex. Science 1989:245:1385-7, Iriki A, Pavlides C. Keller A, Asanuma H. Long-term potentiation of thalamic input to the motor cortex induced by coactivation fof thalamocortical and corticacortcal afferents. J Newrophy'- {io 1991;65:1435-41 Jones EG. Some aspects ofthe organization of the thalamic reticular ‘complex. J Comp Newol 1975: 162:285-308, Jones EG. The thalamus, New York: Plenum Pres, 1985. Jones EG, Coulter 1D, Burton H, Porter R. Cells of origin and ter ‘minal distribution of comicosriatal fibers arising in the senso- sy-motor cortex of monkeys. J Comp Neurol 1977;173:53-80, Kato M, Kimura M. Effects of reversible blockade of basal ganglia ‘ona voluntary arm movement, J Neurophysiol 1992;68:15 16 34, Kawaguchi Y. Large agpiny cells in the matrix ofthe rat neostia- ‘um in vitro: physiological identification, relation to the com- partments and excitatory postsynaptic currents. J Newrophy To 1992567: 1669-82. Kawaguchi ¥, Wilson Cl, Emson PC. Intracellular recording of ‘dentiied neostratal patch and matrix spiny cells in a slice preparation preserving cortical inputs. J Neurophysiol 1989;62:1082-68, kim R, Nakano K, Jayaraman A, Carpenter MB, Projections of the slobus pallidus and adjacent structures: an autoradiographic study inthe monkey. J Comp Newol 1976; 169:263-0. Kimura M. The role of primate putamen neuronsiin the association ‘of sensory stimuli with movement, Neurosci Res 1986;3:436- 3 Kimura M, Aostki T, Hu Y, Ishida A, Watanabe K. Activity of ‘primate putamen neuronsisselective tothe mode of voluntary ‘movement: visually guided, selFinitated or memory guided. Exp Brain Res 1992:89:073-7. Kimura M, Kato M, Shimazaki H. Physiological properties of roj- ection neurons in the monkey striatum tothe globus pallidus. Exp Brain Res 1990:82:672-6. Kita H, Kitai ST. Efferent projections ofthe subthalamic nucleus in the rat: light and cleetton microscopic analysis with the PHA-L method. J Comp Neurol 1987;260:435~52. Kita H, Kitai ST. Glutamate decarboxylase immunoreactive neu- soasin rat neosriatum: their morphological types and popula- tions. Brain Res 1988;447:346-5 Kitai ST, Deniau JM. Cortical inputs to the subthalamus: an intra- cellular analysis. Bran Res 1981;214:41 1-5, Kobayashi T, Nagao T, Fukuda H, Hicks TP, Oka JIL NMDA re- ceptors mediate neuronal burst firing in at somatosensory cor- texin vivo, Newroreport 1993;4:735-8, J.Clin Newophysiol. Vo. 1, No.4, 1994 430 G. E, ALEXANDER Kiinzle H. Bilateral projections from precentral motor cortex tothe putamen and other parts of the basal ganglia. An autoradio~ raphie study in Macaca fasciculars. Brain Res 1975;88:195: 208. Knale H. Projections from the primary somatosensory cortex to ‘basil ganglia and thalamus in the monkey. Exp Brain Res 1977;30:481-92. Kinale H. An autoradiographic analysis ofthe efferent connections from premotor and adjacent prefrontal regions (areas 6 and 9) in Macaca fasicularis. Brain Behav Evol \978:15:183-234 Laitinen LV, Bergenheim AT. Hariz MI, Leksell's posteroventral pallidotomy in the treatment of Parkinson's disease. J Neuro Surg 1992: 16-53-61. Lee HJ, Rye DB, Hallanger AE, Levey Al, Wainer BH. Cholinersic 's-noncholinerpc efferent from the mesopontine tegmentum to the extrapyramidal motor system nuclei. J Comp Newol 1988;275:469-92, Liles SL. Activity of neurons in putamen during active and passive "movements of wrist J Newophysiol 1985;53:217-36. Linsker R. Perceptual neural organization: some approaches based ‘on network models and information theory. Anni Rev New ‘rosci 1990; 13257-81 Mercuri NB, Strata F, Calabresi P, Berardi G. A voltage-clamp, ‘analysis of NMDA-induced responses on dopaminergic neu rons of the rat substantia nigra 2ona compacta and ventral teg- mental area. Brain Res 1992;593:51-6. Miller WC, DeLong MR. Parkinsonian symptomatology: an ana- {omical and physiological analysis. Ann NY Acad Sei 1988;515:287-302. Mink JW, Thach W', Basal ganalia motor contro. Il Pallidal ab Tation: normal reaction time, muscle cocontraction, and slow movement. J Neurophysiol 1991:65:530-S1. Mitchell SJ, Richardson RT, Baker FH, DeLong MR. The primate plobus pallidus: neuronal acvity related to direction of move- ment. Exp Brain Res 1987;68:491 ~505, [Nakamura Y, Kudo M, Tokuno H. Monosynaptic projection from the pedunculopontine tegmental region to the reticulospinal feutons of the medulla oblongata: an electron microscopic study in the cat. Brain Res 1990;524:353-6. Nakamura ¥. Tokuno H, Moriizumi T, Kitao Y. Kudo M. Mono~ Synaptic nigral inputs to the pedunculopontine tegmental n- ‘leus neurons which send their axons to the medial reticular formation in the medulla oblongata. An electron microscopic study in the eat, Neurosci Lett 1989; 103:145-80. [Nakanishi H, Kita H, Kita ST. An Nomethyl-o-aspartate receptor imediated excitatory postsynaptic potential evoked in subtha- Tamic neurons in an io vitro slice preparation of the rat. New sei Lett 1988;95:130-4, Nakano K, Hasegawa Y, Tokushige A, Nakagawa S, Kayahara T, ‘Mizuno N. Topographicel projections from the thalamus, sub- thalamic nucleus and pedunculopontine tegmental nucleus to the stratum in the Japanese monkey, Macaca fuscata, Brain ‘Res 1990;537:54-68 Nambu A, Yoshida S, Finnai K, Projection on the motor cortex of thalamic neurons with pallial mputin the monkey. Exp Brain Res 1988;71:658-62. Nambu A, Yoshida S,Jinnai K. Discharge patterns of pallial neu Tons with input ftom various cortical areas during movement inthe monkey, Brain Res 1990;519:183-91 Nambu A. Yoshida S,Jinnai K. Movementrelated activity of thal- ‘amie neurons with input from the globus pallidus and proj- ction to the motor cortex of the monkey. Exp Brain Res 1991;84:279-84 Nauta HIW. A proposed conceptual reorganization of the basal fzanglia and telencephalon, Neuroscience 1979:4: 1875-81 Parent A. The subcortical afferents to caudate nucleus and putamen im primate: a fluorescence retrograde double labeling study. Neuroscience 1983;10:1137-50. J.Clin Neurophysiol, Vol 11 No.4 1994 Parent A. Comparative neurabiology of the basal ganglia. New ‘York: Wiles, 1985, arent A. Extrinsic connections ofthe basal ganglia. Trends New ase! 1990; 13:254-8, Penney JBJe, Young AB, GABA as the pallidothalamie neurotrans- milter: implications for besal ganglia function. Brain Res 1981;207:195-9. Phelps PE, Houser CR, Vaughn JE. fmmunorytochemicallocaliza- Tion of choline acesltransferase within the rat neostriatum: & correlated light and electron microscopic study of cholinergic ‘neurons and synapses. J Comp Neurol 1985;238:286~307. Rauschecher JP. Mechanisms of visual plasticity: Hebb synapses, NMDA receptors, and beyond, Physiol Rev 1991;71:587~615. Robledo P, Fes J Excitatory influence of rat subthalamic nucleas tothe substantia nigra pars reticulata and the pallidal complex: tlectrophysiological data, Brain Res 1990;518:47—54, ‘Romo R, Schultz W. Dopamine neurons of the monkey midbrain: ‘Sondingencics of responses to active touch during slF-initated farm movements. J Neurophysiol 1990;63:592-606. ye DB, Saper CB, Lee HJ, Wainer BH. Pedunculopontine ter- ‘ental nucleus ofthe rat: cytoarchitecture, eytochemistry and Some extrapyramidal connections ofthe mesopontine tegmen= tum. J Comp Neurol 1981;259:483-528. ‘ScarnattE, Prioa A, Campana E, Pacitti C. A mieroiontophoretic ‘aud on the nature of the putative synaptic neurotransmitter in the pedunculopontine-substantia agra pars compacta excit- ‘tory pathvay inthe rat. Exp Brain Res 1986;62:470-8. SScarnatiE, Prioa A, DeLoreto S, Patt C. The reciprocal electto- physiological influence between the nucleus tezmenti pedun- Eulopontinus and the substantia nigra in normal and decort- tated rats Brain Res 1987;423:116-24 Schroder KE, Hopf A, Lange H, Thorner G. Morphemetrisch-sat- che Strukturenalysen des Stratum, Pallidom und Nucleus Sabihalamicus beim Menschen, 1 Striatum, J £fimforsch 1975; 16:333-50. Schultz W, Romo R. Dopamine neurons of the monkey midbrain: ‘contingencies of responses to stimuli eliciting immediate be- hhavioral reactions. J Newophysiol 1990;63:607-24. Selemon LD, Goldman-RakicPS. Longitudinal topography and in- {erdigitation of cortco-stratal projections inthe rhesus mon ey. d Neuroscl 1985;5:776-94, Selemon LD, Goldman-Rakic PS. Topographical intermingling of “Seiatonigral and striatopallidal ncurons in the rhesus monkey 3 Comp Newrol 1990;297359-16, ‘Smiley JF, Williams SM, Srigeti K, Goldman-Rakie PS. Light and ‘letron microscopic characterization of dopamine-immmuno- feactive axons inhuman cerebral cortex. J Comp Neurol 1992;321:325-35. ‘Smith ¥, Haztati L-N, Parent A. Efferent projections ofthe subtha- amie nucleus in the squirrel monkey as studied by PHHA-Lan- terograde tracing method. J Comp Neurol 1990;294:306-23. ‘Smith Y; Parent A. Neurons ofthe subthalamic nucteus in primates “display slutamate but not GABA immunoreactivity. Brain Res 1988;453:353-0 Steriade M, Llinas RR. The functional sates ofthe thalamus and the associated neuronal interplay. Physio! Rev 1988;68:649 742. “Thach WT. Correlation of neural discharge with pattern and force ‘of muscular activity, joint postion, and direction of intended ‘peat movement in motor cortex and cerebellum, J Newrophys- {ol L97B;41:654-76, “Thomer G, Lange H, Hopf A. Morphometrisch-statische Struktur- Tinalysen des Striatum, Pallidum und Nucleus Subthalamicus beim Menschen. Il Palidum, J Hirnforsch 1975:16:401-13. Ueki A, Uno M, Anderson A, Yoshida M. Monosynaptic inhibition ‘of thalamic neurons produced by stimulation of the substantia pigra. Experientia 1977; 38 1480-2. Uno, Ozawa N, Yoshida M. The mode of pllido-thalamic trans- BASAL GANGLIA-THALAMOCORTICAL CIRCUITS IN MOVEMENT 431 mission investigated with intracellular recording from cat thal amus. Exp Brain Res 1978;33:483-507 ‘Uno M, Yoshida M, Monosynaptic inhibition of thalamic neurons ‘produced by stimulation ofthe pallial nucteus in cat. Brain Res 1995,99:377-80. Vatschanoff 1G, Weinberg RJ, Rustioni A. Amino acid immunoreac- tivity in corticospinal terminals. Exp Brain Res 1993:9395~103. Walsh JP. Depression of excitatory synaptic input in rat striatal ‘neurons. Brain Res 1993;608:123-8. Wilson Cl. Postsynaptic potentials evoked in spiny neostriatal proj- fection neurons by stimulation of ipsilateral and contralateral neocortex. Bra; Res 1986;367:201-13. Wilson CI, Chang HT, Kita ST. Disfaciitation and long-lasting inhibition of neostriatal neurons in the rat, Exp Brain Res 1983,51:227-35. Wilson CJ, Chang HT, Kitai ST. Firing patterns and synaptic po- Tentals of identified giant aspiny inthe rat neostriatum. J New ‘Wilson CJ, Groves PM, Fine structure and synaptic connections ‘of the common spiny neuron of the rat neostiatum: a study ‘employing intracellular injection of horseradish peroxidase. J ‘Comp Neurol 1980; 194:599-615. ‘Yamamoto T, Samejima A, Oka H. An intracellular analysis of the ‘entopeduncolar inputs on the centrum medianum-parafasci- cular nuclear complex in cats. Brain Res 1985;348:343-T. ‘J, Percheron G, Francois. A Golgi analysis ofthe primate elobus pallidus. 1. Quantitative morphology and spatial orien- tation of dendritic arborizations. J Comp Neurol 1984;227: 200-13, Yeterian EH, Pandya DN. Stital connections ofthe parietal asso- Cation cortices in chesus monkeys. J Comp Neurol 1993;332: 175-97 ‘Young AB, Dauth GW, Hollingsworth Z, Penney JB, Kaatz K, Gil ‘man 8, Quisqualate- and NMDA-sensitive (3HJglutamate binding in primate brain. J Neurose! Res 1990;27:512-21 Yeln J.Clin,Nearophysol, Vol. 1, No.4, 1984

You might also like