Comprehensive Physiology Wiley Online Library

Basal Ganglia—A Motion Perspective

Full Article on Wiley Online Library



Abstract

The basal ganglia represent an ancient part of the nervous system that have remained organized in a similar way over the last 500 million years and are of importance for our ability to determine which actions to choose at any given moment in time. Salient or reward stimuli act via the dopamine system and contribute to motor or procedural learning (reinforcement learning). The input stage of the basal ganglia, the striatum, is shaped by glutamatergic input from the cortex and thalamus and by the dopamine system. All intrinsic neurons of the striatum are GABAergic and inhibitory except for the cholinergic interneurons. Too little dopamine and all vertebrates show symptoms similar to that of a Parkinsonian patient, whereas too much dopamine results in hyperkinesia with involuntary movements. In this article, we discuss the detailed organization of the basal ganglia, with the different cell types, their properties, and contributions to basal ganglia functions. The striatal projection neurons represent 95% of all neurons in the striatum and are subdivided into two types, one that projects directly to the output stage, referred to as the “direct” pathway that promotes action, and the other subtype that targets the output nuclei via intercalated basal ganglia nuclei. This “indirect” pathway has an opposite effect. The striatal projection neurons express a set of ion channels that give them a high threshold for activation, whereas neurons in all other parts of the basal ganglia have a resting discharge that allows for modulation in both an increased and decreased direction. © 2020 American Physiological Society. Compr Physiol 10:1241‐1275, 2020.

Figure 1. Figure 1. Common motor infrastructure from lamprey to man. Throughout the vertebrates, several basic motor behaviors are controlled by neuronal networks (CPGs) located in the brainstem and spinal cord. The basal ganglia play a crucial role in the selection of motor behaviors and are similarly organized in lamprey and primate. In primates, the addition of a well‐developed cerebral cortex provides a locus for networks controlling fine motor skills.
Figure 2. Figure 2. The basal ganglia subnuclei in the human brain. (A) The location of the different basal ganglia subnuclei at the level of thalamus. (B) A sagittal view of the brain showing the shape of the caudate‐putamen. (C) Schematic of the striatum indicating the matrix and striosome compartments.
Figure 3. Figure 3. The organization of the basal ganglia. The striatum consists of GABAergic neurons, as do GPe, GPi, and SNr. SNr and GPi represent the output level of the basal ganglia, and it projects via different subpopulations of neurons to the superior colliculus (SC), the mesencephalic (MLR), and diencephalic (DLR) locomotor command regions and other brainstem motor centers, as well as back to thalamus with efference copies of information sent to the brainstem. The dSPNs that target SNr/GPi express the dopamine D1 receptor (D1) and substance P (SP), while the iSPNs express the dopamine D2 receptor (D2) and enkephalin (Enk). The indirect loop is represented by the GPe, the STN, and the output level (SNr/GPi)—the net effect being an enhancement of activity in these nuclei. Also indicated is the dopamine input from the SNc (green) to striatum and brainstem centers. Excitatory glutamatergic neurons are shown in pink and GABAergic structures in blue color.
Figure 4. Figure 4. Striatal interneurons and the striatal microcircuit. (A) Each subtype of striatal interneurons identified by their neurotransmitter expression (inner circle), other molecular markers (middle circle), and electrophysiological properties (outer circle) are represented in the circular plot. Redrawn and modified, with permission, from Burke DA, et al., 2017 50, Figure 1. (B) The striatal microcircuit with the connectivity between the striatal projection neurons (SPNs) and their input from FS, LTS and ChIN interneurons. Abbreviations: ACh, acetylcholine; ChAT, choline‐acetyl transferase; ChIN, cholinergic interneuron; CR, calretinin; FA, fast adapting; FS, fast spiking; GABA, gamma‐butyric acid; 5‐HT3R, serotonin type‐3 receptor; LTS, low‐threshold spiking; NOS, nitric oxide synthase; NPY, neuropeptide Y; PV, parvalbumin; SOM, somatostatin; TAN, tonically active neurons; TH, tyrosine hydroxylase.
Figure 5. Figure 5. Input to different neuronal subpopulations in striatum. (A) Many cortical/pallial axons that target the brainstem and spinal cord (PT‐type) give off collaterals to neurons within the striatum. There is a subset of pyramidal neurons that have intratelencephalic axons projecting to the contralateral cortex/pallium (IT‐type) that also target the striatum. (B) Cortical and thalamic neurons target both direct and indirect striatal projection neurons (d/iSPNs) and the ChINs, FS, and LTS interneurons. The glutamatergic pedunculopontine (PPN) neurons only project to the interneurons, whereas the cholinergic PPN target the d/iSPNs. The red dashed arrow from cortex to ChINs indicates a variable and weak effect.
Figure 6. Figure 6. The direct, indirect, and hyperdirect pathways. Striatal projection neurons of the direct pathway (dSPNs) directly target the output level (SNr) and will enhance the excitability of brainstem motor targets through disinhibition and thus promote action. SPNs of the indirect pathway (iSPNs) will inhibit the spontaneously active GPe that in turn disinhibit SNr, thus increasing inhibition of downstream motor targets. The hyperdirect pathway projects to the glutamatergic STN that in turn targets SNr that will then inhibit the motor targets.
Figure 7. Figure 7. Connectivity of the globus pallidus externa (GPe) and the subthalamic nucleus (STN) with target structures. The GPe has two subpopulations of GABAergic cells, prototypical and arkypallidal cells. The prototypical cells receive input from iSPNs and STN. They project to the STN and GPi/SNr. The arkypallidal cells project back to the striatum's dSPNs, iSPNs, and interneurons, and receive input from the STN, cortex, and dSPNs. The STN receives input from the cortex, thalamus, PPN, SNc, and GPe. Like the GPe, the STN projects to the output nuclei GPi/SNr.
Figure 8. Figure 8. The activity of striatal projection neurons of the direct and indirect pathway during a goal‐directed push‐pull task. (A) Shows the activity pattern (spiking frequency) of a dSPN during a push/pull task. The red trace demonstrates a correct response (reward), and the blue trace an incorrect response (no reward). Upon the GO signal, the neuron is activated and remains active until a sound signals if the response will lead to a reward or not. The actual reward occurs with a further delay. Note that after the reward signal, the level of activity remains high, whereas with no reward the activity drops immediately. (B) The corresponding data for an indirect pathway neuron (iSPN). Note that immediately after the GO signal there is a marked increase of activity that rapidly decreases, while after the no‐reward signal there is a marked increase from base‐line. (C) Simplified scheme of the basal ganglia. Two separate populations of dSPNs and iSPNs control the push and the pull motion, respectively. The action is mediated by the basal ganglia output nuclei SNr and GPi to downstream motor circuits. Reused, with permission, from Grillner S, 2018 140.
Figure 9. Figure 9. The effects of enhanced or decreased dopamine activity on the direct and indirect pathways through the basal ganglia. (A) An enhanced dopamine activity excites the striatal projection neurons of the direct pathway that express dopamine receptors of the D1 subtype, while it inhibits those of the indirect pathway through their D2 receptors. (B) Illustrates the opposite situation with decreased dopamine activity that removes excitation from the direct pathway and reduces the inhibition of the indirect pathway and thereby indirectly increase the net excitation.
Figure 10. Figure 10. Parallel pathways for goal‐directed behavior conveyed via the head of the caudate nucleus (CDh) and habitual behavior produced through the tail of the caudate nucleus (CDt). The CDt and CDh receive input from different cortical regions and both target the superior colliculus to elicit saccadic eye movements, although through separate channels and via separate output neurons of the basal ganglia in Substantia Nigra pars reticulata (SNr). Moreover, separate parts of the Substantia Nigra pars compacta (SNc) supply the two circuits. Abbreviations: CDh, head of the caudate nucleus; CDt, tail of the caudate nucleus; c‐d‐l, caudal‐dorsal‐lateral; GPe(c), caudal part of GPe; GPe(r), rostral part of GPe; r‐v‐m, rostral‐ventral‐medial; SC, superior colliculus. Modified and redrawn, with permission, from Kim HF and Hikosaka O, 2015 183, Figure 6.
Figure 11. Figure 11. Schematic representation of the effect of an increased salient dopamine (DA) burst on striatal cell populations controlling different motor acts. When active, different cortical/thalamic cell populations (blue color) control the excitability of separate populations of striatal neurons, which are depolarized but not firing at the resting level of dopamine (2nd panel from the left). With enhanced DA activity, however, the striatal population becomes activated (3rd panel) and promotes motor action. The 4th panel shows the same condition for another population of striatal cells.
Figure 12. Figure 12. A sequence of tightly coupled discrete movements combined into one integrated movement, termed chunking. Illustration designed as in Figure 11, but with the addition of the salient dopamine burst that accompanies each discrete movement (see text).
Figure 13. Figure 13. Schematic representation of the excitatory globus pallidus projecting to the lateral habenulae (GPh; red), and the inhibitory compartment (GPi; blue) in lamprey, rodent, and primates. GPh and GPi are represented in separate nuclei in lamprey but are merged into two compartments within one nucleus (entopeduncular nucleus) in rodents. In primates, the habenular projecting parts are located mostly at the periphery, also referred to as the border region of the GPi. Modified, with permission, from Stephenson‐Jones M, et al., 2013 317, Figure 7.
Figure 14. Figure 14. Overview of the basal ganglia/habenular circuits underlying the control of motion and evaluation. The lower motion circuit corresponds to the circuits detailed in Figure 3. The dSPNs in the matrix compartment target GPi and SNr, as well as brainstem motor centers and send a collateral back to the thalamus (the lower part of the diagram). Also indicated is the indirect pathway via the GPe and STN. The evaluation circuit, in the upper part of the diagram, shows the lateral habenula (LHb) that target the dopamine (DA) neurons both directly and indirectly via the GABAergic rostromedial tegmental nucleus (RMTg). The LHb receives input from the glutamatergic habenula‐projecting globus pallidus (GPh). The GPh receives excitation from cortex and thalamus, whereas it receives inhibition from iSPNs in the striosome compartment. DA neurons are inhibited by striosomal dSPNs and send projections to the mesencephalic locomotor region (MLR) and optic tectum. The color code is blue for GABAergic, red for glutamatergic, and green for dopaminergic neurons.
Figure 15. Figure 15. Receptor‐induced signaling activating downstream effectors important for long‐term potentiation (LTP) or depression (LTD). Activation of different receptors activates different protein kinases involved in LTP, whereas other sets of receptors have other downstream targets and evoke LTD. There is evidence that eliciting of LTP processes inhibit LTD, and vice versa. CaMKII, Ca2+/calmodulin‐dependent protein kinase II; PKA, protein kinase A; PKC, protein kinase C; PKG, protein kinase G; NO, nitric oxide; 2AG, 2‐arachidonoylglycerol (an endocannabinoid); ERK, extracellular signal‐regulated kinases; sGC, soluble guanylyl cyclase; mGluR, metabotropic glutamate receptor; A2AR, adenosine 2A receptor.
Figure 16. Figure 16. The characteristic motor symptoms of Parkinson's disease. Front and side views of a man portrayed with Parkinson's disease. These are woodcut reproductions. From Paul de Saint‐Leger's 1879 doctoral thesis, Paralysie agitante. Fig. 145) published by Gowers 128, p. 591.
Figure 17. Figure 17. Data‐driven detailed simulation of striatum. (A) Shows the somata of the number of cells (71,242) contained in 1 mm3, distributed according to estimated densities. (B) Shows a limited number of cells with their somatodendritic arbor, based on detailed reconstruction of SPNs of both types, ChINs, and FS and LTS interneurons. The high density of dendritic and axonal branches within the tissue is evident. For each type of cell, the detailed membrane properties are simulated and validated versus their biological counterparts. (C) Simulated cells with dendrites are shown with different magnification from left to right. A touch detection algorithm is used to detect where axonal and dendritic structures come close (red circles as indicated). Depending on the pre‐ and postsynaptic cell types, adjusted pruning rules are applied and validated against established connectivity. In this way, the neuronal microstructure is reconstructed bottom‐up in a data‐driven manner. Also, optimization algorithms are used to fit electrophysiological properties to the different neuronal types. These data‐driven workflows have also been used to predict cortical microcircuits 151,225.
Figure 18. Figure 18. Phylogenetic tree of vertebrates. The lamprey diverged from the line to mammals around 560 million years ago (mya). Adapted, with permission, from Reiner A, et al., 1998 289.
Figure 19. Figure 19. From lamprey to primates—the organization of the basal ganglia is almost identical throughout vertebrate phylogeny. (A) The striatum consists of GABAergic neurons (blue), as do the GPe, GPi, and SNr. The output level of the basal ganglia is represented by the GPi/SNr, and it projects via different subpopulations of neurons to the optic tectum (superior colliculus in mammals), the mesencephalic (MLR) and diencephalic (DLR) locomotor command regions and other brainstem motor centers, and send additional efference copies of the information back to thalamus. The indirect loop is represented by the GPe, the STN, and the output level (SNr/GPi)—the net effect is an enhancement of activity in these nuclei. The dSPNs of the direct pathway express the dopamine D1 receptor and substance P (SP), while the iSPNs express the dopamine D2 receptor and enkephalin (Enk). Excitatory glutamatergic neurons are represented by pink, GABAergic structures in blue and dopamine input from the SNc in green. (B) A table showing the key features of the basal ganglia organization that are found in mammals and lamprey. So far, subtypes of FS striatal interneurons have not been demonstrated in the lamprey.
Figure 20. Figure 20. The SNc connectome in lamprey and mammals. The efferent and afferent connectivity of the SNc is virtually identical in lamprey and mammals. Thus, the dopaminergic neurons within SNc project to the same structures in the basal ganglia as in mammals and the same midbrain motor centers. The input to SNc is similarly identical from the striatum, STN, cortex/pallium, PPN, and the lateral habenulae. Reused, with permission, from Perez‐Fernandez J, et al., 2014 271.
Figure 21. Figure 21. Scheme summarizing the main building‐blocks promoting and inhibiting action in the forebrain. Cortical circuits can directly and via the direct pathway promote action, as can thalamic circuits and the dopamine system. The cerebral cortex can also inhibit action via the STN, the hyperdirect pathway, and via the striatum and the indirect pathway and finally via GPe inhibition of striatum. For simplicity the output nuclei of the basal ganglia are not included—only the net effect of the direct and indirect pathway.


Figure 1. Common motor infrastructure from lamprey to man. Throughout the vertebrates, several basic motor behaviors are controlled by neuronal networks (CPGs) located in the brainstem and spinal cord. The basal ganglia play a crucial role in the selection of motor behaviors and are similarly organized in lamprey and primate. In primates, the addition of a well‐developed cerebral cortex provides a locus for networks controlling fine motor skills.


Figure 2. The basal ganglia subnuclei in the human brain. (A) The location of the different basal ganglia subnuclei at the level of thalamus. (B) A sagittal view of the brain showing the shape of the caudate‐putamen. (C) Schematic of the striatum indicating the matrix and striosome compartments.


Figure 3. The organization of the basal ganglia. The striatum consists of GABAergic neurons, as do GPe, GPi, and SNr. SNr and GPi represent the output level of the basal ganglia, and it projects via different subpopulations of neurons to the superior colliculus (SC), the mesencephalic (MLR), and diencephalic (DLR) locomotor command regions and other brainstem motor centers, as well as back to thalamus with efference copies of information sent to the brainstem. The dSPNs that target SNr/GPi express the dopamine D1 receptor (D1) and substance P (SP), while the iSPNs express the dopamine D2 receptor (D2) and enkephalin (Enk). The indirect loop is represented by the GPe, the STN, and the output level (SNr/GPi)—the net effect being an enhancement of activity in these nuclei. Also indicated is the dopamine input from the SNc (green) to striatum and brainstem centers. Excitatory glutamatergic neurons are shown in pink and GABAergic structures in blue color.


Figure 4. Striatal interneurons and the striatal microcircuit. (A) Each subtype of striatal interneurons identified by their neurotransmitter expression (inner circle), other molecular markers (middle circle), and electrophysiological properties (outer circle) are represented in the circular plot. Redrawn and modified, with permission, from Burke DA, et al., 2017 50, Figure 1. (B) The striatal microcircuit with the connectivity between the striatal projection neurons (SPNs) and their input from FS, LTS and ChIN interneurons. Abbreviations: ACh, acetylcholine; ChAT, choline‐acetyl transferase; ChIN, cholinergic interneuron; CR, calretinin; FA, fast adapting; FS, fast spiking; GABA, gamma‐butyric acid; 5‐HT3R, serotonin type‐3 receptor; LTS, low‐threshold spiking; NOS, nitric oxide synthase; NPY, neuropeptide Y; PV, parvalbumin; SOM, somatostatin; TAN, tonically active neurons; TH, tyrosine hydroxylase.


Figure 5. Input to different neuronal subpopulations in striatum. (A) Many cortical/pallial axons that target the brainstem and spinal cord (PT‐type) give off collaterals to neurons within the striatum. There is a subset of pyramidal neurons that have intratelencephalic axons projecting to the contralateral cortex/pallium (IT‐type) that also target the striatum. (B) Cortical and thalamic neurons target both direct and indirect striatal projection neurons (d/iSPNs) and the ChINs, FS, and LTS interneurons. The glutamatergic pedunculopontine (PPN) neurons only project to the interneurons, whereas the cholinergic PPN target the d/iSPNs. The red dashed arrow from cortex to ChINs indicates a variable and weak effect.


Figure 6. The direct, indirect, and hyperdirect pathways. Striatal projection neurons of the direct pathway (dSPNs) directly target the output level (SNr) and will enhance the excitability of brainstem motor targets through disinhibition and thus promote action. SPNs of the indirect pathway (iSPNs) will inhibit the spontaneously active GPe that in turn disinhibit SNr, thus increasing inhibition of downstream motor targets. The hyperdirect pathway projects to the glutamatergic STN that in turn targets SNr that will then inhibit the motor targets.


Figure 7. Connectivity of the globus pallidus externa (GPe) and the subthalamic nucleus (STN) with target structures. The GPe has two subpopulations of GABAergic cells, prototypical and arkypallidal cells. The prototypical cells receive input from iSPNs and STN. They project to the STN and GPi/SNr. The arkypallidal cells project back to the striatum's dSPNs, iSPNs, and interneurons, and receive input from the STN, cortex, and dSPNs. The STN receives input from the cortex, thalamus, PPN, SNc, and GPe. Like the GPe, the STN projects to the output nuclei GPi/SNr.


Figure 8. The activity of striatal projection neurons of the direct and indirect pathway during a goal‐directed push‐pull task. (A) Shows the activity pattern (spiking frequency) of a dSPN during a push/pull task. The red trace demonstrates a correct response (reward), and the blue trace an incorrect response (no reward). Upon the GO signal, the neuron is activated and remains active until a sound signals if the response will lead to a reward or not. The actual reward occurs with a further delay. Note that after the reward signal, the level of activity remains high, whereas with no reward the activity drops immediately. (B) The corresponding data for an indirect pathway neuron (iSPN). Note that immediately after the GO signal there is a marked increase of activity that rapidly decreases, while after the no‐reward signal there is a marked increase from base‐line. (C) Simplified scheme of the basal ganglia. Two separate populations of dSPNs and iSPNs control the push and the pull motion, respectively. The action is mediated by the basal ganglia output nuclei SNr and GPi to downstream motor circuits. Reused, with permission, from Grillner S, 2018 140.


Figure 9. The effects of enhanced or decreased dopamine activity on the direct and indirect pathways through the basal ganglia. (A) An enhanced dopamine activity excites the striatal projection neurons of the direct pathway that express dopamine receptors of the D1 subtype, while it inhibits those of the indirect pathway through their D2 receptors. (B) Illustrates the opposite situation with decreased dopamine activity that removes excitation from the direct pathway and reduces the inhibition of the indirect pathway and thereby indirectly increase the net excitation.


Figure 10. Parallel pathways for goal‐directed behavior conveyed via the head of the caudate nucleus (CDh) and habitual behavior produced through the tail of the caudate nucleus (CDt). The CDt and CDh receive input from different cortical regions and both target the superior colliculus to elicit saccadic eye movements, although through separate channels and via separate output neurons of the basal ganglia in Substantia Nigra pars reticulata (SNr). Moreover, separate parts of the Substantia Nigra pars compacta (SNc) supply the two circuits. Abbreviations: CDh, head of the caudate nucleus; CDt, tail of the caudate nucleus; c‐d‐l, caudal‐dorsal‐lateral; GPe(c), caudal part of GPe; GPe(r), rostral part of GPe; r‐v‐m, rostral‐ventral‐medial; SC, superior colliculus. Modified and redrawn, with permission, from Kim HF and Hikosaka O, 2015 183, Figure 6.


Figure 11. Schematic representation of the effect of an increased salient dopamine (DA) burst on striatal cell populations controlling different motor acts. When active, different cortical/thalamic cell populations (blue color) control the excitability of separate populations of striatal neurons, which are depolarized but not firing at the resting level of dopamine (2nd panel from the left). With enhanced DA activity, however, the striatal population becomes activated (3rd panel) and promotes motor action. The 4th panel shows the same condition for another population of striatal cells.


Figure 12. A sequence of tightly coupled discrete movements combined into one integrated movement, termed chunking. Illustration designed as in Figure 11, but with the addition of the salient dopamine burst that accompanies each discrete movement (see text).


Figure 13. Schematic representation of the excitatory globus pallidus projecting to the lateral habenulae (GPh; red), and the inhibitory compartment (GPi; blue) in lamprey, rodent, and primates. GPh and GPi are represented in separate nuclei in lamprey but are merged into two compartments within one nucleus (entopeduncular nucleus) in rodents. In primates, the habenular projecting parts are located mostly at the periphery, also referred to as the border region of the GPi. Modified, with permission, from Stephenson‐Jones M, et al., 2013 317, Figure 7.


Figure 14. Overview of the basal ganglia/habenular circuits underlying the control of motion and evaluation. The lower motion circuit corresponds to the circuits detailed in Figure 3. The dSPNs in the matrix compartment target GPi and SNr, as well as brainstem motor centers and send a collateral back to the thalamus (the lower part of the diagram). Also indicated is the indirect pathway via the GPe and STN. The evaluation circuit, in the upper part of the diagram, shows the lateral habenula (LHb) that target the dopamine (DA) neurons both directly and indirectly via the GABAergic rostromedial tegmental nucleus (RMTg). The LHb receives input from the glutamatergic habenula‐projecting globus pallidus (GPh). The GPh receives excitation from cortex and thalamus, whereas it receives inhibition from iSPNs in the striosome compartment. DA neurons are inhibited by striosomal dSPNs and send projections to the mesencephalic locomotor region (MLR) and optic tectum. The color code is blue for GABAergic, red for glutamatergic, and green for dopaminergic neurons.


Figure 15. Receptor‐induced signaling activating downstream effectors important for long‐term potentiation (LTP) or depression (LTD). Activation of different receptors activates different protein kinases involved in LTP, whereas other sets of receptors have other downstream targets and evoke LTD. There is evidence that eliciting of LTP processes inhibit LTD, and vice versa. CaMKII, Ca2+/calmodulin‐dependent protein kinase II; PKA, protein kinase A; PKC, protein kinase C; PKG, protein kinase G; NO, nitric oxide; 2AG, 2‐arachidonoylglycerol (an endocannabinoid); ERK, extracellular signal‐regulated kinases; sGC, soluble guanylyl cyclase; mGluR, metabotropic glutamate receptor; A2AR, adenosine 2A receptor.


Figure 16. The characteristic motor symptoms of Parkinson's disease. Front and side views of a man portrayed with Parkinson's disease. These are woodcut reproductions. From Paul de Saint‐Leger's 1879 doctoral thesis, Paralysie agitante. Fig. 145) published by Gowers 128, p. 591.


Figure 17. Data‐driven detailed simulation of striatum. (A) Shows the somata of the number of cells (71,242) contained in 1 mm3, distributed according to estimated densities. (B) Shows a limited number of cells with their somatodendritic arbor, based on detailed reconstruction of SPNs of both types, ChINs, and FS and LTS interneurons. The high density of dendritic and axonal branches within the tissue is evident. For each type of cell, the detailed membrane properties are simulated and validated versus their biological counterparts. (C) Simulated cells with dendrites are shown with different magnification from left to right. A touch detection algorithm is used to detect where axonal and dendritic structures come close (red circles as indicated). Depending on the pre‐ and postsynaptic cell types, adjusted pruning rules are applied and validated against established connectivity. In this way, the neuronal microstructure is reconstructed bottom‐up in a data‐driven manner. Also, optimization algorithms are used to fit electrophysiological properties to the different neuronal types. These data‐driven workflows have also been used to predict cortical microcircuits 151,225.


Figure 18. Phylogenetic tree of vertebrates. The lamprey diverged from the line to mammals around 560 million years ago (mya). Adapted, with permission, from Reiner A, et al., 1998 289.


Figure 19. From lamprey to primates—the organization of the basal ganglia is almost identical throughout vertebrate phylogeny. (A) The striatum consists of GABAergic neurons (blue), as do the GPe, GPi, and SNr. The output level of the basal ganglia is represented by the GPi/SNr, and it projects via different subpopulations of neurons to the optic tectum (superior colliculus in mammals), the mesencephalic (MLR) and diencephalic (DLR) locomotor command regions and other brainstem motor centers, and send additional efference copies of the information back to thalamus. The indirect loop is represented by the GPe, the STN, and the output level (SNr/GPi)—the net effect is an enhancement of activity in these nuclei. The dSPNs of the direct pathway express the dopamine D1 receptor and substance P (SP), while the iSPNs express the dopamine D2 receptor and enkephalin (Enk). Excitatory glutamatergic neurons are represented by pink, GABAergic structures in blue and dopamine input from the SNc in green. (B) A table showing the key features of the basal ganglia organization that are found in mammals and lamprey. So far, subtypes of FS striatal interneurons have not been demonstrated in the lamprey.


Figure 20. The SNc connectome in lamprey and mammals. The efferent and afferent connectivity of the SNc is virtually identical in lamprey and mammals. Thus, the dopaminergic neurons within SNc project to the same structures in the basal ganglia as in mammals and the same midbrain motor centers. The input to SNc is similarly identical from the striatum, STN, cortex/pallium, PPN, and the lateral habenulae. Reused, with permission, from Perez‐Fernandez J, et al., 2014 271.


Figure 21. Scheme summarizing the main building‐blocks promoting and inhibiting action in the forebrain. Cortical circuits can directly and via the direct pathway promote action, as can thalamic circuits and the dopamine system. The cerebral cortex can also inhibit action via the STN, the hyperdirect pathway, and via the striatum and the indirect pathway and finally via GPe inhibition of striatum. For simplicity the output nuclei of the basal ganglia are not included—only the net effect of the direct and indirect pathway.
References
 1.Abdi A, Mallet N, Mohamed FY, Sharott A, Dodson PD, Nakamura KC, Suri S, Avery SV, Larvin JT, Garas FN, Garas SN, Vinciati F, Morin S, Bezard E, Baufreton J, Magill PJ. Prototypic and arkypallidal neurons in the dopamine‐intact external globus pallidus. J Neurosci 35: 6667‐6688, 2015.
 2.Abecassis ZA, Berceau BL, Win PH, Garcia D, Xenias HS, Cui Q, Pamucku A, Cherian S, Hernández VM, Chon U, Lim BK, Justice NJ, Awatramani R, Kim Y, Hooks BM, Gerfen CR, Boca SM, Chan CS. Npas1+ – Nkx2.1+ neurons are an integral part of the cortico‐pallido‐cortical loop. J Neurosci 40: 743‐768, 2020.
 3.Alexander GE, DeLong MR. Microstimulation of the primate neostriatum. I. Physiological properties of striatal microexcitable zones. J Neurophysiol 53: 1401‐1416, 1985.
 4.Alexander GE, DeLong MR. Microstimulation of the primate neostriatum. II. Somatotopic organization of striatal microexcitable zones and their relation to neuronal response properties. J Neurophysiol 53: 1417‐1430, 1985.
 5.Alexander GE, DeLong MR, Strick PL. Parallel organization of functionally segregated circuits linking basal ganglia and cortex. Annu Rev Neurosci 9: 357‐381, 1986.
 6.Alloway KD, Smith JB, Mowery TM, Watson GDR. Sensory processing in the dorsolateral striatum: The contribution of thalamostriatal pathways. Front Syst Neurosci 11: 53, 2017.
 7.Amemori S, Amemori KI, Yoshida T, Papageorgiou GK, Xu R, Shimazu H, Desimone R, Graybiel AM. Microstimulation of primate neocortex targeting striosomes induces negative decision‐making. Eur J Neurosci 51: 731‐741, 2020.
 8.Ammari R, Lopez C, Bioulac B, Garcia L, Hammond C. Subthalamic nucleus evokes similar long lasting glutamatergic excitations in pallidal, entopeduncular and nigral neurons in the basal ganglia slice. Neuroscience 166: 808‐818, 2010.
 9.An X, Mohan H, Matho K, Kepecs A, Huang J. A cortical command circuit coordinates food handling and manipulation. Society for Neuroscience Meeting abstract number 493.13 https://www.abstractsonline.com/pp8/#!/7883/presentation/49433, 2019
 10.Aosaki T, Kimura M, Graybiel AM. Temporal and spatial characteristics of tonically active neurons of the primate's striatum. J Neurophysiol 73: 1234‐1252, 1995.
 11.Apicella AJ, Wickersham IR, Seung HS, Shepherd GM. Laminarly orthogonal excitation of fast‐spiking and low‐threshold‐spiking interneurons in mouse motor cortex. J Neurosci 32: 7021‐7033, 2012.
 12.Apicella P, Legallet E, Trouche E. Responses of tonically discharging neurons in the monkey striatum to primary rewards delivered during different behavioral states. Exp Brain Res 116: 456‐466, 1997.
 13.Arber S, Costa RM. Connecting neuronal circuits for movement. Science 360: 1403‐1404, 2018.
 14.Assous M, Dautan D, Tepper JM, Mena‐Segovia J. Pedunculopontine glutamatergic neurons provide a novel source of feedforward inhibition in the striatum by selectively targeting interneurons. J Neurosci 39: 4727‐4737, 2019.
 15.Assous M, Kaminer J, Shah F, Garg A, Koos T, Tepper JM. Differential processing of thalamic information via distinct striatal interneuron circuits. Nat Commun 8: 15860, 2017.
 16.Assous M, Tepper JM. Excitatory extrinsic afferents to striatal interneurons and interactions with striatal microcircuitry. Eur J Neurosci 49: 593‐603, 2019.
 17.Atallah HE, McCool AD, Howe MW, Graybiel AM. Neurons in the ventral striatum exhibit cell‐type‐specific representations of outcome during learning. Neuron 82: 1145‐1156, 2014.
 18.Baker PM, Jhou T, Li B, Matsumoto M, Mizumori SJ, Stephenson‐Jones M, Vicentic A. The lateral habenula circuitry: Reward processing and cognitive control. J Neurosci 36: 11482‐11488, 2016.
 19.Ballion B, Mallet N, Bezard E, Lanciego JL, Gonon F. Intratelencephalic corticostriatal neurons equally excite striatonigral and striatopallidal neurons and their discharge activity is selectively reduced in experimental parkinsonism. Eur J Neurosci 27: 2313‐2321, 2008.
 20.Baufreton J, Atherton JF, Surmeier DJ, Bevan MD. Enhancement of excitatory synaptic integration by GABAergic inhibition in the subthalamic nucleus. J Neurosci 25: 8505‐8517, 2005.
 21.Baufreton J, Bevan MD. D2‐like dopamine receptor‐mediated modulation of activity‐dependent plasticity at GABAergic synapses in the subthalamic nucleus. J Physiol 586: 2121‐2142, 2008.
 22.Baufreton J, Kirkham E, Atherton JF, Menard A, Magill PJ, Bolam JP, Bevan MD. Sparse but selective and potent synaptic transmission from the globus pallidus to the subthalamic nucleus. J Neurophysiol 102: 532‐545, 2009.
 23.Belic JJ, Kumar A, Hellgren Kotaleski J. Interplay between periodic stimulation and GABAergic inhibition in striatal network oscillations. PLoS One 12: e0175135, 2017.
 24.Benabid AL, Chabardes S, Mitrofanis J, Pollak P. Deep brain stimulation of the subthalamic nucleus for the treatment of Parkinson's disease. Lancet Neurol 8: 67‐81, 2009.
 25.Benabid AL, Chabardes S, Torres N, Piallat B, Krack P, Fraix V, Pollak P. Functional neurosurgery for movement disorders: A historical perspective. Prog Brain Res 175: 379‐391, 2009.
 26.Bergman H, Wichmann T, DeLong MR. Reversal of experimental parkinsonism by lesions of the subthalamic nucleus. Science 249: 1436‐1438, 1990.
 27.Bergman H, Wichmann T, Karmon B, DeLong MR. The primate subthalamic nucleus. II. Neuronal activity in the MPTP model of parkinsonism. J Neurophysiol 72: 507‐520, 1994.
 28.Berke JD. What does dopamine mean? Nat Neurosci 21: 787‐793, 2018.
 29.Berthet P, Hellgren‐Kotaleski J, Lansner A. Action selection performance of a reconfigurable basal ganglia inspired model with Hebbian‐Bayesian Go‐NoGo connectivity. Front Behav Neurosci 6: 65, 2012.
 30.Berthet P, Lindahl M, Tully PJ, Hellgren‐Kotaleski J, Lansner A. Functional relevance of different basal ganglia pathways investigated in a spiking model with reward dependent plasticity. Front Neural Circuits 10: 53, 2016.
 31.Betolngar DB, Mota E, Fabritius A, Nielsen J, Hougaard C, Christoffersen CT, Yang J, Kehler J, Griesbeck O, Castro LRV, Vincent P. Phosphodiesterase 1 bridges glutamate inputs with NO‐ and dopamine‐induced cyclic nucleotide signals in the striatum. Cereb Cortex 29: 5022‐5036, 2019.
 32.Bevan MD. The subthalamic nucleus. In: Steiner H, Tseng KY, editors. Handbook of Basal Ganglia Structure and Function. 2nd edition, London, Oxford, San Diego, Cambridge MA: Academic Press, 2017, p. 277‐291.
 33.Bevan MD, Bolam JP, Crossman AR. Convergent synaptic input from the neostriatum and the subthalamus onto identified nigrothalamic neurons in the rat. Eur J Neurosci 6: 320‐334, 1994.
 34.Bevan MD, Francis CM, Bolam JP. The glutamate‐enriched cortical and thalamic input to neurons in the subthalamic nucleus of the rat: Convergence with GABA‐positive terminals. J Comp Neurol 361: 491‐511, 1995.
 35.Bjursten LM, Norrsell K, Norrsell U. Behavioural repertory of cats without cerebral cortex from infancy. Exp Brain Res 25: 115‐130, 1976.
 36.Blackwell KT, Salinas AG, Tewatia P, English B, Hellgren Kotaleski J, Lovinger DM. Molecular mechanisms underlying striatal synaptic plasticity: Relevance to chronic alcohol consumption and seeking. Eur J Neurosci 49: 768‐783, 2019.
 37.Bobadilla AC, Heinsbroek JA, Gipson CD, Griffin WC, Fowler CD, Kenny PJ, Kalivas PW. Corticostriatal plasticity, neuronal ensembles, and regulation of drug‐seeking behavior. Prog Brain Res 235: 93‐112, 2017.
 38.Bolam JP, Ellender TJ. Histamine and the striatum. Neuropharmacology 106: 74‐84, 2016.
 39.Bolam JP, Pissadaki EK. Living on the edge with too many mouths to feed: Why dopamine neurons die. Mov Disord 27: 1478‐1483, 2012.
 40.Borgkvist A, Lieberman OJ, Sulzer D. Synaptic plasticity may underlie l‐DOPA induced dyskinesia. Curr Opin Neurobiol 48: 71‐78, 2018.
 41.Boyd LA, Edwards JD, Siengsukon CS, Vidoni ED, Wessel BD, Linsdell MA. Motor sequence chunking is impaired by basal ganglia stroke. Neurobiol Learn Mem 92: 35‐44, 2009.
 42.Brandalise F, Carta S, Helmchen F, Lisman J, Gerber U. Dendritic NMDA spikes are necessary for timing‐dependent associative LTP in CA3 pyramidal cells. Nat Commun 7: 13480, 2016.
 43.Brimblecombe KR, Cragg SJ. The striosome and matrix compartments of the striatum: A path through the labyrinth from neurochemistry toward function. ACS Chem Nerosci 8: 235‐242, 2017.
 44.Brissaud É. Leçons sur les maladies nerveuses. Paris: Masson, 1925.
 45.Brodin L, Hokfelt T, Grillner S, Panula P. Distribution of histaminergic neurons in the brain of the lamprey Lampetra fluviatilis as revealed by histamine‐immunohistochemistry. J Comp Neurol 292: 435‐442, 1990.
 46.Brodin L, Theordorsson E, Christenson J, Cullheim S, Hokfelt T, Brown JC, Buchan A, Panula P, Verhofstad AA, Goldstein M. Neurotensin‐like peptides in the CNS of lampreys: Chromatographic characterization and immunohistochemical localization with reference to aminergic markers. Eur J Neurosci 2: 1095‐1109, 1990.
 47.Bromberg‐Martin ES, Hikosaka O. Lateral habenula neurons signal errors in the prediction of reward information. Nat Neurosci 14: 1209‐1216, 2011.
 48.Bromberg‐Martin ES, Matsumoto M, Hikosaka O. Dopamine in motivational control: Rewarding, aversive, and alerting. Neuron 68: 815‐834, 2010.
 49.Bruce NJ, Narzi D, Trpevski D, van Keulen SC, Nair AG, Rothlisberger U, Wade RC, Carloni P, Hellgren Kotaleski J. Regulation of adenylyl cyclase 5 in striatal neurons confers the ability to detect coincident neuromodulatory signals. PLoS Comput Biol 15: e1007382, 2019.
 50.Burke DA, Rotstein HG, Alvarez VA. Striatal local circuitry: A new framework for lateral inhibition. Neuron 96: 267‐284, 2017.
 51.Caboche J, Vanhoutte P, Boussicault L, Betuing S. Cellular and molecular mechanisms of neuronal dysfunction in Huntington's disease. In: Steiner H, Tseng KY, editors. Handbook of Basal Ganglia Structure and Function. 2nd edition, London, Oxford, San Diego, Cambridge MA: Academic Press, 2017, p. 889‐906.
 52.Cachope R, Mateo Y, Mathur BN, Irving J, Wang HL, Morales M, Lovinger DM, Cheer JF. Selective activation of cholinergic interneurons enhances accumbal phasic dopamine release: Setting the tone for reward processing. Cell Rep 2: 33‐41, 2012.
 53.Caligiore D, Pezzulo G, Baldassarre G, Bostan AC, Strick PL, Doya K, Helmich RC, Dirkx M, Houk J, Jorntell H, Lago‐Rodriguez A, Galea JM, Miall RC, Popa T, Kishore A, Verschure PF, Zucca R, Herreros I. Consensus paper: Towards a systems‐level view of cerebellar function: The interplay between cerebellum, basal ganglia, and cortex. Cerebellum 16: 203‐229, 2017.
 54.Carlsson A. Evidence for a role of dopamine in extrapyramidal functions. Acta Neuroveg (Wien) 26: 484‐493, 1964.
 55.Carlsson A. A paradigm shift in brain research. Science 294: 1021‐1024, 2001.
 56.Carlsson A, Lindqvist M, Magnusson T, Waldeck B. On the presence of 3‐hydroxytyramine in brain. Science 127: 471, 1958.
 57.Cenci MA. Molecular mechanisms of l‐DOPA‐induced dyskinesia. In: Steiner H, Tseng KY, editors. Handbook of Basal Ganglia Structure and Function. 2nd edition, London, Oxford, San Diego, Cambridge MA: Academic Press, 2017, p. 857‐871.
 58.Chakravarthy VS, Joseph D, Bapi RS. What do the basal ganglia do? A modeling perspective. Biol Cybern 103: 237‐253, 2010.
 59.Chen CH, Fremont R, Arteaga‐Bracho EE, Khodakhah K. Short latency cerebellar modulation of the basal ganglia. Nat Neurosci 17: 1767‐1775, 2014.
 60.Chen MC, Ferrari L, Sacchet MD, Foland‐Ross LC, Qiu MH, Gotlib IH, Fuller PM, Arrigoni E, Lu J. Identification of a direct GABAergic pallidocortical pathway in rodents. Eur J Neurosci 41: 748‐759, 2015.
 61.Chergui K, Charlety PJ, Akaoka H, Saunier CF, Brunet JL, Buda M, Svensson TH, Chouvet G. Tonic activation of NMDA receptors causes spontaneous burst discharge of rat midbrain dopamine neurons in vivo. Eur J Neurosci 5: 137‐144, 1993.
 62.Chiken S, Nambu A. Mechanism of deep brain stimulation: Inhibition, excitation, or disruption? Neuroscientist 22: 313‐322, 2016.
 63.Cho YT, Ernst M, Fudge JL. Cortico‐amygdala‐striatal circuits are organized as hierarchical subsystems through the primate amygdala. J Neurosci 33: 14017‐14030, 2013.
 64.Chu HY, McIver EL, Kovaleski RF, Atherton JF, Bevan MD. Loss of hyperdirect pathway cortico‐subthalamic inputs following degeneration of midbrain dopamine neurons. Neuron 95: 1306‐1318, e1305, 2017.
 65.Chuhma N, Mingote S, Yetnikoff L, Kalmbach A, Ma T, Ztaou S, Sienna AC, Tepler S, Poulin JF, Ansorge M, Awatramani R, Kang UJ, Rayport S. Dopamine neuron glutamate cotransmission evokes a delayed excitation in lateral dorsal striatal cholinergic interneurons. eLife 7: e39786, 2018. DOI: 10.7554/eLife.39786.
 66.Coizet V, Graham JH, Moss J, Bolam JP, Savasta M, McHaffie JG, Redgrave P, Overton PG. Short‐latency visual input to the subthalamic nucleus is provided by the midbrain superior colliculus. J Neurosci 29: 5701‐5709, 2009.
 67.Corbit LH, Leung BK, Balleine BW. The role of the amygdala‐striatal pathway in the acquisition and performance of goal‐directed instrumental actions. J Neurosci 33: 17682‐17690, 2013.
 68.Costa RM, Lin SC, Sotnikova TD, Cyr M, Gainetdinov RR, Caron MG, Nicolelis MA. Rapid alterations in corticostriatal ensemble coordination during acute dopamine‐dependent motor dysfunction. Neuron 52: 359‐369, 2006.
 69.Cragg SJ, Baufreton J, Xue Y, Bolam JP, Bevan MD. Synaptic release of dopamine in the subthalamic nucleus. Eur J Neurosci 20: 1788‐1802, 2004.
 70.Crittenden JR, Graybiel AM. Basal ganglia disorders associated with imbalances in the striatal striosome and matrix compartments. Front Neuroanat 5: 59, 2011.
 71.Crittenden JR, Graybiel AM. Disease‐associated changes in the striosome and matrix compartments of the dorsal striatum. In: Steiner H, Tseng KY, editors. Handbook of Basal Ganglia Structure and Function. 2nd edition, London, Oxford, San Diego, Cambridge MA: Academic Press, 2017, p. 783‐802.
 72.Crossman AR, Sambrook MA, Jackson A. Experimental hemichorea/hemiballismus in the monkey. Studies on the intracerebral site of action in a drug‐induced dyskinesia. Brain 107 (Pt 2): 579‐596, 1984.
 73.Cui G, Jun SB, Jin X, Pham MD, Vogel SS, Lovinger DM, Costa RM. Concurrent activation of striatal direct and indirect pathways during action initiation. Nature 494: 238‐242, 2013.
 74.da Silva JA, Tecuapetla F, Paixao V, Costa RM. Dopamine neuron activity before action initiation gates and invigorates future movements. Nature 554: 244‐248, 2018.
 75.Damodaran S, Cressman JR, Jedrzejewski‐Szmek Z, Blackwell KT. Desynchronization of fast‐spiking interneurons reduces beta‐band oscillations and imbalance in firing in the dopamine‐depleted striatum. J Neurosci 35: 1149‐1159, 2015.
 76.Damodaran S, Evans RC, Blackwell KT. Synchronized firing of fast‐spiking interneurons is critical to maintain balanced firing between direct and indirect pathway neurons of the striatum. J Neurophysiol 111: 836‐848, 2014.
 77.Dautan D, Huerta‐Ocampo I, Valencia M, Kondabolu K, Gerdjikov TV, Mena‐Segovia J. Cholinergic midbrain afferents modulate striatal circuits and shape encoding of action control. Nat Commun 11: 1739, 2020.
 78.Dautan D, Huerta‐Ocampo I, Witten IB, Deisseroth K, Bolam JP, Gerdjikov T, Mena‐Segovia J. A major external source of cholinergic innervation of the striatum and nucleus accumbens originates in the brainstem. J Neurosci 34: 4509‐4518, 2014.
 79.Davis MI, Crittenden JR, Feng AY, Kupferschmidt DA, Naydenov A, Stella N, Graybiel AM, Lovinger DM. The cannabinoid‐1 receptor is abundantly expressed in striatal striosomes and striosome‐dendron bouquets of the substantia nigra. PLoS One 13: e0191436, 2018.
 80.Day M, Wokosin D, Plotkin JL, Tian X, Surmeier DJ. Differential excitability and modulation of striatal medium spiny neuron dendrites. J Neurosci 28: 11603‐11614, 2008.
 81.DeLong MR. Primate models of movement disorders of basal ganglia origin. Trends Neurosci 13: 281‐285, 1990.
 82.DeLong MR, Crutcher MD, Georgopoulos AP. Primate globus pallidus and subthalamic nucleus: Functional organization. J Neurophysiol 53: 530‐543, 1985.
 83.Deng P, Zhang Y, Xu ZC. Involvement of I(h) in dopamine modulation of tonic firing in striatal cholinergic interneurons. J Neurosci 27: 3148‐3156, 2007.
 84.Deniau JM, Mailly P, Maurice N, Charpier S. The pars reticulata of the substantia nigra: A window to basal ganglia output. Prog Brain Res 160: 151‐172, 2007.
 85.Di Filippo M, Picconi B, Tantucci M, Ghiglieri V, Bagetta V, Sgobio C, Tozzi A, Parnetti L, Calabresi P. Short‐term and long‐term plasticity at corticostriatal synapses: Implications for learning and memory. Behav Brain Res 199: 108‐118, 2009.
 86.Diederich NJ, James Surmeier D, Uchihara T, Grillner S, Goetz CG. Parkinson's disease: Is it a consequence of human brain evolution? Mov Disord 34: 453‐459, 2019.
 87.Ding JB, Guzman JN, Peterson JD, Goldberg JA, Surmeier DJ. Thalamic gating of corticostriatal signaling by cholinergic interneurons. Neuron 67: 294‐307, 2010.
 88.Dodson PD, Larvin JT, Duffell JM, Garas FN, Doig NM, Kessaris N, Duguid IC, Bogacz R, Butt SJ, Magill PJ. Distinct developmental origins manifest in the specialized encoding of movement by adult neurons of the external globus pallidus. Neuron 86: 501‐513, 2015.
 89.Doig NM, Magill PJ, Apicella P, Bolam JP, Sharott A. Cortical and thalamic excitation mediate the multiphasic responses of striatal cholinergic interneurons to motivationally salient stimuli. J Neurosci 34: 3101‐3117, 2014.
 90.Dorman DB, Jedrzejewska‐Szmek J, Blackwell KT. Inhibition enhances spatially‐specific calcium encoding of synaptic input patterns in a biologically constrained model. eLife 7: e38588, 2018. DOI: 10.7554/eLife.38588.
 91.Dreyer JK. Three mechanisms by which striatal denervation causes breakdown of dopamine signaling. J Neurosci 34: 12444‐12456, 2014.
 92.Du K, Wu YW, Lindroos R, Liu Y, Rozsa B, Katona G, Ding JB, Kotaleski JH. Cell‐type‐specific inhibition of the dendritic plateau potential in striatal spiny projection neurons. Proc Natl Acad Sci USA 114: E7612‐E7621, 2017.
 93.Duval C, Panisset M, Strafella AP, Sadikot AF. The impact of ventrolateral thalamotomy on tremor and voluntary motor behavior in patients with Parkinson's disease. Exp Brain Res 170: 160‐171, 2006.
 94.Duvoisin RC. Cholinergic‐anticholinergic antagonism in parkinsonism. Arch Neurol 17: 124‐136, 1967.
 95.Eblen F, Graybiel AM. Highly restricted origin of prefrontal cortical inputs to striosomes in the macaque monkey. J Neurosci 15: 5999‐6013, 1995.
 96.Edgerton JR, Hanson JE, Gunay C, Jaeger D. Dendritic sodium channels regulate network integration in globus pallidus neurons: A modeling study. J Neurosci 30: 15146‐15159, 2010.
 97.Edgerton JR, Jaeger D. Dendritic sodium channels promote active decorrelation and reduce phase locking to parkinsonian input oscillations in model globus pallidus neurons. J Neurosci 31: 10919‐10936, 2011.
 98.Elghaba R, Vautrelle N, Bracci E. Mutual control of cholinergic and low‐threshold spike interneurons in the striatum. Front Cell Neurosci 10: 111, 2016.
 99.Ellender TJ, Harwood J, Kosillo P, Capogna M, Bolam JP. Heterogeneous properties of central lateral and parafascicular thalamic synapses in the striatum. J Physiol 591: 257‐272, 2013.
 100.Ellender TJ, Huerta‐Ocampo I, Deisseroth K, Capogna M, Bolam JP. Differential modulation of excitatory and inhibitory striatal synaptic transmission by histamine. J Neurosci 31: 15340‐15351, 2011.
 101.Endepols H, Helmbold F, Walkowiak W. GABAergic projection neurons in the basal ganglia of the green tree frog (Hyla cinerea). Brain Res 1138: 76‐85, 2007.
 102.Ericsson J, Silberberg G, Robertson B, Wikstrom MA, Grillner S. Striatal cellular properties conserved from lampreys to mammals. J Physiol 589: 2979‐2992, 2011.
 103.Ericsson J, Stephenson‐Jones M, Kardamakis A, Robertson B, Silberberg G, Grillner S. Evolutionarily conserved differences in pallial and thalamic short‐term synaptic plasticity in striatum. J Physiol 591: 859‐874, 2013.
 104.Ericsson J, Stephenson‐Jones M, Perez‐Fernandez J, Robertson B, Silberberg G, Grillner S. Dopamine differentially modulates the excitability of striatal neurons of the direct and indirect pathways in lamprey. J Neurosci 33: 8045‐8054, 2013.
 105.Fernandez E, Schiappa R, Girault JA, Le Novere N. DARPP‐32 is a robust integrator of dopamine and glutamate signals. PLoS Comput Biol 2: e176, 2006.
 106.Filipovic M, Ketzef M, Reig R, Aertsen A, Silberberg G, Kumar A. Direct pathway neurons in mouse dorsolateral striatum in vivo receive stronger synaptic input than indirect pathway neurons. J Neurophysiol 122: 2294‐2303, 2019.
 107.Fino E, Vandecasteele M, Perez S, Saudou F, Venance L. Region‐specific and state‐dependent action of striatal GABAergic interneurons. Nat Commun 9: 3339, 2018.
 108.Fino E, Venance L. Spike‐timing dependent plasticity in striatal interneurons. Neuropharmacology 60: 780‐788, 2011.
 109.Fishell G, van der Kooy D. Pattern formation in the striatum: Developmental changes in the distribution of striatonigral neurons. J Neurosci 7: 1969‐1978, 1987.
 110.Fisher SD, Robertson PB, Black MJ, Redgrave P, Sagar MA, Abraham WC, Reynolds JNJ. Reinforcement determines the timing dependence of corticostriatal synaptic plasticity in vivo. Nat Commun 8: 334, 2017.
 111.Foncelle A, Mendes A, Jedrzejewska‐Szmek J, Valtcheva S, Berry H, Blackwell KT, Venance L. Modulation of spike‐timing dependent plasticity: Towards the inclusion of a third factor in computational models. Front Comput Neurosci 12: 49, 2018.
 112.Fonseca MS, Murakami M, Mainen ZF. Activation of dorsal raphe serotonergic neurons promotes waiting but is not reinforcing. Curr Biol 25: 306‐315, 2015.
 113.Francois C, Savy C, Jan C, Tande D, Hirsch EC, Yelnik J. Dopaminergic innervation of the subthalamic nucleus in the normal state, in MPTP‐treated monkeys, and in Parkinson's disease patients. J Comp Neurol 425: 121‐129, 2000.
 114.Frank MJ. Computational models of motivated action selection in corticostriatal circuits. Curr Opin Neurobiol 21: 381‐386, 2011.
 115.Fuentes R, Petersson P, Siesser WB, Caron MG, Nicolelis MA. Spinal cord stimulation restores locomotion in animal models of Parkinson's disease. Science 323: 1578‐1582, 2009.
 116.Fujiyama F, Takahashi S, Karube F. Morphological elucidation of basal ganglia circuits contributing reward prediction. Front Neurosci 9: 6, 2015.
 117.Galvan A, Hu X, Rommelfanger KS, Pare JF, Khan ZU, Smith Y, Wichmann T. Localization and function of dopamine receptors in the subthalamic nucleus of normal and parkinsonian monkeys. J Neurophysiol 112: 467‐479, 2014.
 118.Galvan A, Smith Y. The primate thalamostriatal systems: Anatomical organization, functional roles and possible involvement in Parkinson's disease. Basal Ganglia 1: 179‐189, 2011.
 119.Ganz J, Kaslin J, Freudenreich D, Machate A, Geffarth M, Brand M. Subdivisions of the adult zebrafish subpallium by molecular marker analysis. J Comp Neurol 520: 633‐655, 2012.
 120.Gerfen CR, Bolam JP. The neuroanatomical organization of the basal ganglia. In: Steiner H, Tseng KY, editors. Handbook of Basal Ganglia Structure and Function. 2nd edition, London, Oxford, San Diego, Cambridge MA: Academic Press, 2017, p. 3‐32.
 121.Gerfen CR, Surmeier DJ. Modulation of striatal projection systems by dopamine. Annu Rev Neurosci 34: 441‐466, 2011.
 122.Gironell A, Pascual‐Sedano B, Aracil I, Marin‐Lahoz J, Pagonabarraga J, Kulisevsky J. Tremor types in Parkinson disease: A descriptive study using a new classification. Parkinsons Dis 2018: 4327597, 2018.
 123.Goetz CG. The history of Parkinson's disease: Early clinical descriptions and neurological therapies. Cold Spring Harb Perspect Med 1: a008862, 2011.
 124.Gonzalez A, Lopez JM, Marin O. Expression pattern of the homeobox protein NKX2‐1 in the developing Xenopus forebrain. Brain Res Gene Expr Patterns 1: 181‐185, 2002.
 125.Gonzalez A, Morona R, Moreno N, Bandin S, Lopez JM. Identification of striatal and pallidal regions in the subpallium of anamniotes. Brain Behav Evol 83: 93‐103, 2014.
 126.Gorodetski L, Zeira R, Lavian H, Korngreen A. Long‐term plasticity of glutamatergic input from the subthalamic nucleus to the entopeduncular nucleus. Eur J Neurosci 48: 2139‐2151, 2018.
 127.Gouty‐Colomer LA, Michel FJ, Baude A, Lopez‐Pauchet C, Dufour A, Cossart R, Hammond C. Mouse subthalamic nucleus neurons with local axon collaterals. J Comp Neurol 526: 275‐284, 2018.
 128.Gowers WR. A Manual of Diseases of the Nervous System II. London: J. & A. Churchill, 1886.
 129.Graybiel AM. Neurotransmitters and neuromodulators in the basal ganglia. Trends Neurosci 13: 244‐254, 1990.
 130.Graybiel AM. The basal ganglia and chunking of action repertoires. Neurobiol Learn Mem 70: 119‐136, 1998.
 131.Graybiel AM. The basal ganglia: learning new tricks and loving it. Curr Opin Neurobiol 15: 638‐644, 2005.
 132.Graybiel AM, Hickey TL. Chemospecificity of ontogenetic units in the striatum: Demonstration by combining [3H]thymidine neuronography and histochemical staining. Proc Natl Acad Sci U S A 79: 198‐202, 1982.
 133.Graybiel AM, Ragsdale CW Jr. Histochemically distinct compartments in the striatum of human, monkeys, and cat demonstrated by acetylthiocholinesterase staining. Proc Natl Acad Sci U S A 75: 5723‐5726, 1978.
 134.Greengard P, Allen PB, Nairn AC. Beyond the dopamine receptor: The DARPP‐32/protein phosphatase‐1 cascade. Neuron 23: 435‐447, 1999.
 135.Grillner P, Svensson TH. Nicotine‐induced excitation of midbrain dopamine neurons in vitro involves ionotropic glutamate receptor activation. Synapse 38: 1‐9, 2000.
 136.Grillner S. Locomotion in vertebrates: Central mechanisms and reflex interaction. Physiol Rev 55: 247‐304, 1975.
 137.Grillner S. Control of locomotion in bipeds, tetrapods and fish. In: Brooks VB, editor. Handbook of Physiology, Sect 1 The Nervous System II Motor Contol. Maryland: American Physiological Society, Waverly Press, 1981, p. 1179‐1236.
 138.Grillner S. The motor infrastructure: From ion channels to neuronal networks. Nat Rev Neurosci 4: 573‐586, 2003.
 139.Grillner S. Biological pattern generation: The cellular and computational logic of networks in motion. Neuron 52: 751‐766, 2006.
 140.Grillner S. The enigmatic “indirect pathway” of the basal ganglia: A new role. Neuron 99: 1105‐1107, 2018.
 141.Grillner S, Hellgren J, Menard A, Saitoh K, Wikstrom MA. Mechanisms for selection of basic motor programs—Roles for the striatum and pallidum. Trends Neurosci 28: 364‐370, 2005.
 142.Grillner S, Robertson B. The basal ganglia over 500 million years. Curr Biol 26: R1088‐R1100, 2016.
 143.Gurney K, Prescott TJ, Redgrave P. A computational model of action selection in the basal ganglia. I. A new functional anatomy. Biol Cybern 84: 401‐410, 2001.
 144.Gurney K, Prescott TJ, Redgrave P. A computational model of action selection in the basal ganglia. II. Analysis and simulation of behaviour. Biol Cybern 84: 411‐423, 2001.
 145.Gurney KN, Humphries MD, Redgrave P. A new framework for cortico‐striatal plasticity: Behavioural theory meets in vitro data at the reinforcement‐action interface. PLoS Biol 13: e1002034, 2015.
 146.Gutierrez‐Arenas O, Eriksson O, Kotaleski JH. Segregation and crosstalk of D1 receptor‐mediated activation of ERK in striatal medium spiny neurons upon acute administration of psychostimulants. PLoS Comput Biol 10: e1003445, 2014.
 147.Haas H, Panula P. The role of histamine and the tuberomamillary nucleus in the nervous system. Nat Rev Neurosci 4: 121‐130, 2003.
 148.Haas HL, Konnerth A. Histamine and noradrenaline decrease calcium‐activated potassium conductance in hippocampal pyramidal cells. Nature 302: 432‐434, 1983.
 149.Haber SN. Corticostriatal circuitry. Dialogues Clin Neurosci 18: 7‐21, 2016.
 150.Hawes SL, Gillani F, Evans RC, Benkert EA, Blackwell KT. Sensitivity to theta‐burst timing permits LTP in dorsal striatal adult brain slice. J Neurophysiol 110: 2027‐2036, 2013.
 151.Hawrylycz M, Anastassiou C, Arkhipov A, Berg J, Buice M, Cain N, Gouwens NW, Gratiy S, Iyer R, Lee JH, Mihalas S, Mitelut C, Olsen S, Reid RC, Teeter C, de Vries S, Waters J, Zeng H, Koch C, MindScope. Inferring cortical function in the mouse visual system through large‐scale systems neuroscience. Proc Natl Acad Sci U S A 113: 7337‐7344, 2016.
 152.Haynes WI, Haber SN. The organization of prefrontal‐subthalamic inputs in primates provides an anatomical substrate for both functional specificity and integration: Implications for Basal Ganglia models and deep brain stimulation. J Neurosci 33: 4804‐4814, 2013.
 153.Heimer L, Wilson RD. The subcortical projections of the allocortex: Similarities in the neuronal associations of the hippocampus, the pyriform cortex and the neocortex. In: Santini M, editor. Golgi Centennial Symposium: Perspectives in Neurobiology. New York: Raven Press, 1975, p. 177‐193.
 154.Hernandez LF, Obeso I, Costa RM, Redgrave P, Obeso JA. Dopaminergic vulnerability in Parkinson disease: The cost of humans' habitual performance. Trends Neurosci 42: 375‐383, 2019.
 155.Hikosaka O. The habenula: From stress evasion to value‐based decision‐making. Nat Rev Neurosci 11: 503‐513, 2010.
 156.Hikosaka O, Wurtz RH. Visual and oculomotor functions of monkey substantia nigra pars reticulata. I. Relation of visual and auditory responses to saccades. J Neurophysiol 49: 1230‐1253, 1983.
 157.Hikosaka O, Wurtz RH. Visual and oculomotor functions of monkey substantia nigra pars reticulata. II. Visual responses related to fixation of gaze. J Neurophysiol 49: 1254‐1267, 1983.
 158.Hikosaka O, Wurtz RH. Visual and oculomotor functions of monkey substantia nigra pars reticulata. III. Memory‐contingent visual and saccade responses. J Neurophysiol 49: 1268‐1284, 1983.
 159.Hikosaka O, Wurtz RH. Visual and oculomotor functions of monkey substantia nigra pars reticulata. IV. Relation of substantia nigra to superior colliculus. J Neurophysiol 49: 1285‐1301, 1983.
 160.Hjorth J, Blackwell KT, Kotaleski JH. Gap junctions between striatal fast‐spiking interneurons regulate spiking activity and synchronization as a function of cortical activity. J Neurosci 29: 5276‐5286, 2009.
 161.Hjorth JJJ, Kozlov A, Carannante I, Frost Nylén J, Carannante I, Lindroos R, Johansson Y, Tokarska A, Dorst MC, Suryanarayana SM, Silberberg G, Hellgren Kotaleski J, Grillners. The microcircuits of striatum in silico. Proc Natl Acad Sci USA 117: 9554‐9565, 2020.
 162.Hong S, Amemori S, Chung E, Gibson DJ, Amemori KI, Graybiel AM. Predominant striatal input to the lateral habenula in macaques comes from striosomes. Curr Biol 29: 51‐61, e55, 2019.
 163.Hong S, Hikosaka O. The globus pallidus sends reward‐related signals to the lateral habenula. Neuron 60: 720‐729, 2008.
 164.Hornykiewicz O. Dopamine (3‐hydroxytyramine) in the central nervous system and its relation to the Parkinson syndrome in man. Dtsch Med Wochenschr 87: 1807‐1810, 1962.
 165.Hoshi E, Tremblay L, Feger J, Carras PL, Strick PL. The cerebellum communicates with the basal ganglia. Nat Neurosci 8: 1491‐1493, 2005.
 166.Humphries MD, Obeso JA, Dreyer JK. Insights into Parkinson's disease from computational models of the basal ganglia. J Neurol Neurosurg Psychiatry 89: 1181‐1188, 2018.
 167.Humphries MD, Stewart RD, Gurney KN. A physiologically plausible model of action selection and oscillatory activity in the basal ganglia. J Neurosci 26: 12921‐12942, 2006.
 168.Hwang EJ, Link TD, Hu YY, Lu S, Wang EH, Lilascharoen V, Aronson S, O'Neil K, Lim BK, Komiyama T. Corticostriatal flow of action selection bias. Neuron 104: 1126‐1140, e1126, 2019.
 169.Ichinohe N, Mori F, Shoumura K. A di‐synaptic projection from the lateral cerebellar nucleus to the laterodorsal part of the striatum via the central lateral nucleus of the thalamus in the rat. Brain Res 880: 191‐197, 2000.
 170.Iigaya K, Fonseca MS, Murakami M, Mainen ZF, Dayan P. An effect of serotonergic stimulation on learning rates for rewards apparent after long intertrial intervals. Nat Commun 9: 2477, 2018.
 171.Ito M. Control of mental activities by internal models in the cerebellum. Nat Rev Neurosci 9: 304‐313, 2008.
 172.Jacobs BL, Fornal CA. Serotonin and motor activity. Curr Opin Neurobiol 7: 820‐825, 1997.
 173.Jankovic J. Parkinson's disease: Clinical features and diagnosis. J Neurol Neurosurg Psychiatry 79: 368‐376, 2008.
 174.Jin X, Costa RM. Shaping action sequences in basal ganglia circuits. Curr Opin Neurobiol 33: 188‐196, 2015.
 175.Jin X, Tecuapetla F, Costa RM. Basal ganglia subcircuits distinctively encode the parsing and concatenation of action sequences. Nat Neurosci 17: 423‐430, 2014.
 176.Johnels B, Ingvarsson PE, Steg G, Olsson T. The posturo‐locomotion‐manual test. A simple method for the characterization of neurological movement disturbances. Adv Neurol 87: 91‐100, 2001.
 177.Karachi C, Grabli D, Bernard FA, Tande D, Wattiez N, Belaid H, Bardinet E, Prigent A, Nothacker HP, Hunot S, Hartmann A, Lehericy S, Hirsch EC, Francois C. Cholinergic mesencephalic neurons are involved in gait and postural disorders in Parkinson disease. J Clin Invest 120: 2745‐2754, 2010.
 178.Karube F, Takahashi S, Kobayashi K, Fujiyama F. Motor cortex can directly drive the globus pallidal neurons in a projection neuron type dependent manner in rat. eLife 8: e49511, 2019. DOI: 10.7554/eLife.49511.
 179.Kawai R, Markman T, Poddar R, Ko R, Fantana AL, Dhawale AK, Kampff AR, Olveczky BP. Motor cortex is required for learning but not for executing a motor skill. Neuron 86: 800‐812, 2015.
 180.Ketzef M, Silberberg G. Differential synaptic input to external globus pallidus neuronal subpopulations in vivo. bioRxiv, DOI: 10.1101/2020.02.27.967869, 2020.
 181.Ketzef M, Spigolon G, Johansson Y, Bonito‐Oliva A, Fisone G, Silberberg G. Dopamine depletion impairs bilateral sensory processing in the striatum in a pathway‐dependent manner. Neuron 94: 855‐865, e855, 2017.
 182.Kim HF, Amita H, Hikosaka O. Indirect pathway of caudal basal ganglia for rejection of valueless visual objects. Neuron 94: 920‐930, e923, 2017.
 183.Kim HF, Hikosaka O. Distinct basal ganglia circuits controlling behaviors guided by flexible and stable values. Neuron 79: 1001‐1010, 2013.
 184.Kim HF, Hikosaka O. Parallel basal ganglia circuits for voluntary and automatic behaviour to reach rewards. Brain 138: 1776‐1800, 2015.
 185.Kita H, Jaeger D. Organization of the globus pallidus. In: Steiner H, Tseng KY, editors. Handbook of Basal Ganglia Structure and Function. 2nd edition, London, Oxford, San Diego, Cambridge MA: Academic Press, 2017, p. 259‐276.
 186.Kita H, Kitai ST. Intracellular study of rat globus pallidus neurons: Membrane properties and responses to neostriatal, subthalamic and nigral stimulation. Brain Res 564: 296‐305, 1991.
 187.Kita T, Kita H. Cholinergic and non‐cholinergic mesopontine tegmental neurons projecting to the subthalamic nucleus in the rat. Eur J Neurosci 33: 433‐443, 2011.
 188.Klaus A, da Silva JA, Costa RM. What, if, and when to move: Basal ganglia circuits and self‐paced action initiation. Annu Rev Neurosci 42: 459‐483, 2019.
 189.Klaus A, Martins GJ, Paixao VB, Zhou P, Paninski L, Costa RM. The spatiotemporal organization of the striatum encodes action space. Neuron 95: 1171‐1180, e1177, 2017.
 190.Klug JR, Engelhardt MD, Cadman CN, Li H, Smith JB, Ayala S, Williams EW, Hoffman H, Jin X. Differential inputs to striatal cholinergic and parvalbumin interneurons imply functional distinctions. eLife 7: e38588, 2018. DOI: 10.7554/eLife.38588.
 191.Koos T, Tepper JM. Inhibitory control of neostriatal projection neurons by GABAergic interneurons. Nat Neurosci 2: 467‐472, 1999.
 192.Koralek AC, Costa RM, Carmena JM. Temporally precise cell‐specific coherence develops in corticostriatal networks during learning. Neuron 79: 865‐872, 2013.
 193.Koralek AC, Jin X, Long JD II, Costa RM, Carmena JM. Corticostriatal plasticity is necessary for learning intentional neuroprosthetic skills. Nature 483: 331‐335, 2012.
 194.Korzeniewska A, Kasicki S, Kaminski M, Blinowska KJ. Information flow between hippocampus and related structures during various types of rat's behavior. J Neurosci Methods 73: 49‐60, 1997.
 195.Kotaleski JH, Plenz D, Blackwell KT. Using potassium currents to solve signal‐to‐noise problems in inhibitory feedforward networks of the striatum. J Neurophysiol 95: 331‐341, 2006.
 196.Kozlov A, Huss M, Lansner A, Kotaleski JH, Grillner S. Simple cellular and network control principles govern complex patterns of motor behavior. Proc Natl Acad Sci U S A 106: 20027‐20032, 2009.
 197.Kravitz AV, Freeze BS, Parker PR, Kay K, Thwin MT, Deisseroth K, Kreitzer AC. Regulation of parkinsonian motor behaviours by optogenetic control of basal ganglia circuitry. Nature 466: 622‐626, 2010.
 198.Kravitz AV, Tye LD, Kreitzer AC. Distinct roles for direct and indirect pathway striatal neurons in reinforcement. Nat Neurosci 15: 816‐818, 2012.
 199.Kreitzer AC, Malenka RC. Striatal plasticity and basal ganglia circuit function. Neuron 60: 543‐554, 2008.
 200.Krienen FM, Goldman M, Zhang Q, del Rosario R, Florio M, Machold R, Saunders A, Levandowski K, Zaniewski H, Schuman B, Wu C, Lutservitz A, Mullally CD, Reed N, Bien E, Bortolin L, Fernandez‐Otero M, Lin J, Wysoker A, Nemesh J, Kulp D, Burns M, Tkachev V, Smith R, Walsh CA, Dimidschstein J, Rudy B, Kean L, Berretta S, Fishell G, Feng G, McCarroll SA. Innovations in primate interneuron repertoire. bioRxiv: 709501, 2019.
 201.Kumar A, Cardanobile S, Rotter S, Aertsen A. The role of inhibition in generating and controlling Parkinson's disease oscillations in the Basal Ganglia. Front Syst Neurosci 5: 86, 2011.
 202.Lacey CJ, Bolam JP, Magill PJ. Novel and distinct operational principles of intralaminar thalamic neurons and their striatal projections. J Neurosci 27: 4374‐4384, 2007.
 203.Lammel S, Lim BK, Ran C, Huang KW, Betley MJ, Tye KM, Deisseroth K, Malenka RC. Input‐specific control of reward and aversion in the ventral tegmental area. Nature 491: 212‐217, 2012.
 204.Lawrence DG, Kuypers HG. The functional organization of the motor system in the monkey. I. The effects of bilateral pyramidal lesions. Brain 91: 1‐14, 1968.
 205.Lawrence DG, Kuypers HG. The functional organization of the motor system in the monkey. II. The effects of lesions of the descending brain‐stem pathways. Brain 91: 15‐36, 1968.
 206.Lazaridis I, Tzortzi O, Weglage M, Martin A, Xuan Y, Parent M, Johansson Y, Fuzik J, Furth D, Fenno LE, Ramakrishnan C, Silberberg G, Deisseroth K, Carlen M, Meletis K. A hypothalamus‐habenula circuit controls aversion. Mol Psychiatry 24: 1351‐1368, 2019.
 207.Lee KJ, Shim I, Sung JH, Hong JT, Kim IS, Cho CB. Striatal glutamate and GABA after high frequency subthalamic stimulation in Parkinsonian rat. J Korean Neurosurg Soc 60: 138‐145, 2017.
 208.Lepeta K, Lourenco MV, Schweitzer BC, Martino Adami PV, Banerjee P, Catuara‐Solarz S, de La Fuente Revenga M, Guillem AM, Haidar M, Ijomone OM, Nadorp B, Qi L, Perera ND, Refsgaard LK, Reid KM, Sabbar M, Sahoo A, Schaefer N, Sheean RK, Suska A, Verma R, Vicidomini C, Wright D, Zhang XD, Seidenbecher C. Synaptopathies: Synaptic dysfunction in neurological disorders—A review from students to students. J Neurochem 138: 785‐805, 2016.
 209.Lepora NF, Gurney KN. The basal ganglia optimize decision making over general perceptual hypotheses. Neural Comput 24: 2924‐2945, 2012.
 210.Lerner TN, Kreitzer AC. RGS4 is required for dopaminergic control of striatal LTD and susceptibility to parkinsonian motor deficits. Neuron 73: 347‐359, 2012.
 211.Lindahl M, Hellgren Kotaleski J. Untangling Basal Ganglia network dynamics and function: Role of dopamine depletion and inhibition investigated in a spiking network model. eNeuro 3(6): ENEURO.0156‐16, 2016.
 212.Lindroos R, Dorst MC, Du K, Filipovic M, Keller D, Ketzef M, Kozlov AK, Kumar A, Lindahl M, Nair AG, Perez‐Fernandez J, Grillner S, Silberberg G, Hellgren Kotaleski J. Basal Ganglia neuromodulation over multiple temporal and structural scales‐simulations of direct pathway MSNs investigate the fast onset of dopaminergic effects and predict the role of Kv4.2. Front Neural Circuits 12: 3, 2018.
 213.Lindskog M, Kim M, Wikstrom MA, Blackwell KT, Kotaleski JH. Transient calcium and dopamine increase PKA activity and DARPP‐32 phosphorylation. PLoS Comput Biol 2: e119, 2006.
 214.Lisman J. Glutamatergic synapses are structurally and biochemically complex because of multiple plasticity processes: Long‐term potentiation, long‐term depression, short‐term potentiation and scaling. Philos Trans R Soc Lond B Biol Sci 372(1715): 20160260, 2017.
 215.Lottem E, Banerjee D, Vertechi P, Sarra D, Lohuis MO, Mainen ZF. Activation of serotonin neurons promotes active persistence in a probabilistic foraging task. Nat Commun 9: 1000, 2018.
 216.Lovinger DM. Neurotransmitter roles in synaptic modulation, plasticity and learning in the dorsal striatum. Neuropharmacology 58: 951‐961, 2010.
 217.Magill PJ, Bolam JP, Bevan MD. Dopamine regulates the impact of the cerebral cortex on the subthalamic nucleus‐globus pallidus network. Neuroscience 106: 313‐330, 2001.
 218.Maier S, Walkowiak W, Luksch H, Endepols H. An indirect basal ganglia pathway in anuran amphibians? J Chem Neuroanat 40: 21‐35, 2010.
 219.Mallet N, Micklem BR, Henny P, Brown MT, Williams C, Bolam JP, Nakamura KC, Magill PJ. Dichotomous organization of the external globus pallidus. Neuron 74: 1075‐1086, 2012.
 220.Mallet N, Pogosyan A, Marton LF, Bolam JP, Brown P, Magill PJ. Parkinsonian beta oscillations in the external globus pallidus and their relationship with subthalamic nucleus activity. J Neurosci 28: 14245‐14258, 2008.
 221.Mallet N, Pogosyan A, Sharott A, Csicsvari J, Bolam JP, Brown P, Magill PJ. Disrupted dopamine transmission and the emergence of exaggerated beta oscillations in subthalamic nucleus and cerebral cortex. J Neurosci 28: 4795‐4806, 2008.
 222.Mallet N, Schmidt R, Leventhal D, Chen F, Amer N, Boraud T, Berke JD. Arkypallidal cells send a stop signal to striatum. Neuron 89: 308‐316, 2016.
 223.Mandelbaum G, Taranda J, Haynes TM, Hochbaum DR, Huang KW, Hyun M, Umadevi Venkataraju K, Straub C, Wang W, Robertson K, Osten P, Sabatini BL. Distinct cortical‐thalamic‐striatal circuits through the parafascicular nucleus. Neuron 102: 636‐652, e637, 2019.
 224.Marin O, Smeets WJ, Gonzalez A. Basal ganglia organization in amphibians: Chemoarchitecture. J Comp Neurol 392: 285‐312, 1998.
 225.Markowitz JE, Gillis WF, Beron CC, Neufeld SQ, Robertson K, Bhagat ND, Peterson RE, Peterson E, Hyun M, Linderman SW, Sabatini BL, Datta SR. The striatum organizes 3D behavior via moment‐to‐moment action selection. Cell 174: 44‐58, e17, 2018.
 226.Markram H, Muller E, Ramaswamy S, Reimann MW, Abdellah M, Sanchez CA, Ailamaki A, Alonso‐Nanclares L, Antille N, Arsever S, Kahou GA, Berger TK, Bilgili A, Buncic N, Chalimourda A, Chindemi G, Courcol JD, Delalondre F, Delattre V, Druckmann S, Dumusc R, Dynes J, Eilemann S, Gal E, Gevaert ME, Ghobril JP, Gidon A, Graham JW, Gupta A, Haenel V, Hay E, Heinis T, Hernando JB, Hines M, Kanari L, Keller D, Kenyon J, Khazen G, Kim Y, King JG, Kisvarday Z, Kumbhar P, Lasserre S, Le Be JV, Magalhaes BR, Merchan‐Perez A, Meystre J, Morrice BR, Muller J, Munoz‐Cespedes A, Muralidhar S, Muthurasa K, Nachbaur D, Newton TH, Nolte M, Ovcharenko A, Palacios J, Pastor L, Perin R, Ranjan R, Riachi I, Rodriguez JR, Riquelme JL, Rossert C, Sfyrakis K, Shi Y, Shillcock JC, Silberberg G, Silva R, Tauheed F, Telefont M, Toledo‐Rodriguez M, Trankler T, Van Geit W, Diaz JV, Walker R, Wang Y, Zaninetta SM, DeFelipe J, Hill SL, Segev I, Schurmann F. Reconstruction and simulation of neocortical microcircuitry. Cell 163: 456‐492, 2015.
 227.Martin A, Calvigioni D, Tzortzi O, Fuzik J, Warnberg E, Meletis K. A spatiomolecular map of the striatum. Cell Rep 29: 4320‐4333, e4325, 2019.
 228.Matsumoto M, Hikosaka O. Two types of dopamine neuron distinctly convey positive and negative motivational signals. Nature 459: 837‐841, 2009.
 229.Matsumoto N, Minamimoto T, Graybiel AM, Kimura M. Neurons in the thalamic CM‐Pf complex supply striatal neurons with information about behaviorally significant sensory events. J Neurophysiol 85: 960‐976, 2001.
 230.McElvain LE, Costa RM. Specialized populations of substantia nigra neurons mediate basal ganglia output signaling to brainstem and thalamus. In: Society for Neuroscience. Abstract number 340.07, Chicago, USA, 2015.
 231.McHaffie JG, Stanford TR, Stein BE, Coizet V, Redgrave P. Subcortical loops through the basal ganglia. Trends Neurosci 28: 401‐407, 2005.
 232.Melendez‐Zaidi AE, Lakshminarasimhah H, Surmeier DJ. Cholinergic modulation of striatal nitric oxide‐producing interneurons. Eur J Neurosci 50: 3713‐3731, 2019.
 233.Mendes A, Vignoud G, Perez S, Perrin E, Touboul J, Venance L. Concurrent thalamostriatal and corticostriatal spike‐timing‐dependent plasticity and heterosynaptic interactions shape striatal plasticity map. Cereb Cortex, DOI: 10.1093/cercor/bhaa024, 2020.
 234.Minamimoto T, Hori Y, Kimura M. Complementary process to response bias in the centromedian nucleus of the thalamus. Science 308: 1798‐1801, 2005.
 235.Minamimoto T, Hori Y, Yamanaka K, Kimura M. Neural signal for counteracting pre‐action bias in the centromedian thalamic nucleus. Front Syst Neurosci 8: 3, 2014.
 236.Minamimoto T, Kimura M. Participation of the thalamic CM‐Pf complex in attentional orienting. J Neurophysiol 87: 3090‐3101, 2002.
 237.Mingote S, Chuhma N, Kusnoor SV, Field B, Deutch AY, Rayport S. Functional connectome analysis of dopamine neuron glutamatergic connections in forebrain regions. J Neurosci 35: 16259‐16271, 2015.
 238.Mink JW. The basal ganglia: Focused selection and inhibition of competing motor programs. Prog Neurobiol 50: 381‐425, 1996.
 239.Mink JW. The Basal Ganglia and involuntary movements: Impaired inhibition of competing motor patterns. Arch Neurol 60: 1365‐1368, 2003.
 240.Mink JW. The Basal Ganglia. In: Squire LR, Berg D, Bloom FE, du Lac S, Ghosh A, Spitzer NC, editors. Fundamental Neuroscience. Elsevier, 2017, p. 725‐750.
 241.Miyazaki K, Miyazaki KW, Yamanaka A, Tokuda T, Tanaka KF, Doya K. Reward probability and timing uncertainty alter the effect of dorsal raphe serotonin neurons on patience. Nat Commun 9: 2048, 2018.
 242.Miyazaki KW, Miyazaki K, Tanaka KF, Yamanaka A, Takahashi A, Tabuchi S, Doya K. Optogenetic activation of dorsal raphe serotonin neurons enhances patience for future rewards. Curr Biol 24: 2033‐2040, 2014.
 243.Mohan H, An X, Musall S, Mitra PP, Churchland AK, Huang JZ. Pyramidal neuron types and cortical networks encoding sensorimotor activity coordinating hand‐mouth synergy during eating. Society for Neuroscience Meeting abstract number 227.15, 2019. https://www.abstractsonline.com/pp8/#!/7883/presentation/52434
 244.Moreno N, Gonzalez A, Retaux S. Development and evolution of the subpallium. Semin Cell Dev Biol 20: 735‐743, 2009.
 245.Morris G, Arkadir D, Nevet A, Vaadia E, Bergman H. Coincident but distinct messages of midbrain dopamine and striatal tonically active neurons. Neuron 43: 133‐143, 2004.
 246.Moss J, Bolam JP. A dopaminergic axon lattice in the striatum and its relationship with cortical and thalamic terminals. J Neurosci 28: 11221‐11230, 2008.
 247.Moyer JT, Halterman BL, Finkel LH, Wolf JA. Lateral and feedforward inhibition suppress asynchronous activity in a large, biophysically‐detailed computational model of the striatal network. Front Comput Neurosci 8: 152, 2014.
 248.Moyer JT, Wolf JA, Finkel LH. Effects of dopaminergic modulation on the integrative properties of the ventral striatal medium spiny neuron. J Neurophysiol 98: 3731‐3748, 2007.
 249.Munoz‐Manchado AB, Bengtsson Gonzales C, Zeisel A, Munguba H, Bekkouche B, Skene NG, Lonnerberg P, Ryge J, Harris KD, Linnarsson S, Hjerling‐Leffler J. Diversity of interneurons in the dorsal striatum revealed by single‐cell RNA sequencing and PatchSeq. Cell Rep 24: 2179‐2190, e2177, 2018.
 250.Nair AG, Bhalla US, Hellgren Kotaleski J. Role of DARPP‐32 and ARPP‐21 in the emergence of temporal constraints on striatal calcium and dopamine integration. PLoS Comput Biol 12: e1005080, 2016.
 251.Nair AG, Castro LRV, El Khoury M, Gorgievski V, Giros B, Tzavara ET, Hellgren‐Kotaleski J, Vincent P. The high efficacy of muscarinic M4 receptor in D1 medium spiny neurons reverses striatal hyperdopaminergia. Neuropharmacology 146: 74‐83, 2019.
 252.Nair AG, Gutierrez‐Arenas O, Eriksson O, Jauhiainen A, Blackwell KT, Kotaleski JH. Modeling intracellular signaling underlying striatal function in health and disease. Prog Mol Biol Transl Sci 123: 277‐304, 2014.
 253.Nair AG, Gutierrez‐Arenas O, Eriksson O, Vincent P, Hellgren Kotaleski J. Sensing positive versus negative reward signals through adenylyl cyclase‐coupled GPCRs in direct and indirect pathway striatal medium spiny neurons. J Neurosci 35: 14017‐14030, 2015.
 254.Nakano T, Yoshimoto J, Doya K. A model‐based prediction of the calcium responses in the striatal synaptic spines depending on the timing of cortical and dopaminergic inputs and post‐synaptic spikes. Front Comput Neurosci 7: 119, 2013.
 255.Nambu A, Takada M, Inase M, Tokuno H. Dual somatotopical representations in the primate subthalamic nucleus: Evidence for ordered but reversed body‐map transformations from the primary motor cortex and the supplementary motor area. J Neurosci 16: 2671‐2683, 1996.
 256.Nambu A, Tokuno H, Takada M. Functional significance of the cortico‐subthalamo‐pallidal ‘hyperdirect’ pathway. Neurosci Res 43: 111‐117, 2002.
 257.Navarro G, Cordomi A, Casado‐Anguera V, Moreno E, Cai NS, Cortes A, Canela EI, Dessauer CW, Casado V, Pardo L, Lluis C, Ferre S. Evidence for functional pre‐coupled complexes of receptor heteromers and adenylyl cyclase. Nat Commun 9: 1242, 2018.
 258.Navntoft CA, Dreyer JK. How compensation breaks down in Parkinson's disease: Insights from modeling of denervated striatum. Mov Disord 31: 280‐289, 2016.
 259.Nieto Mendoza E, Hernandez Echeagaray E. Dopaminergic modulation of striatal inhibitory transmission and long‐term plasticity. Neural Plast 2015: 789502, 2015.
 260.Nisenbaum ES, Wilson CJ. Potassium currents responsible for inward and outward rectification in rat neostriatal spiny projection neurons. J Neurosci 15: 4449‐4463, 1995.
 261.Nonomura S, Nishizawa K, Sakai Y, Kawaguchi Y, Kato S, Uchigashima M, Watanabe M, Yamanaka K, Enomoto K, Chiken S, Sano H, Soma S, Yoshida J, Samejima K, Ogawa M, Kobayashi K, Nambu A, Isomura Y, Kimura M. Monitoring and updating of action selection for goal‐directed behavior through the striatal direct and indirect pathways. Neuron 99: 1302‐1314, e1305, 2018.
 262.Ocana FM, Suryanarayana SM, Saitoh K, Kardamakis AA, Capantini L, Robertson B, Grillner S. The lamprey pallium provides a blueprint of the mammalian motor projections from cortex. Curr Biol 25: 413‐423, 2015.
 263.Okada K, Toyama K, Inoue Y, Isa T, Kobayashi Y. Different pedunculopontine tegmental neurons signal predicted and actual task rewards. J Neurosci 29: 4858‐4870, 2009.
 264.Oorschot DE. Total number of neurons in the neostriatal, pallidal, subthalamic, and substantia nigral nuclei of the rat basal ganglia: A stereological study using the cavalieri and optical disector methods. J Comp Neurol 366: 580‐599, 1996.
 265.Paille V, Fino E, Du K, Morera‐Herreras T, Perez S, Kotaleski JH, Venance L. GABAergic circuits control spike‐timing‐dependent plasticity. J Neurosci 33: 9353‐9363, 2013.
 266.Papathanou M, Dumas S, Pettersson H, Olson L, Wallen‐Mackenzie A. Off‐target effects in transgenic mice: Characterization of dopamine transporter (DAT)‐Cre transgenic mouse lines exposes multiple non‐dopaminergic neuronal clusters available for selective targeting within limbic neurocircuitry. eNeuro 6(5): ENEURO.0198-19, 2019.
 267.Parent A. Comparative Neurobiology of the Basal Ganglia. New York: John Wiley & Sons, 1986, p. 335.
 268.Parievsky A, Cepeda C, Levine MS. Alterations of synaptic function in Huntington's disease. In: Steiner H, Tseng KY, editors. Handbook of Basal Ganglia Structure and Function. 2nd edition, London, Oxford, San Diego, Cambridge MA: Academic Press, 2017, p. 907‐927.
 269.Parker JG, Marshall JD, Ahanonu B, Wu YW, Kim TH, Grewe BF, Zhang Y, Li JZ, Ding JB, Ehlers MD, Schnitzer MJ. Diametric neural ensemble dynamics in parkinsonian and dyskinetic states. Nature 557: 177‐182, 2018.
 270.Parkinson J. An essay on the shaking palsy. J Neuropsychiatry Clin Neurosci 14: 223‐236, 1817.
 271.Perez‐Fernandez J, Kardamakis AA, Suzuki DG, Robertson B, Grillner S. Direct dopaminergic projections from the SNc modulate visuomotor transformation in the lamprey tectum. Neuron 96: 910‐924, e915, 2017.
 272.Perez‐Fernandez J, Stephenson‐Jones M, Suryanarayana SM, Robertson B, Grillner S. Evolutionarily conserved organization of the dopaminergic system in lamprey: SNc/VTA afferent and efferent connectivity and D2 receptor expression. J Comp Neurol 522: 3775‐3794, 2014.
 273.Perrin E, Venance L. Bridging the gap between striatal plasticity and learning. Curr Opin Neurobiol 54: 104‐112, 2019.
 274.Petersson P, Halje P, Cenci MA. Significance and translational value of high‐frequency cortico‐basal ganglia oscillations in Parkinson's disease. J Parkinsons Dis 9: 183‐196, 2019.
 275.Picconi B, Piccoli G, Calabresi P. Synaptic dysfunction in Parkinson's disease. Adv Exp Med Biol 970: 553‐572, 2012.
 276.Planert H, Szydlowski SN, Hjorth JJ, Grillner S, Silberberg G. Dynamics of synaptic transmission between fast‐spiking interneurons and striatal projection neurons of the direct and indirect pathways. J Neurosci 30: 3499‐3507, 2010.
 277.Plotkin JL, Day M, Surmeier DJ. Synaptically driven state transitions in distal dendrites of striatal spiny neurons. Nat Neurosci 14: 881‐888, 2011.
 278.Plotkin JL, Shen W, Rafalovich I, Sebel LE, Day M, Chan CS, Surmeier DJ. Regulation of dendritic calcium release in striatal spiny projection neurons. J Neurophysiol 110: 2325‐2336, 2013.
 279.Plotkin JL, Surmeier DJ. Corticostriatal synaptic adaptations in Huntington's disease. Curr Opin Neurobiol 33: 53‐62, 2015.
 280.Pombal MA, El Manira A, Grillner S. Afferents of the lamprey striatum with special reference to the dopaminergic system: A combined tracing and immunohistochemical study. J Comp Neurol 386: 71‐91, 1997.
 281.Pombal MA, El Manira A, Grillner S. Organization of the lamprey striatum–Transmitters and projections. Brain Res 766: 249‐254, 1997.
 282.Pombal MA, Marin O, Gonzalez A. Distribution of choline acetyltransferase‐immunoreactive structures in the lamprey brain. J Comp Neurol 431: 105‐126, 2001.
 283.Ponzi A, Wickens JR. Optimal balance of the striatal medium spiny neuron network. PLoS Comput Biol 9: e1002954, 2013.
 284.Ranade SP, Mainen ZF. Transient firing of dorsal raphe neurons encodes diverse and specific sensory, motor, and reward events. J Neurophysiol 102: 3026‐3037, 2009.
 285.Raz A, Feingold A, Zelanskaya V, Vaadia E, Bergman H. Neuronal synchronization of tonically active neurons in the striatum of normal and parkinsonian primates. J Neurophysiol 76: 2083‐2088, 1996.
 286.Redgrave P, Rodriguez M, Smith Y, Rodriguez‐Oroz MC, Lehericy S, Bergman H, Agid Y, DeLong MR, Obeso JA. Goal‐directed and habitual control in the basal ganglia: Implications for Parkinson's disease. Nat Rev Neurosci 11: 760‐772, 2010.
 287.Reig R, Silberberg G. Multisensory integration in the mouse striatum. Neuron 83: 1200‐1212, 2014.
 288.Reig R, Silberberg G. Distinct corticostriatal and intracortical pathways mediate bilateral sensory responses in the striatum. Cereb Cortex 26: 4405‐4415, 2016.
 289.Reiner A. The conserved evolution of the vertebrate basal ganglia. In: Steiner H, Tseng KY, editors. Handbook of Basal Ganglia Structure and Function. Burlington, MA, Academic Press, 2010, p. 29‐62.
 290.Reiner A, Medina L, Veenman CL. Structural and functional evolution of the basal ganglia in vertebrates. Brain Res Brain Res Rev 28: 235‐285, 1998.
 291.Robertson B, Huerta‐Ocampo I, Ericsson J, Stephenson‐Jones M, Perez‐Fernandez J, Bolam JP, Diaz‐Heijtz R, Grillner S. The dopamine D2 receptor gene in lamprey, its expression in the striatum and cellular effects of D2 receptor activation. PLoS One 7: e35642, 2012.
 292.Rolland AS, Tande D, Herrero MT, Luquin MR, Vazquez‐Claverie M, Karachi C, Hirsch EC, Francois C. Evidence for a dopaminergic innervation of the pedunculopontine nucleus in monkeys, and its drastic reduction after MPTP intoxication. J Neurochem 110: 1321‐1329, 2009.
 293.Roseberry TK, Lalive AL, Margolin BD, Kreitzer AC. Locomotor suppression by a monosynaptic amygdala to brainstem circuit. bioRxiv: 724252, 2019.
 294.Roseberry TK, Lee AM, Lalive AL, Wilbrecht L, Bonci A, Kreitzer AC. Cell‐type‐specific control of brainstem locomotor circuits by basal ganglia. Cell 164: 526‐537, 2016.
 295.Ryczko D, Cone JJ, Alpert MH, Goetz L, Auclair F, Dube C, Parent M, Roitman MF, Alford S, Dubuc R. A descending dopamine pathway conserved from basal vertebrates to mammals. Proc Natl Acad Sci U S A 113: E2440‐E2449, 2016.
 296.Ryczko D, Gratsch S, Auclair F, Dube C, Bergeron S, Alpert MH, Cone JJ, Roitman MF, Alford S, Dubuc R. Forebrain dopamine neurons project down to a brainstem region controlling locomotion. Proc Natl Acad Sci U S A 110: E3235‐E3242, 2013.
 297.Ryczko D, Gratsch S, Schlager L, Keuyalian A, Boukhatem Z, Garcia C, Auclair F, Buschges A, Dubuc R. Nigral glutamatergic neurons control the speed of locomotion. J Neurosci 37: 9759‐9770, 2017.
 298.Santana MB, Halje P, Simplicio H, Richter U, Freire MAM, Petersson P, Fuentes R, Nicolelis MAL. Spinal cord stimulation alleviates motor deficits in a primate model of Parkinson disease. Neuron 84: 716‐722, 2014.
 299.Saunders A, Oldenburg IA, Berezovskii VK, Johnson CA, Kingery ND, Elliott HL, Xie T, Gerfen CR, Sabatini BL. A direct GABAergic output from the basal ganglia to frontal cortex. Nature 521: 85‐89, 2015.
 300.Schultz W. Multiple dopamine functions at different time courses. Annu Rev Neurosci 30: 259‐288, 2007.
 301.Schwab RS, Chafetz ME, Walker S. Control of two simultaneous voluntary motor acts in normals and Parkinsonism. Arch Neurol Psychol 72: 591‐598, 1954.
 302.Shen W, Flajolet M, Greengard P, Surmeier DJ. Dichotomous dopaminergic control of striatal synaptic plasticity. Science 321: 848‐851, 2008.
 303.Shen W, Plotkin JL, Francardo V, Ko WK, Xie Z, Li Q, Fieblinger T, Wess J, Neubig RR, Lindsley CW, Conn PJ, Greengard P, Bezard E, Cenci MA, Surmeier DJ. M4 muscarinic receptor signaling ameliorates striatal plasticity deficits in models of L‐DOPA‐induced dyskinesia. Neuron 88: 762‐773, 2015.
 304.Shepherd GM. Corticostriatal connectivity and its role in disease. Nat Rev Neurosci 14: 278‐291, 2013.
 305.Shiflett MW, Balleine BW. Contributions of ERK signaling in the striatum to instrumental learning and performance. Behav Brain Res 218: 240‐247, 2011.
 306.Shindou T, Ochi‐Shindou M, Wickens JR. A Ca(2+) threshold for induction of spike‐timing‐dependent depression in the mouse striatum. J Neurosci 31: 13015‐13022, 2011.
 307.Shindou T, Shindou M, Watanabe S, Wickens J. A silent eligibility trace enables dopamine‐dependent synaptic plasticity for reinforcement learning in the mouse striatum. Eur J Neurosci 49: 726‐736, 2019.
 308.Shonesy BC, Wang X, Rose KL, Ramikie TS, Cavener VS, Rentz T, Baucum AJ II, Jalan‐Sakrikar N, Mackie K, Winder DG, Patel S, Colbran RJ. CaMKII regulates diacylglycerol lipase‐alpha and striatal endocannabinoid signaling. Nat Neurosci 16: 456‐463, 2013.
 309.Smeets WJ, Marin O, Gonzalez A. Evolution of the basal ganglia: New perspectives through a comparative approach. J Anat 196 (Pt 4): 501‐517, 2000.
 310.Smith JB, Klug JR, Ross DL, Howard CD, Hollon NG, Ko VI, Hoffman H, Callaway EM, Gerfen CR, Jin X. Genetic‐based dissection unveils the inputs and outputs of striatal patch and matrix compartments. Neuron 91: 1069‐1084, 2016.
 311.Smith Y, Galvan A, Ellender TJ, Doig N, Villalba RM, Huerta‐Ocampo I, Wichmann T, Bolam JP. The thalamostriatal system in normal and diseased states. Front Syst Neurosci 8: 5, 2014.
 312.Stanley G, Gokce O, Malenka RC, Sudhof TC, Quake SR. Continuous and discrete neuron types of the adult murine striatum. Neuron 105: 688‐699, e688, 2020.
 313.Stein E, Bar‐Gad I. beta oscillations in the cortico‐basal ganglia loop during parkinsonism. Exp Neurol 245: 52‐59, 2013.
 314.Steiner H, Tseng KY. Handbook of Basal Ganglia Structure and Function. 2nd edition, London, Oxford, San Diego, Cambridge MA: Academic Press, 2017.
 315.Steiner LA, Barreda Tomas FJ, Planert H, Alle H, Vida I, Geiger JRP. Connectivity and dynamics underlying synaptic control of the subthalamic nucleus. J Neurosci 39: 2470‐2481, 2019.
 316.Stephenson‐Jones M, Ericsson J, Robertson B, Grillner S. Evolution of the basal ganglia: Dual‐output pathways conserved throughout vertebrate phylogeny. J Comp Neurol 520: 2957‐2973, 2012.
 317.Stephenson‐Jones M, Floros O, Robertson B, Grillner S. Evolutionary conservation of the habenular nuclei and their circuitry controlling the dopamine and 5‐hydroxytryptophan (5‐HT) systems. Proc Natl Acad Sci USA 109: E164‐E173, 2012.
 318.Stephenson‐Jones M, Kardamakis AA, Robertson B, Grillner S. Independent circuits in the basal ganglia for the evaluation and selection of actions. Proc Natl Acad Sci USA 110: E3670‐E3679, 2013.
 319.Stephenson‐Jones M, Samuelsson E, Ericsson J, Robertson B, Grillner S. Evolutionary conservation of the basal ganglia as a common vertebrate mechanism for action selection. Curr Biol 21: 1081‐1091, 2011.
 320.Stephenson‐Jones M, Yu K, Ahrens S, Tucciarone JM, van Huijstee AN, Mejia LA, Penzo MA, Tai LH, Wilbrecht L, Li B. A basal ganglia circuit for evaluating action outcomes. Nature 539: 289‐293, 2016.
 321.Strausfeld NJ, Hirth F. Deep homology of arthropod central complex and vertebrate basal ganglia. Science 340: 157‐161, 2013.
 322.Stuber GD, Hnasko TS, Britt JP, Edwards RH, Bonci A. Dopaminergic terminals in the nucleus accumbens but not the dorsal striatum corelease glutamate. J Neurosci 30: 8229‐8233, 2010.
 323.Surmeier DJ. Determinants of dopaminergic neuron loss in Parkinson's disease. FEBS J 285: 3657‐3668, 2018.
 324.Surmeier DJ, Graves SM, Shen W. Dopaminergic modulation of striatal networks in health and Parkinson's disease. Curr Opin Neurobiol 29: 109‐117, 2014.
 325.Surmeier DJ, Schumacker PT, Guzman JD, Ilijic E, Yang B, Zampese E. Calcium and Parkinson's disease. Biochem Biophys Res Commun 483: 1013‐1019, 2017.
 326.Suryanarayana SM, Hellgren Kotaleski J, Grillner S, Gurney KN. Roles for globus pallidus externa revealed in a computational model of action selection in the basal ganglia. Neural Netw 109: 113‐136, 2019.
 327.Suryanarayana SM, Pérez‐Fernández J, Robertson B, Grillner S. The evolutionary origin of visual and somatosensory representation in the vertebrate cortex. Nat Ecol Evol 4: 639‐651, 2020.
 328.Suryanarayana SM, Robertson B, Wallen P, Grillner S. The lamprey pallium provides a blueprint of the mammalian layered cortex. Curr Biol 27: 3264‐3277, e3265, 2017.
 329.Swanson LW. Understanding the Basic Plan. Oxford: Oxford University Press, 2003.
 330.Takakusaki K. Forebrain control of locomotor behaviors. Brain Res Rev 57: 192‐198, 2008.
 331.Takakusaki K, Oohinata‐Sugimoto J, Saitoh K, Habaguchi T. Role of basal ganglia‐brainstem systems in the control of postural muscle tone and locomotion. Prog Brain Res 143: 231‐237, 2004.
 332.Taverna S, Ilijic E, Surmeier DJ. Recurrent collateral connections of striatal medium spiny neurons are disrupted in models of Parkinson's disease. J Neurosci 28: 5504‐5512, 2008.
 333.Tecuapetla F, Jin X, Lima SQ, Costa RM. Complementary contributions of striatal projection pathways to action initiation and execution. Cell 166: 703‐715, 2016.
 334.Tecuapetla F, Patel JC, Xenias H, English D, Tadros I, Shah F, Berlin J, Deisseroth K, Rice ME, Tepper JM, Koos T. Glutamatergic signaling by mesolimbic dopamine neurons in the nucleus accumbens. J Neurosci 30: 7105‐7110, 2010.
 335.Tepper JM, Koos T, Ibanez‐Sandoval O, Tecuapetla F, Faust TW, Assous M. Heterogeneity and diversity of striatal GABAergic interneurons: Update 2018. Front Neuroanat 12: 91, 2018.
 336.Tepper JM, Tecuapetla F, Koos T, Ibanez‐Sandoval O. Heterogeneity and diversity of striatal GABAergic interneurons. Front Neuroanat 4: 150, 2010.
 337.Terao Y, Fukuda H, Ugawa Y, Hikosaka O. New perspectives on the pathophysiology of Parkinson's disease as assessed by saccade performance: A clinical review. Clin Neurophysiol 124: 1491‐1506, 2013.
 338.Thiele SL, Chen B, Lo C, Gertler TS, Warre R, Surmeier JD, Brotchie JM, Nash JE. Selective loss of bi‐directional synaptic plasticity in the direct and indirect striatal output pathways accompanies generation of parkinsonism and l‐DOPA induced dyskinesia in mouse models. Neurobiol Dis 71: 334‐344, 2014.
 339.Thompson RH, Menard A, Pombal M, Grillner S. Forebrain dopamine depletion impairs motor behavior in lamprey. Eur J Neurosci 27: 1452‐1460, 2008.
 340.Threlfell S, Lalic T, Platt NJ, Jennings KA, Deisseroth K, Cragg SJ. Striatal dopamine release is triggered by synchronized activity in cholinergic interneurons. Neuron 75: 58‐64, 2012.
 341.Tiklova K, Bjorklund AK, Lahti L, Fiorenzano A, Nolbrant S, Gillberg L, Volakakis N, Yokota C, Hilscher MM, Hauling T, Holmstrom F, Joodmardi E, Nilsson M, Parmar M, Perlmann T. Single‐cell RNA sequencing reveals midbrain dopamine neuron diversity emerging during mouse brain development. Nat Commun 10: 581, 2019.
 342.Tomkins A, Vasilaki E, Beste C, Gurney K, Humphries MD. Transient and steady‐state selection in the striatal microcircuit. Front Comput Neurosci 7: 192, 2013.
 343.Tritsch NX, Granger AJ, Sabatini BL. Mechanisms and functions of GABA co‐release. Nat Rev Neurosci 17: 139‐145, 2016.
 344.Ungerstedt U, Arbuthnott GW. Quantitative recording of rotational behavior in rats after 6‐hydroxy‐dopamine lesions of the nigrostriatal dopamine system. Brain Res 24: 485‐493, 1970.
 345.van der Kooy D, Fishell G. Neuronal birthdate underlies the development of striatal compartments. Brain Res 401: 155‐161, 1987.
 346.van der Meer MMA, Ito R, Lansink CS, Pennartz CMA. Hippocampal projections to the ventral striatum: From spatial memory to motivated behavior. In: Knierim JJ, Derdikman D, editors. Space, Time and Memory in the Hippocampal Formation. Vienna: Springer, 2014, p. 497‐516.
 347.van Keulen SC, Narzi D, Rothlisberger U. Association of both inhibitory and stimulatory Galpha subunits implies adenylyl cyclase 5 deactivation. Biochemistry 58: 4317‐4324, 2019.
 348.van Vulpen EH, van der Kooy D. Striatal cholinergic interneurons: Birthdates predict compartmental localization. Brain Res Dev Brain Res 109: 51‐58, 1998.
 349.Vicente AM, Galvao‐Ferreira P, Tecuapetla F, Costa RM. Direct and indirect dorsolateral striatum pathways reinforce different action strategies. Curr Biol 26: R267‐R269, 2016.
 350.Volkow ND, Morales M. The brain on drugs: From reward to addiction. Cell 162: 712‐725, 2015.
 351.von Twickel A, Kowatschew D, Salturk M, Schauer M, Robertson B, Korsching S, Walkowiak W, Grillner S, Perez‐Fernandez J. Individual dopaminergic neurons of lamprey SNc/VTA project to both the striatum and optic tectum but restrict co‐release of glutamate to striatum only. Curr Biol 29: 677‐685, e676, 2019.
 352.Wallace ML, Saunders A, Huang KW, Philson AC, Goldman M, Macosko EZ, McCarroll SA, Sabatini BL. Genetically distinct parallel pathways in the entopeduncular nucleus for limbic and sensorimotor output of the basal ganglia. Neuron 94: 138‐152, e135, 2017.
 353.Watabe‐Uchida M, Zhu L, Ogawa SK, Vamanrao A, Uchida N. Whole‐brain mapping of direct inputs to midbrain dopamine neurons. Neuron 74: 858‐873, 2012.
 354.Wolf JA, Moyer JT, Lazarewicz MT, Contreras D, Benoit‐Marand M, O'Donnell P, Finkel LH. NMDA/AMPA ratio impacts state transitions and entrainment to oscillations in a computational model of the nucleus accumbens medium spiny projection neuron. J Neurosci 25: 9080‐9095, 2005.
 355.Wolff SBE, Ko R, Ölveczky BP. Distinct roles for motor cortical and thalamic inputs to striatum during motor learning and execution. bioRxiv: 825810, 2019.
 356.Wu YW, Kim JI, Tawfik VL, Lalchandani RR, Scherrer G, Ding JB. Input‐ and cell‐type‐specific endocannabinoid‐dependent LTD in the striatum. Cell Rep 10: 75‐87, 2015.
 357.Xenias HS, Ibanez‐Sandoval O, Koos T, Tepper JM. Are striatal tyrosine hydroxylase interneurons dopaminergic? J Neurosci 35: 6584‐6599, 2015.
 358.Xiao L, Bornmann C, Hatstatt‐Burkle L, Scheiffele P. Regulation of striatal cells and goal‐directed behavior by cerebellar outputs. Nat Commun 9: 3133, 2018.
 359.Xie K, Masuho I, Shih CC, Cao Y, Sasaki K, Lai CW, Han PL, Ueda H, Dessauer CW, Ehrlich ME, Xu B, Willardson BM, Martemyanov KA. Stable G protein‐effector complexes in striatal neurons: Mechanism of assembly and role in neurotransmitter signaling. eLife 4: e10451, 2015.
 360.Yager LM, Garcia AF, Wunsch AM, Ferguson SM. The ins and outs of the striatum: Role in drug addiction. Neuroscience 301: 529‐541, 2015.
 361.Yagishita S, Hayashi‐Takagi A, Ellis‐Davies GC, Urakubo H, Ishii S, Kasai H. A critical time window for dopamine actions on the structural plasticity of dendritic spines. Science 345: 1616‐1620, 2014.
 362.Yamamoto K, Vernier P. The evolution of dopamine systems in chordates. Front Neuroanat 5: 21, 2011.
 363.Yamanaka K, Hori Y, Minamimoto T, Yamada H, Matsumoto N, Enomoto K, Aosaki T, Graybiel AM, Kimura M. Roles of centromedian parafascicular nuclei of thalamus and cholinergic interneurons in the dorsal striatum in associative learning of environmental events. J Neural Transm (Vienna) 125: 501‐513, 2018.
 364.Yapo C, Nair AG, Clement L, Castro LR, Hellgren Kotaleski J, Vincent P. Detection of phasic dopamine by D1 and D2 striatal medium spiny neurons. J Physiol 595: 7451‐7475, 2017.
 365.Yasuda M, Hikosaka O. Functional territories in primate substantia nigra pars reticulata separately signaling stable and flexible values. J Neurophysiol 113: 1681‐1696, 2015.
 366.Yin HH, Knowlton BJ. The role of the basal ganglia in habit formation. Nat Rev Neurosci 7: 464‐476, 2006.
 367.Zhai S, Tanimura A, Graves SM, Shen W, Surmeier DJ. Striatal synapses, circuits, and Parkinson's disease. Curr Opin Neurobiol 48: 9‐16, 2018.
 368.Zhang S, Qi J, Li X, Wang HL, Britt JP, Hoffman AF, Bonci A, Lupica CR, Morales M. Dopaminergic and glutamatergic microdomains in a subset of rodent mesoaccumbens axons. Nat Neurosci 18: 386‐392, 2015.
 369.Zhou FM. The substantia nigra pars reticulata. In: Steiner H, Tseng KY, editors. Handbook of Basal Ganglia Structure and Function. 2nd edition, London, Oxford, San Diego, Cambridge MA: Academic Press, 2017, p. 293‐316.

Contact Editor

Submit a note to the editor about this article by filling in the form below.

* Required Field

How to Cite

Sten Grillner, Brita Robertson, Jeanette Hellgren Kotaleski. Basal Ganglia—A Motion Perspective. Compr Physiol 2020, 10: 1241-1275. doi: 10.1002/cphy.c190045