Comprehensive Physiology Wiley Online Library

Functioning of Circuits Connecting Thalamus and Cortex

Full Article on Wiley Online Library



ABSTRACT

Glutamatergic pathways in thalamus and cortex are divided into two distinct classes: driver, which carries the main information between cells, and modulator, which modifies how driver inputs function. Identifying driver inputs helps to reveal functional computational circuits, and one set of such circuits identified by this approach are cortico‐thalamo‐cortical (or transthalamic corticocortical) circuits. This, in turn, leads to the conclusion that there are two types of thalamic relay: first order nuclei (such as the lateral geniculate nucleus) that relay driver input from a subcortical source (i.e., retina), and higher order nuclei (such as the pulvinar) which are involved in these transthalamic pathways by relaying driver input from layer 5 of one cortical area to another. This thalamic division is also seen in other sensory pathways and beyond these so that most of thalamus by volume consists of higher‐order relays. Many, and perhaps all, direct driver connections between cortical areas are paralleled by an indirect cortico‐thalamo‐cortical (transthalamic) driver route involving higher order thalamic relays. Such thalamic relays represent a heretofore unappreciated role in cortical functioning, and this assessment challenges and extends conventional views regarding both the role of thalamus and mechanisms of corticocortical communication. Finally, many and perhaps the vast majority of driver inputs relayed through thalamus arrive via branching axons, with extrathalamic targets often being subcortical motor centers. This raises the possibility that inputs relayed by thalamus to cortex also serve as efference copies, and this may represent an important feature of information relayed up the cortical hierarchy via transthalamic circuits. © 2017 American Physiological Society. Compr Physiol 7:713‐739, 2017.

Comprehensive Physiology offers downloadable PowerPoint presentations of figures for non-profit, educational use, provided the content is not modified and full credit is given to the author and publication.

Download a PowerPoint presentation of all images


Figure 1. Figure 1. Schematic three‐dimensional view of right thalamus with many of its major nuclei. A cut is placed in the posterior part to reveal a representative cross‐section. To prevent obscuring the dorsal thalamus, only the rostral tip of the TRN is shown. Abbreviations: A, anterior nucleus; CM, centromedian nucleus; IL, intralaminar nuclei; IML, internal medullary lamina; LD, lateral dorsal nucleus; LP, lateral posterior nucleus; LGN, lateral geniculate nucleus; MGN, medial geniculate nucleus; MD, mediodorsal nucleus; MI, midline nuclei; P, pulvinar; PO, posterior nucleus; TRN, thalamic reticular nucleus; VA, ventral anterior nucleus; VPl, ventral posterolateral nucleus; VPm, ventral posteromedial nucleus. See Jones (75) for details of connectivity of these nuclei. Redrawn, with permission, from (149).
Figure 2. Figure 2. Reconstruction of representative thalamic cells types from the lateral geniculate nucleus of the cat based on intracellular dye filling of individual, physiologically identified neurons. (A) Relay X and Y cell. (B) Interneuron. The inset shows the presynaptic bouton terminals emanating from dendrites. (C) Cell of the thalamic reticular nucleus. The larger scale applies to the inset for the interneuron. A and B, with permission, from (46); and C, with permission, from (166).
Figure 3. Figure 3. (A) Schematic and simplified view of thalamic circuitry. The various inputs to the different thalamic cell types are displayed, and the excitatory or inhibitory postsynaptic effect. (B and C) Schematic view of different possible circuits involving layer 6 corticothalamic input to reticular cells, interneurons, and relay cells. See text for details. Abbreviations: 5‐HT, serotonin; ACh, acetylcholine; BRF, brainstem reticular formation; GABA, γ‐aminobutyric acid; Glu, glutamate; NA, noradrenaline; TRN, thalamic reticular nucleus.
Figure 4. Figure 4. Schematic views of features of connectivity in the A layers of the cat's lateral geniculate nucleus. (A) Synaptic inputs in and near a glomerulus. Shown are the various synaptic contacts (arrows), whether they are inhibitory or excitatory, and the related postsynaptic receptors. The conventional triad includes the lower interneuronal dendritic terminal and involves three synapses (from the retinal terminal to the dendritic terminal, from the retinal terminal to an appendage of the X cell dendrite, and from the dendritic terminal to the same appendage). Another type of triad includes the upper interneuronal dendritic terminal and also involves three synapses: a branched (cholinergic) brainstem axon produces one synaptic terminal onto an X cell relay dendrite and another onto the dendritic terminal, and a third synapse is formed from the dendritic terminal onto the same relay cell dendrite. For simplicity, the NMDA receptor on the relay cell postsynaptic to the retinal input has been left off. (B) Synaptic inputs onto an X and a Y cell. For simplicity, only one, unbranched dendrite is shown. Synaptic types are shown in relative numbers. Abbreviations: ACh, acetylcholine; AMPAR, (R,S)‐α‐amino‐3‐hydroxy‐5‐methyl‐4‐isoxazolepropionic acid receptor; GABA, γ‐aminobutyric acid; GABAAR, type A receptor for GABA; Glu, glutamate; M1R and M2R, two types of muscarinic receptor; mGluR5, type 5 metabotropic glutamate receptor; NicR, nicotinic receptor. Redrawn, with permission, from (148).
Figure 5. Figure 5. Overview of circuitry of LGN. (A and B) Detailed circuitry for X and Y relay cells of the LGN of the cat. The inhibitory inputs from axons of interneurons to relay cells is of the opposite center/surround type. For the center/surround receptive field icons, plusses refer to on areas, and minuses, to off areas. Redrawn, with permission, from (148). Abbreviations: I, interneuron; LGN, lateral geniculate nucleus; R, LGN relay cell; TRN, thalamic reticular nucleus.
Figure 6. Figure 6. Burst and tonic firing based on IT properties. Adapted, with permission, from (141). (A) IT becomes inactivated following ≥∼100 ms of membrane depolarization relative to about −60 to 65 mV, and the inactivation is removed (or, IT is deinactivated) by an equivalent time of relative hyperpolarization. (A) Tonic firing results when IT is inactivated. (B) Burst firing results when IT is deinactivated. (C) Input–output relationship for a single cell. The abscissa is the amplitude of the depolarizing current pulse and the ordinate is the firing frequency of the cell. The firing frequency was determined by the first six action potentials of the response, burst or tonic, because this cell normally exhibited six action potentials per burst in this experiment. The initial holding potentials are shown: −47 and −59 mV reflects tonic mode (blue points and curves), whereas −77 and −83 mV reflects burst mode (red points and curves).
Figure 7. Figure 7. Different receptive field properties involved in responses of geniculate relay cells. Those of the retinal input display the classic center/surround structure, and these are monocularly represented. Those of the layer 6 cortical input have more complex features, including orientation and direction selectivity, and these are binocularly represented. Those of the geniculate cell have features much like those of its retinal input and unlike its cortical input.
Figure 8. Figure 8. Drivers and modulators. (A) Modulators (dashed green inputs) shown contacting more peripheral dendrites than do drivers (solid green input). Also, drivers activate only ionotropic glutamate receptors, whereas modulators also activate metabotropic glutamate receptors. (B) Effects of repetitive stimulation on EPSP amplitude: for modulators, this produces paired pulse facilitation (increasing EPSP amplitudes during the stimulus train), whereas for drivers, this produces paired pulse depression (decreasing EPSP amplitudes during the stimulus train). Also, increasing stimulus intensity for modulators (shown as different colors) increases EPSPs more than is the case for drivers; this indicates more convergence of modulator inputs compared to driver inputs. (C) Light microscopic tracings of a driver afferent (a retinogeniculate axon from the cat) and a modulator afferent (a corticogeniculate axon from layer 6 of the cat). Redrawn, with permission, from (147). (D) Three‐dimensional scatterplot for inputs classified as driver or modulator to cells of thalamus and cortex; data from in vitro slice experiments in mice from the author's laboratory. The three parameters are: (1) the amplitude of the first EPSP elicited in a train at a stimulus level just above threshold; (2) a measure of paired‐pulse effects (the amplitude of the second EPSP divided by the first) for stimulus trains of 10 to 20 Hz; and (3) a measure of the response to synaptic activation of metabotropic glutamate receptors, taken as the maximum voltage deflection (i.e., depolarization or hyperpolarization) during the 300 ms postsynaptic response period to tetanic stimulation in the presence of AMPA and NMDA blockers. Pathways tested here include various inputs to thalamus from cortex and subcortical sources, various thalamocortical pathways, and various intracortical pathways. Adapted, with permission, from (145).
Figure 9. Figure 9. Schematic summary of synaptic effects of layer 6 corticothalamic cells on thalamocortical transmission. Note that these cells have bifurcating axons that innervate both layer 4 cells postsynaptic to thalamic input as well as thalamic circuitry. Four distinct effects have been documented, each of which serves to reduce the gain of thalamocortical transmission. See text for details.
Figure 10. Figure 10. Schematic diagrams showing organizational features of first and higher order thalamic nuclei. A first‐order nucleus (left) represents the first relay of a particular type of subcortical information to a first‐order or primary cortical area. A higher‐order nucleus (center and right) relays information from layer 5 of one cortical area up the hierarchy to another cortical area. This relay can be from a primary area to a higher one (center) or between two higher‐order cortical areas (right). The important difference between first‐ and higher‐order nuclei is the driver input, which is subcortical for a first‐order relay and from layer 5 of cortex for a higher‐order relay. Note that all thalamic nuclei receive an input from layer 6 of cortex, which is mostly organized in a reciprocal feedback manner, but higher‐order nuclei in addition receive a layer 5 input from cortex, which is feedforward. Note that the driver inputs, both subcortical and from layer 5, are typically from branching axons, with some extrathalamic targets being subcortical motor centers, and the significance this is elaborated in the text. Abbreviations: BRF, brainstem reticular formation; FO, first order; HO, higher order; TRN, thalamic reticular nucleus. Redrawn, with permission, from (143).
Figure 11. Figure 11. Two views of tectothalamic inputs in auditory pathways. Shown are projections from the core region of the inferior colliculus (ICc) to the ventral part of the medial geniculate nucleus (MGNv) and from the shell region of the inferior colliculus (ICs) to the dorsal part of the medial geniculate nucleus (MGNd). (A) View of parallel processing. Two information streams are shown, one from ICc through MGNv to primary auditory cortex (A1) and the other from ICs through MGNd to secondary auditory cortex (A2). (B) View incorporating identification of drivers and modulators. The stream from ICc through MGNv involves driver paths and thus represents an information stream. However, the input from ICs to MGNd is a modulator, whereas the driver input to MGNd arises from layer 5 of cortex. Thus, in this view, the projection from ICs to MGNd modulates this transthalamic circuit.
Figure 12. Figure 12. Two views of the relationship between the basal ganglia and cortex. (A) Textbook view. This depicts a simple information loop, the information flowing from thalamus to cortex to basal ganglia to thalamus, etc. Adapted, with permission, from (78). (B) Different view incorporating the idea that information is carried by glutamatergic driver pathways. Since the basal ganglia outputs are strictly GABAergic, this input to thalamus serves not as an information route but rather modulates a transthalamic pathway through the higher‐order portion of the motor thalamus. When active, the basal ganglia input would shut down the higher‐order thalamic relays, providing a gating mechanism. (C) One example of this is that the basal ganglia input to thalamus can determine which combinations of transthalamic pathways are active at any given time; dashed thalamocortical pathways indicate that these are nonactive due to basal ganglia inhibition. (D) A related example is that the basal ganglia can determine which cortical areas are actively connected by both direct and transthalamic pathways, and not just the former. See text for details.
Figure 13. Figure 13. Branching driver inputs to representative first order thalamic relays. (A) Example from retinogeniculate axon of cat; redrawn, with permission, from (163). (B) Example of cerebellar inputs to the ventral anterior and ventral lateral nuclei (VA/VL); redrawn, with permission, from (19), and thanks to Javier deFelipe for providing this image. (C) Example of mammillary inputs to the anterior dorsal nucleus (AD); redrawn, with permission, from (81). Red arrows in B and C indicate branch points.
Figure 14. Figure 14. Example from layer 5 pyramidal tract cell of rat motor cortex; redrawn, with permission, from (80); tracing of reconstruction generously supplied by H. Kita. Branches innervating thalamus are indicated by the dashed blue circle, and brainstem motor regions are indicated by red arrows. Abbreviations: cp, cerebral peduncle; DpMe, deep mesencephalic nuclei; Gi, gigantocellular reticular nucleus; GPe, Globus pallidus external segment; ic, internal capsule; IO, inferior olive; Pn, pontine nucleus; PnO, pontine reticular nucleus, oral part; py, medullary pyramid; pyd, pyramidal decussation; Rt, thalamic reticular nucleus; SC, superior colliculus; SN, substantia nigra; Str, striatum; VL, ventrolateral thalamic nucleus; VM, ventromedial thalamic nucleus.
Figure 15. Figure 15. Branching axons. (A) Cajal illustration (19) of primary axons entering the spinal cord and branching to innervate the spinal gray matter and brain areas. The red arrows indicate branch points. Thanks to Javier deFelipe for providing this image. (B) Schematic interpretation of A.
Figure 16. Figure 16. Comparison of conventional view (A) with the alternative view proposed here (B). The question marks in A indicate higher order thalamic relays, for which no specific function is suggested in this scheme. Further details in text. Abbreviations: FO, first order; HO, higher order.


Figure 1. Schematic three‐dimensional view of right thalamus with many of its major nuclei. A cut is placed in the posterior part to reveal a representative cross‐section. To prevent obscuring the dorsal thalamus, only the rostral tip of the TRN is shown. Abbreviations: A, anterior nucleus; CM, centromedian nucleus; IL, intralaminar nuclei; IML, internal medullary lamina; LD, lateral dorsal nucleus; LP, lateral posterior nucleus; LGN, lateral geniculate nucleus; MGN, medial geniculate nucleus; MD, mediodorsal nucleus; MI, midline nuclei; P, pulvinar; PO, posterior nucleus; TRN, thalamic reticular nucleus; VA, ventral anterior nucleus; VPl, ventral posterolateral nucleus; VPm, ventral posteromedial nucleus. See Jones (75) for details of connectivity of these nuclei. Redrawn, with permission, from (149).


Figure 2. Reconstruction of representative thalamic cells types from the lateral geniculate nucleus of the cat based on intracellular dye filling of individual, physiologically identified neurons. (A) Relay X and Y cell. (B) Interneuron. The inset shows the presynaptic bouton terminals emanating from dendrites. (C) Cell of the thalamic reticular nucleus. The larger scale applies to the inset for the interneuron. A and B, with permission, from (46); and C, with permission, from (166).


Figure 3. (A) Schematic and simplified view of thalamic circuitry. The various inputs to the different thalamic cell types are displayed, and the excitatory or inhibitory postsynaptic effect. (B and C) Schematic view of different possible circuits involving layer 6 corticothalamic input to reticular cells, interneurons, and relay cells. See text for details. Abbreviations: 5‐HT, serotonin; ACh, acetylcholine; BRF, brainstem reticular formation; GABA, γ‐aminobutyric acid; Glu, glutamate; NA, noradrenaline; TRN, thalamic reticular nucleus.


Figure 4. Schematic views of features of connectivity in the A layers of the cat's lateral geniculate nucleus. (A) Synaptic inputs in and near a glomerulus. Shown are the various synaptic contacts (arrows), whether they are inhibitory or excitatory, and the related postsynaptic receptors. The conventional triad includes the lower interneuronal dendritic terminal and involves three synapses (from the retinal terminal to the dendritic terminal, from the retinal terminal to an appendage of the X cell dendrite, and from the dendritic terminal to the same appendage). Another type of triad includes the upper interneuronal dendritic terminal and also involves three synapses: a branched (cholinergic) brainstem axon produces one synaptic terminal onto an X cell relay dendrite and another onto the dendritic terminal, and a third synapse is formed from the dendritic terminal onto the same relay cell dendrite. For simplicity, the NMDA receptor on the relay cell postsynaptic to the retinal input has been left off. (B) Synaptic inputs onto an X and a Y cell. For simplicity, only one, unbranched dendrite is shown. Synaptic types are shown in relative numbers. Abbreviations: ACh, acetylcholine; AMPAR, (R,S)‐α‐amino‐3‐hydroxy‐5‐methyl‐4‐isoxazolepropionic acid receptor; GABA, γ‐aminobutyric acid; GABAAR, type A receptor for GABA; Glu, glutamate; M1R and M2R, two types of muscarinic receptor; mGluR5, type 5 metabotropic glutamate receptor; NicR, nicotinic receptor. Redrawn, with permission, from (148).


Figure 5. Overview of circuitry of LGN. (A and B) Detailed circuitry for X and Y relay cells of the LGN of the cat. The inhibitory inputs from axons of interneurons to relay cells is of the opposite center/surround type. For the center/surround receptive field icons, plusses refer to on areas, and minuses, to off areas. Redrawn, with permission, from (148). Abbreviations: I, interneuron; LGN, lateral geniculate nucleus; R, LGN relay cell; TRN, thalamic reticular nucleus.


Figure 6. Burst and tonic firing based on IT properties. Adapted, with permission, from (141). (A) IT becomes inactivated following ≥∼100 ms of membrane depolarization relative to about −60 to 65 mV, and the inactivation is removed (or, IT is deinactivated) by an equivalent time of relative hyperpolarization. (A) Tonic firing results when IT is inactivated. (B) Burst firing results when IT is deinactivated. (C) Input–output relationship for a single cell. The abscissa is the amplitude of the depolarizing current pulse and the ordinate is the firing frequency of the cell. The firing frequency was determined by the first six action potentials of the response, burst or tonic, because this cell normally exhibited six action potentials per burst in this experiment. The initial holding potentials are shown: −47 and −59 mV reflects tonic mode (blue points and curves), whereas −77 and −83 mV reflects burst mode (red points and curves).


Figure 7. Different receptive field properties involved in responses of geniculate relay cells. Those of the retinal input display the classic center/surround structure, and these are monocularly represented. Those of the layer 6 cortical input have more complex features, including orientation and direction selectivity, and these are binocularly represented. Those of the geniculate cell have features much like those of its retinal input and unlike its cortical input.


Figure 8. Drivers and modulators. (A) Modulators (dashed green inputs) shown contacting more peripheral dendrites than do drivers (solid green input). Also, drivers activate only ionotropic glutamate receptors, whereas modulators also activate metabotropic glutamate receptors. (B) Effects of repetitive stimulation on EPSP amplitude: for modulators, this produces paired pulse facilitation (increasing EPSP amplitudes during the stimulus train), whereas for drivers, this produces paired pulse depression (decreasing EPSP amplitudes during the stimulus train). Also, increasing stimulus intensity for modulators (shown as different colors) increases EPSPs more than is the case for drivers; this indicates more convergence of modulator inputs compared to driver inputs. (C) Light microscopic tracings of a driver afferent (a retinogeniculate axon from the cat) and a modulator afferent (a corticogeniculate axon from layer 6 of the cat). Redrawn, with permission, from (147). (D) Three‐dimensional scatterplot for inputs classified as driver or modulator to cells of thalamus and cortex; data from in vitro slice experiments in mice from the author's laboratory. The three parameters are: (1) the amplitude of the first EPSP elicited in a train at a stimulus level just above threshold; (2) a measure of paired‐pulse effects (the amplitude of the second EPSP divided by the first) for stimulus trains of 10 to 20 Hz; and (3) a measure of the response to synaptic activation of metabotropic glutamate receptors, taken as the maximum voltage deflection (i.e., depolarization or hyperpolarization) during the 300 ms postsynaptic response period to tetanic stimulation in the presence of AMPA and NMDA blockers. Pathways tested here include various inputs to thalamus from cortex and subcortical sources, various thalamocortical pathways, and various intracortical pathways. Adapted, with permission, from (145).


Figure 9. Schematic summary of synaptic effects of layer 6 corticothalamic cells on thalamocortical transmission. Note that these cells have bifurcating axons that innervate both layer 4 cells postsynaptic to thalamic input as well as thalamic circuitry. Four distinct effects have been documented, each of which serves to reduce the gain of thalamocortical transmission. See text for details.


Figure 10. Schematic diagrams showing organizational features of first and higher order thalamic nuclei. A first‐order nucleus (left) represents the first relay of a particular type of subcortical information to a first‐order or primary cortical area. A higher‐order nucleus (center and right) relays information from layer 5 of one cortical area up the hierarchy to another cortical area. This relay can be from a primary area to a higher one (center) or between two higher‐order cortical areas (right). The important difference between first‐ and higher‐order nuclei is the driver input, which is subcortical for a first‐order relay and from layer 5 of cortex for a higher‐order relay. Note that all thalamic nuclei receive an input from layer 6 of cortex, which is mostly organized in a reciprocal feedback manner, but higher‐order nuclei in addition receive a layer 5 input from cortex, which is feedforward. Note that the driver inputs, both subcortical and from layer 5, are typically from branching axons, with some extrathalamic targets being subcortical motor centers, and the significance this is elaborated in the text. Abbreviations: BRF, brainstem reticular formation; FO, first order; HO, higher order; TRN, thalamic reticular nucleus. Redrawn, with permission, from (143).


Figure 11. Two views of tectothalamic inputs in auditory pathways. Shown are projections from the core region of the inferior colliculus (ICc) to the ventral part of the medial geniculate nucleus (MGNv) and from the shell region of the inferior colliculus (ICs) to the dorsal part of the medial geniculate nucleus (MGNd). (A) View of parallel processing. Two information streams are shown, one from ICc through MGNv to primary auditory cortex (A1) and the other from ICs through MGNd to secondary auditory cortex (A2). (B) View incorporating identification of drivers and modulators. The stream from ICc through MGNv involves driver paths and thus represents an information stream. However, the input from ICs to MGNd is a modulator, whereas the driver input to MGNd arises from layer 5 of cortex. Thus, in this view, the projection from ICs to MGNd modulates this transthalamic circuit.


Figure 12. Two views of the relationship between the basal ganglia and cortex. (A) Textbook view. This depicts a simple information loop, the information flowing from thalamus to cortex to basal ganglia to thalamus, etc. Adapted, with permission, from (78). (B) Different view incorporating the idea that information is carried by glutamatergic driver pathways. Since the basal ganglia outputs are strictly GABAergic, this input to thalamus serves not as an information route but rather modulates a transthalamic pathway through the higher‐order portion of the motor thalamus. When active, the basal ganglia input would shut down the higher‐order thalamic relays, providing a gating mechanism. (C) One example of this is that the basal ganglia input to thalamus can determine which combinations of transthalamic pathways are active at any given time; dashed thalamocortical pathways indicate that these are nonactive due to basal ganglia inhibition. (D) A related example is that the basal ganglia can determine which cortical areas are actively connected by both direct and transthalamic pathways, and not just the former. See text for details.


Figure 13. Branching driver inputs to representative first order thalamic relays. (A) Example from retinogeniculate axon of cat; redrawn, with permission, from (163). (B) Example of cerebellar inputs to the ventral anterior and ventral lateral nuclei (VA/VL); redrawn, with permission, from (19), and thanks to Javier deFelipe for providing this image. (C) Example of mammillary inputs to the anterior dorsal nucleus (AD); redrawn, with permission, from (81). Red arrows in B and C indicate branch points.


Figure 14. Example from layer 5 pyramidal tract cell of rat motor cortex; redrawn, with permission, from (80); tracing of reconstruction generously supplied by H. Kita. Branches innervating thalamus are indicated by the dashed blue circle, and brainstem motor regions are indicated by red arrows. Abbreviations: cp, cerebral peduncle; DpMe, deep mesencephalic nuclei; Gi, gigantocellular reticular nucleus; GPe, Globus pallidus external segment; ic, internal capsule; IO, inferior olive; Pn, pontine nucleus; PnO, pontine reticular nucleus, oral part; py, medullary pyramid; pyd, pyramidal decussation; Rt, thalamic reticular nucleus; SC, superior colliculus; SN, substantia nigra; Str, striatum; VL, ventrolateral thalamic nucleus; VM, ventromedial thalamic nucleus.


Figure 15. Branching axons. (A) Cajal illustration (19) of primary axons entering the spinal cord and branching to innervate the spinal gray matter and brain areas. The red arrows indicate branch points. Thanks to Javier deFelipe for providing this image. (B) Schematic interpretation of A.


Figure 16. Comparison of conventional view (A) with the alternative view proposed here (B). The question marks in A indicate higher order thalamic relays, for which no specific function is suggested in this scheme. Further details in text. Abbreviations: FO, first order; HO, higher order.
References
 1.Ahmed B, Anderson JC, Douglas RJ, Martin KAC, Nelson JC. Polyneuronal innervation of spiny stellate neurons in cat visual cortex. J Comp Neurol 341: 39‐49, 1994.
 2.Alitto HJ, Usrey WM. Corticothalamic feedback and sensory processing. Curr Opin Neurobiol 13: 440‐445, 2003.
 3.Allen NJ. Astrocyte regulation of synaptic behavior. Annu Rev Cell Dev Biol 30: 439‐463, 2014.
 4.Andersen RA, Cui H. Intention, action planning, and decision making in parietal‐frontal circuits. Neuron 63: 568‐583, 2009.
 5.Andolina IM, Jones HE, Sillito AM. Effects of cortical feedback on the spatial properties of relay cells in the lateral geniculate nucleus. J Neurophysiol 109: 889‐899, 2013.
 6.Arcelli P, Frassoni C, Regondi MC, De Biasi S, Spreafico R. GABAergic neurons in mammalian thalamus: A marker of thalamic complexity? Brain Res Bull 42: 27‐37, 1997.
 7.Baker FH, Malpeli JG. Effects of cryogenic blockade of visual cortex on the responses of lateral geniculate neurons in the monkey. Exp Brain Res 29: 433‐444, 1977.
 8.Bender DB. Visual activation of neurons in the primate pulvinar depends on cortex but not colliculus. Brain Res 279: 258‐261, 1983.
 9.Berman RA, Wurtz RH. Functional identification of a pulvinar path from superior colliculus to cortical area MT. J Neurosci 30: 6342‐6354, 2010.
 10.Bickford ME, Zhou N, Krahe TE, Govindaiah G, Guido W. Retinal and tectal “driver‐Like” inputs converge in the shell of the mouse dorsal lateral geniculate nucleus. J Neurosci 35: 10523‐10534, 2015.
 11.Bokor H, Frere SGA, Eyre MD, Slezia A, Ulbert I, Luthi A, Acsády L. Selective GABAergic control of higher‐order thalamic relays. Neuron 45: 929‐940, 2005.
 12.Bourassa J, Deschênes M. Corticothalamic projections from the primary visual cortex in rats: A single fiber study using biocytin as an anterograde tracer. Neurosci 66: 253‐263, 1995.
 13.Bourassa J, Pinault D, Deschênes M. Corticothalamic projections from the cortical barrel field to the somatosensory thalamus in rats: A single‐fibre study using biocytin as an anterograde tracer. Eur J Neurosci 7: 19‐30, 1995.
 14.Brickman AM, Buchsbaum MS, Shihabuddin L, Byne W, Newmark RE, Brand J, Ahmed S, Mitelman SA, Hazlett EA. Thalamus size and outcome in schizophrenia. Schizophr Res 71: 473‐484, 2004.
 15.Briggs F, Kiley CW, Callaway EM, Usrey WM. Morphological substrates for parallel streams of corticogeniculate feedback originating in both V1 and V2 of the macaque monkey. Neuron 90: 388‐399, 2016.
 16.Briggs F, Mangun GR, Usrey WM. Attention enhances synaptic efficacy and the signal‐to‐noise ratio in neural circuits. Nature 499: 476‐480, 2013.
 17.Briggs F, Usrey WM. Parallel processing in the corticogeniculate pathway of the macaque monkey. Neuron 62: 135‐146, 2009.
 18.Bull MS, Berkley KJ. Differences in the neurones that project from the dorsal column nuclei to the diencephalon, pretectum, and the tectum in the cat. Somatosens Res 1: 281‐300, 1984.
 19.Cajal SRy. Histologie du Système Nerveaux de l'Homme et des Vertébrés. Paris: Maloine, 1911.
 20.Carandini M. Receptive fields and suppressive fields in the early visual system. In: Gazzaniga MS, editor. The Cognitive Neurosciences. Cambridge, MA: MIT Press, 2004, p. 313‐326.
 21.Casanova C. Response properties of neurons in area 17 projecting to the striate‐recipient zone of the cat's lateralis posterior‐pulvinar complex: Comparison with cortico‐tectal cells. Exp Brain Res 96: 247‐259, 1993.
 22.Catterall WA. Ion channel voltage sensors: Structure, function, and pathophysiology. Neuron 67: 915‐928, 2010.
 23.Chalupa LM, Anchel H, Lindsley DB. Visual input to the pulvinar via lateral geniculate, superior colliculus and visual cortex in the cat. Exp Neurol 36: 449‐462, 1972.
 24.Chalupa LM, Thompson ID. Retinal ganglion cell projections to the superior colliculus of the hamster demonstrated by the horseradish peroxidase technique. Neurosci Lett 19: 13‐19, 1980.
 25.Chance FS, Abbott LF, Reyes A. Gain modulation from background synaptic input. Neuron 35: 773‐782, 2002.
 26.Chauveau F, Celerier A, Ognard R, Pierard C, Beracochea D. Effects of ibotenic acid lesions of the mediodorsal thalamus on memory: Relationship with emotional processes in mice. Behav Brain Res 156: 215‐223, 2005.
 27.Coogan TA, Burkhalter A. Hierarchical organization of areas in rat visual cortex. J Neurosci 13: 3749‐3772, 1993.
 28.Coscia DM, Narr KL, Robinson DG, Hamilton LS, Sevy S, Burdick KE, GunduzBruce H, McCormack J, Bilder RM, Szeszko PR. Volumetric and shape analysis of the thalamus in first‐episode schizophrenia. Hum Brain Mapp 30: 1236‐1245, 2009.
 29.Covic EN, Sherman SM. Synaptic properties of connections between the primary and secondary auditory cortices in mice. Cereb Cortex 21: 2425‐2441, 2011.
 30.Cox CL, Sherman SM. Control of dendritic outputs of inhibitory interneurons in the lateral geniculate nucleus. Neuron 27: 597‐610, 2000.
 31.Crandall SR, Cruikshank SJ, Connors BW. A corticothalamic switch: Controlling the thalamus with dynamic synapses. Neuron 86: 768‐782, 2015.
 32.Cronenwett WJ, Csernansky J. Thalamic pathology in schizophrenia. Curr Top Behav Neurosci 4: 509‐528, 2010.
 33.Dankowski A, Bickford ME. Inhibitory circuitry involving Y cells and Y retinal terminals in the C laminae of the cat dorsal lateral geniculate nucleus. J Comp Neurol 460: 368‐379, 2003.
 34.Danos P, Baumann B, Kramer A, Bernstein HG, Stauch R, Krell D, Falkai P, Bogerts B. Volumes of association thalamic nuclei in schizophrenia: A postmortem study. Schizophr Res 60: 141‐155, 2003.
 35.DePasquale R, Sherman SM. Synaptic properties of corticocortical connections between the primary and secondary visual cortical areas in the mouse. J Neurosci 31: 16494‐16506, 2011.
 36.Deschênes M, Bourassa J, Pinault D. Corticothalamic projections from layer V cells in rat are collaterals of long‐range corticofugal axons. Brain Res 664: 215‐219, 1994.
 37.Diamond ME, Armstrong‐James M, Ebner FF. Somatic sensory responses in the rostral sector of the posterior group (POm) and in the ventral posterior medial nucleus (VPM) of the rat thalamus. J Comp Neurol 318: 462‐476, 1992.
 38.Donishi T, Kimura A, Imbe H, Yokoi I, Kaneoke Y. Sub‐threshold cross‐modal sensory interaction in the thalamus: Lemniscal auditory response in the medial geniculate nucleus is modulated by somatosensory stimulation. Neurosci 174: 200‐215, 2011.
 39.DorphPetersen KA, Caric D, Saghafi R, Zhang W, Sampson AR, Lewis DA. Volume and neuron number of the lateral geniculate nucleus in schizophrenia and mood disorders. Acta Neuropathol 117: 369‐384, 2009.
 40.Duhamel J‐R, Colby CL, Goldberg ME. The updating of the representation of visual space in parietal cortex by intended eye movements. Science 255: 90‐92, 1992.
 41.EriΠir A, Van Horn SC, Bickford ME, Sherman SM. Immunocytochemistry and distribution of parabrachial terminals in the lateral geniculate nucleus of the cat: A comparison with corticogeniculate terminals. J Comp Neurol 377: 535‐549, 1997.
 42.EriΠir A, Van Horn SC, Sherman SM. Relative numbers of cortical and brainstem inputs to the lateral geniculate nucleus. Proc Natl Acad Sci U S A 94: 1517‐1520, 1997.
 43.Felleman DJ, Van Essen DC. Distributed hierarchical processing in the primate cerebral cortex. Cereb Cortex 1: 1‐47, 1991.
 44.Ford JM, Palzes VA, Roach BJ, Mathalon DH. Did I do that? Abnormal predictive processes in schizophrenia when button pressing to deliver a tone. Schizophr Bull 40: 804‐812, 2014.
 45.Ford JM, Roach BJ, Faustman WO, Mathalon DH. Out‐of‐synch and out‐of‐sorts: Dysfunction of motor‐sensory communication in schizophrenia. Biol Psychiatry 63: 736‐743, 2008.
 46.Friedlander MJ, Lin C‐S, Stanford LR, Sherman SM. Morphology of functionally identified neurons in lateral geniculate nucleus of the cat. J Neurophysiol 46: 80‐129, 1981.
 47.Fries P. Neuronal gamma‐band synchronization as a fundamental process in cortical computation. Annu Rev Neurosci 32: 209‐224, 2009.
 48.Fries P, Nikolic D, Singer W. The gamma cycle. Trends Neurosci 30: 309‐316, 2007.
 49.Geisert EE, Langsetmo A, Spear PD. Influence of the cortico‐geniculate pathway on response properties of cat lateral geniculate neurons. Brain Res 208: 409‐415, 1981.
 50.Govindaiah, Cox CL. Synaptic activation of metabotropic glutamate receptors regulates dendritic outputs of thalamic interneurons. Neuron 41: 611‐623, 2004.
 51.Govindaiah G, Wang T, Gillette MU, Cox CL. Activity‐dependent regulation of retinogeniculate signaling by metabotropic glutamate receptors. J Neurosci 32: 12820‐12831, 2012.
 52.Gregoriou GG, Gotts SJ, Zhou H, Desimone R. High‐frequency, long‐range coupling between prefrontal and visual cortex during attention. Science 324: 1207‐1210, 2009.
 53.Groh A, Bokor H, Mease RA, Plattner VM, Hangya B, Stroh A, Deschenes M, Acsady L. Convergence of cortical and sensory driver inputs on single thalamocortical cells. Cereb Cortex 24: 3167‐3179, 2014.
 54.Groh A, Bokor H, Mease RA, Plattner VM, Hangya B, Stroh A, Deschenes M, Acsady L. Convergence of cortical and sensory driver inputs on single thalamocortical cells. Cereb Cortex 24: 3167‐3179, 2014.
 55.Guillery RW. Degeneration in the hypothalamic connexions of the albino rat. J Anat (Lond) 91: 91‐115, 1957.
 56.Guillery RW. Anatomical evidence concerning the role of the thalamus in corticocortical communication: A brief review. J Anat 187: 583‐592, 1995.
 57.Guillery RW, Sherman SM. Branched thalamic afferents: What are the messages that they relay to cortex? Brain Res Brain Res Rev 66: 205‐219, 2011.
 58.Gulcebi MI, Ketenci S, Linke R, Hacioglu H, Yanali H, Veliskova J, Moshe SL, Onat F, Cavdar S. Topographical connections of the substantia nigra pars reticulata to higher‐order thalamic nuclei in the rat. Brain Res Bull 87: 312‐318, 2012.
 59.Hall NJ, Colby CL. Remapping for visual stability. Philos Trans R Soc Lond B Biol Sci 366: 528‐539, 2011.
 60.Hamos JE, Van Horn SC, Raczkowski D, Sherman SM. Synaptic circuits involving an individual retinogeniculate axon in the cat. J Comp Neurol 259: 165‐192, 1987.
 61.Hamos JE, Van Horn SC, Raczkowski D, Uhlrich DJ, Sherman SM. Synaptic connectivity of a local circuit neurone in lateral geniculate nucleus of the cat. Nature 317: 618‐621, 1985.
 62.Helmholtz H. Handbuch der Physiologischen Optik Volume 3. Leipzig: Voss, 1866.
 63.Hirsch JA, Wang X, Sommer FT, Martinez LM. How inhibitory circuits in the thalamus serve vision. Annu Rev Neurosci 38: 309‐329, 2015.
 64.Hu B. Functional organization of lemniscal and nonlemniscal auditory thalamus. Exp Brain Res 153: 543‐549, 2003.
 65.Hubel DH, Wiesel TN. Brain mechanisms of vision. Sci Am 241: 150‐162, 1979.
 66.Huguenard JR, McCormick DA. Electrophysiology of the Neuron. New York: Oxford University Press, 1994.
 67.Isu N, Sakuma A, Kitahara M, Watanabe S, Uchino Y. Extracellular recording of vestibulo‐thalamic neurons projecting to the spinal cord in the cat. Neurosci Lett 104: 25‐30, 1989.
 68.Jahnsen H, Llinás R. Electrophysiological properties of guinea‐pig thalamic neurones: An in vitro study. J Physiol (Lond) 349: 205‐226, 1984.
 69.Jahnsen H, Llinás R. Ionic basis for the electroresponsiveness and oscillatory properties of guinea‐pig thalamic neurones in vitro. J Physiol (Lond) 349: 227‐247, 1984.
 70.Janssen J, Aleman‐Gomez Y, Reig S, Schnack HG, Parellada M, Graell M, Moreno C, Moreno D, Mateos‐Perez JM, Udias JM, Arango C, Desco M. Regional specificity of thalamic volume deficits in male adolescents with early‐onset psychosis. Br J Psychiatry 200: 30‐36, 2012.
 71.Jones EG. A new view of specific and nonspecific thalamocortical connections. Adv Neurol 77: 49‐71, 1998.
 72.Jones EG. Viewpoint: The core and matrix of thalamic organization. Neurosci 85: 331‐345, 1998.
 73.Jones EG. The thalamic matrix and thalamocortical synchrony. Trends in Neurosci 24: 595‐601, 2001.
 74.Jones EG. Thalamic circuitry and thalamocortical synchrony. Philos Trans R Soc Lond [Biol] 357: 1659‐1673, 2002.
 75.Jones EG. The Thalamus: Second Edition. Cambridge, U.K.: Cambridge University Press, 2007.
 76.Jones EG. Synchrony in the interconnected circuitry of the thalamus and cerebral cortex. Ann N Y Acad Sci 1157: 10‐23, 2009.
 77.Kalil RE, Chase R. Corticofugal influence on activity of lateral geniculate neurons in the cat. J Neurophysiol 33: 459‐474, 1970.
 78.Kandel ER, Schwartz JH, Jessell TM. Principles of Neural Science. New York: McGraw Hill, 2000.
 79.Kelly LR, Li J, Carden WB, Bickford ME. Ultrastructure and synaptic targets of tectothalamic terminals in the cat lateral posterior nucleus. J Comp Neurol 464: 472‐486, 2003.
 80.Kita T, Kita H. The subthalamic nucleus is one of multiple innervation sites for long‐range corticofugal axons: A single‐axon tracing study in the rat. J Neurosci 32: 5990‐5999, 2012.
 81.Kölliker A. Handbuch der Gerwebelehre des Menschen. Nervensystemen des Menschen und der Thiere. Leipzig: Engelmann, 1896.
 82.Kultas‐Ilinsky K, Ilinsky IA, Young PA, Smith KR. Ultrastructure of degenerating cerebellothalamic terminals in the ventral medial nucleus of the cat. Exp Brain Res 38: 125‐135, 1980.
 83.Kuramoto E, Fujiyama F, Nakamura KC, Tanaka Y, Hioki H, Kaneko T. Complementary distribution of glutamatergic cerebellar and GABAergic basal ganglia afferents to the rat motor thalamic nuclei. Eur J Neurosci 33: 95‐109, 2011.
 84.Kuramoto E, Ohno S, Furuta T, Unzai T, Tanaka YR, Hioki H, Kaneko T. Ventral medial nucleus neurons send thalamocortical afferents more widely and more preferentially to layer 1 than neurons of the ventral anterior‐ventral lateral nuclear complex in the rat. Cereb Cortex 25: 221‐235, 2015.
 85.Lam YW, Sherman SM. Functional organization of the somatosensory cortical layer 6 feedback to the thalamus. Cereb Cortex 20: 13‐24, 2010.
 86.Lam YW, Sherman SM. Activation of both Group I and Group II metabotropic glutamatergic receptors suppress retinogeniculate transmission. Neurosci 242: 78‐84, 2013.
 87.Lavallée P, Urbain N, Dufresne C, Bokor H, Acsády L, Deschênes M. Feedforward inhibitory control of sensory information in higher‐order thalamic nuclei. J Neurosci 25: 7489‐7498, 2005.
 88.Lee CC, Kishan AU, Winer JA. Wiring of divergent networks in the central auditory system. Front Neuroanat 5: 46, 2011.
 89.Lee CC, Sherman SM. Synaptic properties of thalamic and intracortical inputs to layer 4 of the first‐ and higher‐order cortical areas in the auditory and somatosensory systems. J Neurophysiol 100: 317‐326, 2008.
 90.Lee CC, Sherman SM. Glutamatergic inhibition in sensory neocortex. Cereb Cortex 19: 2281‐2289, 2009.
 91.Lee CC, Sherman SM. Topography and physiology of ascending streams in the auditory tectothalamic pathway. Proc Nat Acad Sci Usa 107: 372‐377, 2010.
 92.Lee CC, Sherman SM. Intrinsic modulators of auditory thalamocortical transmission. Hear Res 287: 43‐50, 2012.
 93.Lesica NA, Weng C, Jin J, Yeh CI, Alonso JM, Stanley GB. Dynamic encoding of natural luminance sequences by LGN bursts. Plos Biol 4: e209, 2006.
 94.Levitan IB, Kaczmarek LK. The Neuron: Cell and Molecular Biology. New York: Oxford University Press, 2002.
 95.Linden R, Perry VH. Massive retinotectal projection in rats. Brain Res 272: 145‐149, 1983.
 96.Llano DA, Sherman SM. Evidence for nonreciprocal organization of the mouse auditory thalamocortical‐corticothalamic projection systems. J Comp Neurol 507: 1209‐1227, 2008.
 97.Lujan R, Nusser Z, Roberts JD, Shigemoto R, Somogyi P. Perisynaptic location of metabotropic glutamate receptors mGluR1 and mGluR5 on dendrites and dendritic spines in the rat hippocampus. Eur J Neurosci 8: 1488‐1500, 1996.
 98.Markov NT, Vezoli J, Chameau P, Falchier A, Quilodran R, Huissoud C, Lamy C, Misery P, Giroud P, Ullman S, Barone P, Dehay C, Knoblauch K, Kennedy H. Anatomy of hierarchy: Feedforward and feedback pathways in macaque visual cortex. J Comp Neurol 522: 225‐259, 2014.
 99.Matsuo S, Hasogai M, Nakao S. Ascending projections of posterior canal activated excitatory and inhibitory secondary vestibular neurons to the mesodiencephalon in cats. Exp Brain Res 100: 7‐17, 1994.
 100.McAlonan K, Cavanaugh J, Wurtz RH. Guarding the gateway to cortex with attention in visual thalamus. Nature 2008.
 101.McClurkin JW, Marrocco RT. Visual cortical input alters spatial tuning in monkey lateral geniculate nucleus cells. J Physiol (Lond) 348: 135‐152, 1984.
 102.McClurkin JW, Optican LM, Richmond BJ. Cortical feedback increases visual information transmitted by monkey parvocellular lateral geniculate nucleus neurons. Visual Neurosci 11: 601‐617, 1994.
 103.McCrea RA, Bishop GA, Kitai ST. Morphological and electrophysiological characteristics of projection neurons in the nucleus interpositus of the cat cerebellum. J Comp Neurol 181: 397‐419, 1978.
 104.McFarland NR, Haber SN. Thalamic relay nuclei of the basal ganglia form both reciprocal and nonreciprocal cortical connections, linking multiple frontal cortical areas. J Neurosci 22: 8117‐8132, 2002.
 105.Means LW, Harrell TH, Mayo ES, Alexander GB. Effects of dorsomedial thalamic lesions on spontaneous alternation, maze, activity and runway performance in the rat. Physiol Behav 12: 973‐979, 1974.
 106.Mease RA, Krieger P, Groh A. Cortical control of adaptation and sensory relay mode in the thalamus. Proc Natl Acad Sci U S A 2014.
 107.Merabet L, Desautels A, Minville K, Casanova C. Motion integration in a thalamic visual nucleus. Nature 396: 265‐268, 1998.
 108.Mitelman SA, Brickman AM, Shihabuddin L, Newmark R, Chu KW, Buchsbaum MS. Correlations between MRI‐assessed volumes of the thalamus and cortical Brodmann's areas in schizophrenia. Schizophr Res 75: 265‐281, 2005.
 109.Mitelman SA, Byne W, Kemether EM, Hazlett EA, Buchsbaum MS. Metabolic disconnection between the mediodorsal nucleus of the thalamus and cortical Brodmann's areas of the left hemisphere in schizophrenia. Am J Psychiatry 162: 1733‐1735, 2005.
 110.Moore T, Armstrong KM. Selective gating of visual signals by microstimulation of frontal cortex. Nature 421: 370‐373, 2003.
 111.Nakamura K, Colby CL. Updating of the visual representation in monkey striate and extrastriate cortex during saccades. Proc Natl Acad Sci U S A 99: 4026‐4031, 2002.
 112.Nicoll RA, Malenka RC, Kauer JA. Functional comparison of neurotransmitter receptor subtypes in mammalian central nervous system. Physiol Rev 70: 513‐565, 1990.
 113.Ojima H. Terminal morphology and distribution of corticothalamic fibers originating from layers 5 and 6 of cat primary auditory cortex. Cereb Cortex 4: 646‐663, 1994.
 114.Ojima H, Murakami K. Triadic synaptic interactions of large corticothalamic terminals in non‐lemniscal thalamic nuclei of the cat auditory system. Hear Res 274: 40‐47, 2011.
 115.Olsen SR, Bortone DS, Adesnik H, Scanziani M. Gain control by layer six in cortical circuits of vision. Nature 483: 47‐52, 2012.
 116.Ortuno T, Grieve KL, Cao R, Cudeiro J, Rivadulla C. Bursting thalamic responses in awake monkey contribute to visual detection and are modulated by corticofugal feedback. Front Behav Neurosci 8: 198, 2014.
 117.Parnaudeau S, O'Neill PK, Bolkan SS, Ward RD, Abbas AI, Roth BL, Balsam PD, Gordon JA, Kellendonk C. Inhibition of mediodorsal thalamus disrupts thalamofrontal connectivity and cognition. Neuron 77: 1151‐1162, 2013.
 118.Perea G, Yang A, Boyden ES, Sur M. Optogenetic astrocyte activation modulates response selectivity of visual cortex neurons in vivo. Nat Commun 5: 3262, 2014.
 119.Pereira de Vasconcelos A, Cassel JC. The nonspecific thalamus: A place in a wedding bed for making memories last? Neurosci Biobehav Rev 54: 175‐196, 2015.
 120.Pesaran B, Nelson MJ, Andersen RA. Free choice activates a decision circuit between frontal and parietal cortex. Nature 453: 406‐409, 2008.
 121.Petrof I, Sherman SM. Synaptic properties of the mammillary and cortical afferents to the anterodorsal thalamic nucleus in the mouse. J Neurosci 29: 7815‐7819, 2009.
 122.Petrof I, Viaene AN, Sherman SM. Synaptic properties of the lemniscal and paralemniscal somatosensory inputs to the mouse thalamus. Soc Neurosci, 2012.
 123.Petrof I, Viaene AN, Sherman SM. Two populations of corticothalamic and interareal corticocortical cells in the subgranular layers of the mouse primary sensory cortices. J Comp Neurol 520: 1678‐1686, 2012.
 124.Purushothaman G, Marion R, Li K, Casagrande VA. Gating and control of primary visual cortex by pulvinar. Nat Neurosci 15: 905‐912, 2012.
 125.Purves D, Augustine GJ, Fitzpatrick D, Hall WC, Lamantia A‐S, White LE. Neuroscience. Sunderland, MA: Sinauer, 2012.
 126.Pynn LK, DeSouza JF. The function of efference copy signals: Implications for symptoms of schizophrenia. Vision Res 76: 124‐133, 2013.
 127.Rafal RD, Posner MI. Deficits in human visual spatial attention following thalamic lesions. Proc Natl Acad Sci U S A 84: 7349‐7353, 1987.
 128.Ramcharan EJ, Gnadt JW, Sherman SM. Higher‐order thalamic relays burst more than first‐order relays. Proc Natl Acad Sci U S A 102: 12236‐12241, 2005.
 129.Reichova I, Sherman SM. Somatosensory corticothalamic projections: Distinguishing drivers from modulators. J Neurophysiol 92: 2185‐2197, 2004.
 130.Reynolds JH, Chelazzi L, Desimone R. Competitive mechanisms subserve attention in macaque areas V2 and V4. J Neurosci 19: 1736‐1753, 1999.
 131.Richard D, Gioanni Y, Kitsikis A, Buser P. A study of geniculate unit activity during cryogenic blockade of the primary visual cortex in the cat. Exp Brain Res 22: 235‐242, 1975.
 132.Rockland KS. Convergence and branching patterns of round, type 2 corticopulvinar axons. J Comp Neurol 390: 515‐536, 1998.
 133.Rosler L, Rolfs M, van der Stigchel S, Neggers SF, Cahn W, Kahn RS, Thakkar KN. Failure to use corollary discharge to remap visual target locations is associated with psychotic symptom severity in schizophrenia. J Neurophysiol 114: 1129‐1136, 2015.
 134.Roth MM, Dahmen JC, Muir DR, Imhof F, Martini FJ, Hofer SB. Thalamic nuclei convey diverse contextual information to layer 1 of visual cortex. Nat Neurosci 19: 299‐307, 2016.
 135.Saalmann YB, Pinsk MA, Wang L, Li X, Kastner S. The pulvinar regulates information transmission between cortical areas based on attention demands. Science 337: 753‐756, 2012.
 136.Sakai ST, Inase M, Tanji J. Comparison of cerebellothalamic and pallidothalamic projections in the monkey (Macaca fuscata): A double anterograde labeling study. J Comp Neurol 368: 215‐228, 1996.
 137.Scannell JW, Blakemore C, Young MP. Analysis of connectivity in the cat cerebral cortex. J Neurosci 15: 1463‐1483, 1995.
 138.Schmielau F, Singer W. The role of visual cortex for binocular interactions in the cat lateral geniculate nucleus. Brain Res 120: 354‐361, 1977.
 139.Seidemann E, Zohary E, Newsome WT. Temporal gating of neural signals during performance of a visual discrimination task. Nature 394: 72‐75, 1998.
 140.Sherman SM. Functional organization of the W‐,X‐, and Y‐cell pathways in the cat: A review and hypothesis. In: Sprague JM, Epstein AN, editors. Progress in Psychobiology and Physiological Psychology, Vol. 11. Orlando: Academic Press, 1985, pp. 233‐314.
 141.Sherman SM. Tonic and burst firing: Dual modes of thalamocortical relay. Trends in Neurosci 24: 122‐126, 2001.
 142.Sherman SM. Interneurons and triadic circuitry of the thalamus. Trends in Neurosci 27: 670‐675, 2004.
 143.Sherman SM. The thalamus is more than just a relay. Curr Opin Neurobiol 17: 1‐6, 2007.
 144.Sherman SM. The function of metabotropic glutamate receptors in thalamus and cortex. Neuroscientist 20: 136‐149, 2014.
 145.Sherman SM. Thalamus plays a central role in ongoing cortical functioning. Nat Neurosci 19: 533‐541, 2016.
 146.Sherman SM, Guillery RW. On the actions that one nerve cell can have on another: Distinguishing “drivers” from “modulators”. Proc Natl Acad Sci U S A 95: 7121‐7126, 1998.
 147.Sherman SM, Guillery RW. Exploring the Thalamus and its Role in Cortical Function. Cambridge, MA: MIT Press, 2006.
 148.Sherman SM, Guillery RW. Thalamocortical Processing: Understanding the Messages that Link the Cortex to the World. Cambridge, MA: MIT Press, 2013.
 149.Sherman SM, Koch C. Thalamus. In: Shepherd GM, editor. The Synaptic Organization of the Brain, Fourth Edition. New York: Oxford University Press, 1998, pp. 289‐328.
 150.Shinoda Y, Futami T, Mitoma H, Yokota J. Morphology of single neurones in the cerebello‐rubrospinal system. Behav Brain Res 28: 59‐64, 1988.
 151.Sillito AM, Jones HE, Gerstein GL, West DC. Feature‐linked synchronization of thalamic relay cell firing induced by feedback from the visual cortex. Nature 369: 479‐482, 1994.
 152.Sirota MG, Swadlow HA, Beloozerova IN. Three channels of corticothalamic communication during locomotion. J Neurosci 25: 5915‐5925, 2005.
 153.Smith GD, Cox CL, Sherman SM, Rinzel J. Fourier analysis of sinusoidally‐driven thalamocortical relay neurons and a minimal integrate‐and‐fire‐or‐burst model. J Neurophysiol 83: 588‐610, 2000.
 154.Smith GD, Sherman SM. Detectability of excitatory versus inhibitory drive in an integrate‐and‐fire‐or‐burst thalamocortical relay neuron model. J Neurosci 22: 10242‐10250, 2002.
 155.Solomon SG, Peirce JW, Dhruv NT, Lennie P. Profound contrast adaptation early in the visual pathway. Neuron 42: 155‐162, 2004.
 156.Sommer MA, Wurtz RH. What the brain stem tells the frontal cortex. II. Role of the SC‐MD‐FEF pathway in corollary discharge. J Neurophysiol 91: 1403‐1423, 2004.
 157.Sommer MA, Wurtz RH. Brain circuits for the internal monitoring of movements. Annu Rev Neurosci 31: 317‐338, 2008.
 158.Spering M, Dias EC, Sanchez JL, Schutz AC, Javitt DC. Efference copy failure during smooth pursuit eye movements in schizophrenia. J Neurosci 33: 11779‐11787, 2013.
 159.Sperry RW. Neural basis of the spontaneous optokinetic response produced by visual inversion. J Comp Neurol 43: 482‐489, 1950.
 160.Stroh A, Adelsberger H, Groh A, Ruhlmann C, Fischer S, Schierloh A, Deisseroth K, Konnerth A. Making waves: Initiation and propagation of corticothalamic ca(2+) waves in vivo. Neuron 77: 1136‐1150, 2013.
 161.Swadlow HA, Gusev AG. The impact of ‘bursting’ thalamic impulses at a neocortical synapse. Nat Neurosci 4: 402‐408, 2001.
 162.Swadlow HA, Gusev AG, Bezdudnaya T. Activation of a cortical column by a thalamocortical impulse. J Neurosci 22: 7766‐7773, 2002.
 163.Tamamaki N, Uhlrich DJ, Sherman SM. Morphology of physiologically identified retinal X and Y axons in the cat's thalamus and midbrain as revealed by intra‐axonal injection of biocytin. J Comp Neurol 354: 583‐607, 1994.
 164.Theyel BB, Llano DA, Sherman SM. The corticothalamocortical circuit drives higher‐order cortex in the mouse. Nat Neurosci 13: 84‐88, 2010.
 165.Tsumoto T, Suda K. Three groups of cortico‐geniculate neurons and their distribution in binocular and monocular segments of cat striate cortex. J Comp Neurol 193: 223‐236, 1980.
 166.Uhlrich DJ, Cucchiaro JB, Humphrey AL, Sherman SM. Morphology and axonal projection patterns of individual neurons in the cat perigeniculate nucleus. J Neurophysiol 65: 1528‐1541, 1991.
 167.Van Essen DC, Anderson CH, Felleman DJ. Information processing in the primate visual system: An integrated systems perspective. Science 255: 419‐423, 1992.
 168.Van Horn SC, EriΠir A, Sherman SM. The relative distribution of synapses in the A‐laminae of the lateral geniculate nucleus of the cat. J Comp Neurol 416: 509‐520, 2000.
 169.Van Horn SC, Sherman SM. Fewer driver synapses in higher order than in first order thalamic relays. Neurosci 475: 406‐415, 2007.
 170.Vaney DI, Levick WR, Thibos LN. Rabbit retinal ganglion cells. Exp Brain Res 44: 27‐33, 1981.
 171.Varela C, Sherman SM. Differences in response to muscarinic agonists between first and higher order thalamic relays. J Neurophysiol 98: 3538‐3547, 2007.
 172.Varela C, Sherman SM. Differences in response to serotonergic activation between first and higher order thalamic nuclei. Cereb Cortex 19: 1776‐1786, 2008.
 173.Veinante P, Jacquin MF, Deschênes M. Thalamic projections from the whisker sensitive regions of the spinal trigeminal complex in the rat. J Comp Neurol 420: 233‐243, 2000.
 174.Veinante P, Lavallée P, Deschênes M. Corticothalamic projections from layer 5 of the vibrissal barrel cortex in the rat. J Comp Neurol 424: 197‐204, 2000.
 175.Vertes RP. Major diencephalic inputs to the hippocampus: Supramammillary nucleus and nucleus reuniens. Circuitry and function. Prog Brain Res 219: 121‐144, 2015.
 176.Vertes RP, Linley SB, Hoover WB. Limbic circuitry of the midline thalamus. Neurosci Biobehav Rev 54: 89‐107, 2015.
 177.Viaene AN, Petrof I, Sherman SM. Properties of the thalamic projection from the posterior medial nucleus to primary and secondary somatosensory cortices in the mouse. Proc Nat Acad Sci Usa 108: 18156‐18161, 2011.
 178.Viaene AN, Petrof I, Sherman SM. Activation requirements for metabotropic glutamate receptors. Neurosci Lett 541: 67‐72, 2013.
 179.von Holst E, Mittelstaedt H. The reafference principle. Interaction between the central nervous system and the periphery. In: Translated by Robert Martin, editor. Selected Papers of Erich von Holst: The Behavioural Physiology of Animals and Man. Coral Gables: University of Miami Press, 1950, p. 139‐173.
 180.Wang S, Eisenback MA, Bickford ME. Relative distribution of synapses in the pulvinar nucleus of the cat: Implications regarding the “driver/modulator” theory of thalamic function. J Comp Neurol 454: 482‐494, 2002.
 181.Wenstrup JJ. The tectothalamic system. In: Winer JA, Schreiner CE, editors. The Inferior Colliculus. New York: Springer, 2005, pp. 200‐230.
 182.Whitmire CJ, Millard DC, Stanley GB. Thalamic state control of cortical paired‐pulse dynamics. J Neurophysiol 117: 163‐177, 2016.
 183.Wilke M, Mueller KM, Leopold DA. Neural activity in the visual thalamus reflects perceptual suppression. Proc Nat Acad Sci Usa 106: 9465‐9470, 2009.
 184.Wilson JR, Friedlander MJ, Sherman SM. Fine structural morphology of identified X‐ and Y‐cells in the cat's lateral geniculate nucleus. Proc Roy Soc Lond B 221: 411‐436, 1984.
 185.Wolpert DM, Flanagan JR. Motor learning. Curr Biol 20: R467‐R472, 2010.
 186.Womelsdorf T, Fries P, Mitra PP, Desimone R. Gamma‐band synchronization in visual cortex predicts speed of change detection. Nature 439: 733‐736, 2006.
 187.Wurtz RH, Sommer MA. Identifying corollary discharges for movement in the primate brain. Prog Brain Res 144: 47‐60, 2004.
 188.Xu XM, Bosking WH, White LE, Fitzpatrick D, Casagrande VA. Functional organization of visual cortex in the prosimian bush baby revealed by optical imaging of intrinsic signals. J Neurophysiol 94: 2748‐2762, 2005.
 189.Zhan XJ, Cox CL, Rinzel J, Sherman SM. Current clamp and modeling studies of low threshold calcium spikes in cells of the cat's lateral geniculate nucleus. J Neurophysiol 81: 2360‐2373, 1999.
 190.Zhou H, Schafer RJ, Desimone R. Pulvinar‐cortex interactions in vision and attention. Neuron 89: 209‐220, 2016.

Contact Editor

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

* Required Field

How to Cite

S. Murray Sherman. Functioning of Circuits Connecting Thalamus and Cortex. Compr Physiol 2017, 7: 713-739. doi: 10.1002/cphy.c160032