Models of thalamocortical system

From Scholarpedia
Richard H. Granger and Robert A. Hearn (2007), Scholarpedia, 2(11):1796. doi:10.4249/scholarpedia.1796 revision #89049 [link to/cite this article]
Jump to: navigation, search
Post-publication activity

Curator: Robert A. Hearn

The thalamocortical system constitutes the vast majority of the mammalian brain, and increases disproportionately (allometrically) with overall brain size. Commensurate with its size, the thalamocortical system has been the subject of extensive neurobiological and computational study.

The thalamus and the neocortex are reciprocally connected via pathways of varying levels of topography. In addition, many areas of cortex and thalamus participate in a cortico-striato-pallido-thalamocortical loop. This article considers possible computational roles for these two kinds of loops (direct, and integrally involving the basal ganglia). In this article "thalamus" without qualification refers to the mammalian dorsal thalamus.

Elucidation of computational function relies on accurate anatomical and physiological data, which this article will begin by surveying. It should be noted that much of the neuroanatomy of thalamocortical pathways is incompletely mapped, leaving many topics of ongoing research, some of which will be noted.

Contents

Thalamocortical Anatomy

The primary sensory signals of vision, audition and touch pass through thalamus en route to cortex. These sensory pathways include the optic tract (vision), the the brachium of the inferior colliculus (audition), and the medial lemniscus (somatosensation), which target, respectively, the dorsal lateral geniculate nucleus, the medial geniculate complex, and the ventral posterior complex of the thalamus. Those thalamic nuclei in turn project topographically to well-defined architectonic fields of cortex, which reciprocally project back topographically to their innervating thalamic nuclei. Both the thalamocortical fibers and the reciprocal corticothalamic fibers also send collaterals to the thalamic reticular nucleus (nucleus reticularis thalami) (Liu and Jones, 1999) which is a thin shell of GABAergic cells surrounding the thalamus. The thalamic reticular nucleus then reciprocally innervates the thalamus, topographically (DeFelipe and Jones, 1991; Jones, 2001). (There is not a one-to-one correspondence; there are fewer reticular cells than thalamic relay cells.) The entirety of the thalamus is also innervated by diffuse brainstem reticular inputs. For details of the internal circuitry of thalamic nuclei see Thalamus.

The three sensory pathways described above (especially vision) are perhaps the best studied, but other thalamic nuclei receive subcortical input, project to cortex, and are innervated by cortex. However, the specific projection patterns are in general not as straightforward as for the primary sensory pathways mentioned.

Perspectives on Thalamocortical and Corticothalamic Connectivity

Two current detailed perspectives on fitting the patterns of thalamocortical and corticothalamic connectivity into a coherent picture are noteworthy.

Core vs. Matrix

In the system of Jones (2001, 2007), thalamic relay cells fall into one of two classes: core and matrix. From a given thalamic nucleus, core cells project topographically to a sharply bounded, associated architectonic field of cortex. For example, the core cells in the dorsal lateral geniculate nucleus project topographically to V1 (Brodmann area 17). Projections from thalamic core cells synapse on neurons in all cortical layers to some extent (Keller and White, 1989) but predominantly in deep layer III and in layer IV in granular cortex, as well as on the apical dendrites of layer VI neurons (Molinari et al., 1995; Jones, 2001). These afferents, which preserve topographic organization, are often described as the primary input to sensory neocortical regions, though quantitative anatomical studies report that these afferents comprise perhaps 6% of the synapses onto layer IV target cells, with the majority of the remaining afferents reportedly arriving from lateral cortico-cortical connections (Freund et al., 1985; Freund et al., 1989; Peters and Payne, 1993; Peters et al., 1994; Ahmed et al., 1997). Projections from a given thalamic core region extend to a cortical area roughly 0.5-1.0 mm wide, somewhat larger than the size of physiologically-delineated functional columns (Jones, 1981). Example thalamic core nuclei include MGv, VPL, VPM, LGd.

In contrast, matrix cells project more broadly and less topographically, crossing architectonic borders, synapsing chiefly in layer I, on apical dendrites of neurons in multiple layers including layers II, III, and V. The earliest detailed reports of these projections emphasize that they occur as a prevalent feature of cortical anatomy, and describe them as “nonspecific”, i.e., projections from small thalamic regions innervate broad cortical areas, and projections to circumscribed cortical areas may originate from a broad expanse of thalamus (Lorente de No, 1938). These initial findings have been confirmed and extended repeatedly (Killackey and Ebner, 1972; Killackey and Ebner, 1973; Herkenham, 1986; Jones, 1998), and it has consistently been found that thalamic cells projecting to a given cortical area receive projections back from layer V of that cortical area without intervening NRt contacts (Conley and Diamond, 1990; Rouiller et al., 1991; Bourassa and Deschênes, 1995; Deschênes et al., 1998). Example thalamic nuclei include MGm, Pul, Pom, AD. Matrix cells are distributed throughout the entire thalamus. In certain nuclei only, a population of core cells is superimposed. The nuclei containing core cells receive subcortical inputs that are highly ordered topographically, whereas the nuclei without a core population receive less topographic subcortical inputs. In addition to the primary sensory pathways mentioned above, other core pathways include cerebellum -> ventral lateral posterior nucleus -> motor cortex, and mammilary nuclei -> anterior thalamic nuclei -> limbic areas.

Core projections to cortex are reciprocated by topographic projections from layer VI of cortex. Although there is a commonly accepted principle of reciprocity (Diamond et al., 1969) stating that every cortical area returns fibers to the thalamic nucleus that provides its dominant thalamic input, most studies have focused on primary sensory relays, which contain core cells. It is unclear what the pattern is for nuclei with no significant core population. Also, Deschênes et al. (1998) note several nonreciprocal corticothalamic layer VI projections, even for primary sensory areas, and propose instead a rule of parity, which states that the distribution of layer VI corticothalamic projections across and within the thalamic nuclei is determined by the branching patterns of the different classes of prethalamic afferent. This rule can explain some of the observed nonreciprocal corticothalamic projections, and can further be seen as congruent with Jones's association of the degree of topography of thalamic input with the degree of topography of corresponding thalamocortical output.

Layer V cells also project to thalamus, as mentioned. However, these projections tend not to return to the primary innervating thalamic nucleus. Instead, they project to different, but usually functionally related, nuclei. These fibers lack collaterals to the thalamic reticular nucleus. Also, the thalamic innervation is itself via a collateral, the layer V fibers continuing on to other subcortical targets. The thalamic nuclei targeted by the layer V projections tend to be core-poor and matrix-rich, suggesting that the matrix relay cells are the targets (Jones, 2001). Jones proposes that one function of this pattern is to synchronize fast cortical oscillations via the matrix projections back to cortex.

The terminology is intended to extend the long-studied distinctions in the literature between “specific” and “nonspecific” nuclei (Killackey and Ebner, 1972; 1973; Berendse and Groenewegen, 1991; Wyss and VanGroen, 1995; Castro-Alamancos and Connors, 1997). There is evidence for some correspondence between matrix cells and calbindin-immunoreactivity, and between core cells and parvalbumin-immunoreactivity in primates (Jones et al., 1989) though specific exceptions to this idea have been found in primates, and it is important to note that the correspondence does not hold in other mammals (Jones, 2007, p. 113; Sherman and Guillery 2006).

First-Order vs. Higher-Order Relays

An important distinction is often made between first-order relays, predominantly denoting primary sensory relay nuclei, as distinct from higher-order thalamic relays. The layer V corticothalamic pathway is also reported to provide higher-order thalamic relays with a driving input that, unlike first-order relays, they otherwise lack (Sherman and Guillery 2002; 2006; see also Thalamus). Since the layer V output represents cortical motor output broadly defined, the thalamic input it provides is an efferent copy of the motor commands issued by a region of cortex. This cortico-thalamo-cortical pathway has been proposed as a largely-unexplored means of cortico-cortical communication, subject to modulatory control; it has been suggested that direct cortico-cortical pathways, rather than driving inputs (e.g., Felleman et al., 1991; Hilgetag et al., 1996), may be primarily modulatory in nature over these cortico-thalamo-cortical paths (Sherman and Guillery 2006).

In comparison to the core vs. matrix perspective, intriguing questions arise regarding the laminar targets of relay cells. Can the cells in higher-order nuclei, which from a core vs. matrix perspective would be matrix cells, be construed as having a parallel function to those in the first-order nuclei, if they target superficial rather than middle cortical layers? Traditionally, the middle layers are viewed at the recipient layers of driving thalamic input. However, there is evidence that the cat pulvinar nucleus, which in this scheme is a higher-order visual relay (with driving input from layer V of primary visual cortical areas), may indeed target layer IV neurons in multiple cortical regions (excluding V1) (Symonds et al., 1981, Abramson et al., 1985).

Figure 1: The primate thalamus.

Number of Cortical Cells

There are (with a few notable exceptions, such as the primary sensory areas) approximately 80,000 neurons beneath each square millimeter of cortical surface, distributed in a stereotypic manner across the cortical layers. Despite dendritic elongation in larger brains, and corresponding increases in numbers of synaptic contacts among pyramidal cells, the number of cells stays constant per region in animals weighing from grams to kilograms (Rockel et al., 1980). These cortical cells are innervated by a much smaller number of thalamic afferents; ratios of roughly 160 cortical cells per corresponding thalamic relay cell are typical (O'Kusky and Colonnier, 1982).

Cortical Cell Types

Excitatory (pyramidal) cells outnumber inhibitory cells by roughly four or five to one throughout most of cortex, again excepting the primary sensory areas (Fitzpatrick et al., 1987; Hendry et al., 1987). Excitatory neurons have axons that can extend millimeters whereas inhibitory cells project only locally (rarely more than 100 μm). Inhibitory axons synapse densely on or near pyramidal cell bodies (Keller and White, 1989). In contrast, excitatory cells receive only sparse afferents from other excitatory cells; it has been estimated that the probability of contact between two neocortical excitatory cells that are 0.2-0.3mm apart is less than 0.1, and between two such cells that are more than 1mm from each other, p < 0.01 (Braitenberg and Schüz, 1998).

Cortical Modules

Within architectonic regions of cortex, neurons are vertically organized into anatomically defined "pyramidal cell modules" consisting of distinct groups of layer V and layer II-III pyramidal cells whose apical dendrites are commingled (White and Peters, 1993; Peters et al., 1994). Architectonically distinguishable areas differ in size and population of cell layers, and there is correspondence between these region boundaries and the site of origin of their thalamic afferents. (In contrast, functional "columns" are physiologically defined, in terms of receptive field properties, rather than anatomical boundaries (Mountcastle, 1957), and are typically described as 400-500 μm in extent, comprising perhaps 200 pyramidal cell modules apiece.

Quantitative data on the microcircuits within cortex are incomplete. However, by extrapolating from known laminar distributions of neuron types in cortex and using several reconstructed neurons to provide morphometrical data, Binzegger et al. (2004) have produced a quantitative map of the circuit of cat primary visual cortex. Questions about the detailed statistics of cortico-cortical connectivity have recently been addressed via computational modeling (Felch & Granger 2008).

Thalamocortical Physiology

The highly recurrent connectivity of the thalamocortical system, in addition to inputs from basal ganglia and thalamic modulation by brainstem reticular input, belie the seemingly simple notion that thalamus is a mere "relay" from the senses to cortex.

Sequential circuit activation

Peripheral inputs activate thalamic core cells which in turn participate in topographic activation of middle cortical layers; e.g., ear > cochlea > auditory brainstem nuclei > ventral subdivision of medial geniculate nucleus (MGv) > A1; in contrast, matrix nuclei are most strongly driven by corticothalamic feedback (Bender, 1983; Diamond et al., 1992b; Diamond et al., 1992a), supporting a system in which peripheral afferents first activate core nuclei, which in turn activate cortex (via a stereotypical vertical pattern: middle layers > superficial layers > deep layers), which then activate both core and matrix nuclei via corticothalamic projections (Mountcastle, 1957; Hubel and Wiesel, 1977; Di et al., 1990; Kenan-Vaknin and Teyler, 1994). Although the majority of excitatory synapses in all thalamic nuclei arise from corticothalamic fibers, the different dendritic locations of subcortical afferent synapses vs. corticothalamic synapses supports this essentially feedforward view in non-matrix nuclei (Jones, 2007, p. 289). (But see Llinás and Paré, 1991, and also Oscillation-Assisted Processing, below, for a different perspective.)

Excitatory and inhibitory interaction

Axons of inhibitory interneurons densely terminate preferentially on the bodies, initial axon segments, and proximal apical dendrites of excitatory pyramidal cells in cortex, and thus are well situated to exert powerful control over the activity of target excitatory neurons. Inhibitory cells receive direct thalamocortical innervation, resulting in feedforward inhibition. When a field of excitatory neurons receives afferent stimulation within a window permitted by this feedforward inhibition, those that are most responsive will activate the local inhibitory cells in their neighborhood, which will in turn inhibit local excitatory cells. The typical time course of an excitatory (depolarizing) postsynaptic potential (PSP) at normal resting potential, in vivo, is brief (15-20 msec), whereas corresponding GABAergic inhibitory PSPs last roughly an order of magnitude longer (~50 msec GABAa response, 200-300 msec GABAb response) (Castro-Alamancos and Connors, 1997). Thus excitation tends to be brief, sparse, and curtailed by longer and stronger feedback inhibition.

Activity rates

The rate of repetitive activation in thalamocortical circuits ranges from the "slow sleep" (.2-1 Hz) and "delta" (1-4 Hz) frequency bands through the "gamma" range (30-80 Hz) (Steriade, 1997; Chrobak and Buzsaki, 1998; Shimono et al., 2000; Sarter and Bruno, 2000; Fries et al., 2001; Rozov et al., 2001; Canales et al., 2002; Knoblauch and Palm, 2002; Pesaran et al., 2002). (Gamma is occasionally considered to range as high as 120 Hz.) There is strong evidence for ascending influences (e.g., basal forebrain) on inhibitory neurons (Freund and Meskenaite, 1992; Gulyas et al., 1996; Blasco-Ibanez et al., 1998; Gulyas et al., 1999) modulating their response properties, in turn affecting the probability of response of excitatory cells during the peaks and troughs of such "clocked" inhibitory cycles. Evidence of intrinsic rhythmic currents in thalamic and cortical cells (Kim et al., 1995; Bush and Sejnowski, 1996; Destexhe et al., 1999; Zhu and Connors, 1999) is compatible with extrinsic ascending influences, acting either independently or in concert with them. Three modes of activity have typically been reported for thalamic neurons: tonic, rhythmic, and arrhythmic bursting. The latter appears predominantly during non-REM sleep whereas the first two appear during waking behavior (McCarley et al., 1983; Steriade and Llinas, 1988; McCormick and Feeser, 1990; Steriade et al., 1990; McCormick and Bal, 1994; Steriade and Contreras, 1995). It has been variously argued that rhythmic burst mode may provide better signal to noise and thus facilitate detection of a stimulus, and that tonic mode contains more detailed information about a stimulus (Guido et al., 1992; Guido et al., 1995; Mukherjee and Kaplan, 1995; Sherman, 2001). Others have suggested that distinctions between modes based on differential information are not warranted (e.g., Reinagel et al., 1999).

Temporal activity

Notable patterns occur in the activity of thalamocortical circuits:

  • Synchronous activity of wide regions of cortex (modulated in part by ascending systems affecting the periodic responsivity of inhibitory cells) makes the probability of excitatory cell spiking lower during peak inhibition and higher during inhibitory troughs.
  • The average time course of excitatory postsynaptic potentials in cortical pyramidal cells (~ 10-15 msec) appears to set limits on the temporal precision of spike trains that such a neuron may emit.
  • Summation characteristics and integration (e.g., capacitance) time constants of dendrites map many distinct spike train input patterns into postsynaptic voltage transients that are difficult to distinguish, limiting the temporal precision with which a target neuron can "read" differences among slightly different spike trains. However, rapid inhibitory feedback from local GABAa-type cortical interneurons contributes to increased synchronization of neighboring excitatory cells, a mechanism by which carefully timed firing may be achieved, potentially enabling spikes with sufficiently precise timing to support temporal coding (Magee, 2000; Magee and Cook, 2000).

Glutamatergic synapses

The vast majority of excitatory synapses throughout telencephalon are glutamatergic (Rodriguez et al., 2004). An excitatory axon targeting the apical dendrite of an excitatory cell typically terminates at a spine, which contains ~500-1000 AMPA- and NMDA-type glutamate receptors (Bekkers and Stevens, 1989). An average neocortical pyramidal cell in humans reportedly receives 25-80 thousand such afferents (Cragg, 1967; Rockel et al., 1980; Braitenberg and Schüz, 1998 pp. 190-191) (with a few notable exceptions such as area 17, which has an unusually high density of neurons per square mm and a correspondingly low number of synapses per neuron (Cragg, 1967; O'Kusky and Colonnier, 1982)), and typical methods may lead to systematic undercounting of synapses (Guillery and Herrup, 1997; von Bartheld, 1999, 2001). In certain regions, notably thalamus and layer IV cortex, as well as glutamatergic synapses onto inhibitory neurons, the NMDA receptors contain the (rare in the adult) NR3A subunit (Wong et al., 2002), which has been shown to inhibit the expression of NMDA receptor ion channels (Das et al., 1998).

Neocortical synaptic potentiation

NMDA-dependent long-term potentiation of synaptic connections in neocortex has been shown in superficial and deep layers of multiple regions (Komatsu et al., 1988; Hirsch and Crepel, 1990; Iriki et al., 1991; Bear and Kirkwood, 1993; Kirkwood et al., 1993; Kimura et al., 1994; Castro-Alamancos et al., 1995; Hess et al., 1996; Kudoh and Shibuki, 1996; Buonomano and Merzenich, 1998; Rioult-Pedotti et al., 2000; Heynen and Bear, 2001; Seki et al., 2001). Memories that are rapidly induced (i.e., with little or no practice), long lasting (potentially for decades) and high-capacity (enough to hold the memories of a lifetime) presumably require a biological mechanism with corresponding characteristics. Biological phenomena that last only for limited duration (decrementing over time), or are slow to induce (e.g., minutes of constant stimulation), or are not synapse-specific (and thus not high capacity) may underlie some form of short-term memory (or other operation) but not rapidly-induced, high-capacity long-term memory. "LTP" here refers specifically to the endogenously occurring synaptic phenomenon that has the properties just listed, enabling it to serve as the substrate of lifelong memories.

Thalamocortical Computation

Feedforward Sensory Processing

Spatiotemporal Coding

From a feedforward perspective, the thalamus relays sensory information to cortex. Several computational models have addressed how the spatial and temporal characteristics of sensory signals are transformed by thalamus and cortex, and how signal transmission can be modulated.

One principle that emerges is that of sparse coding: that sensory information may be encoded using only a small number of potentially active neurons at a time (Olshausen and Field, 2004; Field, 1987; Bell and Sejnowski, 1997; Lewicki 2002). More generally, several researchers have investigated the hypothesis that the thalamocortical system is optimized to efficiently code the statistical properties of the signals to which it is exposed (Attneave, 1954; Barlow, 1961; Simoncelli and Olshausen, 2001).

A critical feature of thalamocortical transmission is feedforward inhibition in cortex. Thalamocortical afferents contact both excitatory projection neurons and fast-spiking, local inhibitory interneurons, which synapse on the same projection neurons (Miller et al., 2001b). This characteristic microcircuit can act as a precise coincidence detector (Hull and Scanziani, 2007), and there are indications that it can also explain such diverse aspects of cortical sensory representation as orientation tuning in visual cortex (Troyer et al., 1998) and temporal tuning (inhibited response to fast-moving visual stimuli, and related temporal low-pass filtering effects in other cortical areas) (Krukowski and Miller, 2001). More generally, spatiotemporal integration across the neural receptive field is a general feature of sensory coding (Boloori and Stanley, 2006; Webber and Stanley, 2004).

Olhausen and Field (2005) sound a cautionary note about the degree to which even the feedforward aspects of sensory processing can ever be understood in terms of combinations of reduced stimuli such as spots, white noise, or sine wave gratings, due to the highly nonlinear response properties of real, as opposed to idealized, cortical neurons.

Ignoring for the moment corticothalamic feedback, the most prominent modulatory input to thalamus and cortex is acetylcholine from the brainstem reticular formation (Hallanger et al., 1987). Models have investigated ways in which acetylcholine can control the transition between thalamic relay tonic and burst modes (Sherman, 2001), affect tuning curves of cortical neurons (Soto et al., Kopell, and Sen, 2006), and modulate the relative importance of thalamocortical and intracortical processing (Gil et al., 1997; Hasselmo, 1995; Kimura, 2000, Clarke 2004).

Regional Specialization

Neocortex consists of multiple modules that share substantial architectural properties. The regularity of thalamocortical circuitry has supported decades of suggestions that it may be composed of functionally similar or even identical circuits, differing only, or predominantly, in their afferent sources and efferent targets (Szentagothai, 1975; Hubel and Wiesel, 1977; Creutzfeldt and Nothdurft, 1978; Mountcastle, 1978; Keller and White, 1989; Galuske et al., 2000; Gazzaniga, 2000, Castro- Alamancos and Connors, 1997; Jones, 1998; Heynen and Bear, 2001; Silberberg et al., 2002; Valverde, 2002).

However, the known architectonic differences between regions, in addition to differing statistical properties arising from distinct afferent sources of input, might suggest regional specializations in sensory processing. Thus, characterizing regional differences in computations performed by the thalamocortical circuit is an important and active area of research.

One way in which such differences could be manifested is by differing receptive field properties in thalamus and corresponding cortex. Miller et al. (2001a) propose a model with three different types of functional convergence from thalamus to cortex: in inheritance, a cortical cell’s receptive field is determined by functionally identical thalamic inputs; in constructive convergence, a cortical cell’s receptive field is a composite of many smaller (in spatiotemporal extent) thalamic inputs; in ensemble convergence, the thalamic inputs have some receptive field properties that are not shared by the cortical target cell.

Miller et al. find that in cat auditory thalamus and cortex, all three types of transformation are present, whereas in the visual system constructive convergence seems to predominate (Alonso et al., 2001), and in the rat somatosensory whisker barrel system ensemble convergence may predominate (Simons and Carvell, 1989).

In auditory cortex, there are important differences in anatomy and synaptic physiology from the columnar organization of other sensory cortices (Linden and Schreiner, 2003). There are indications auditory cortex might be specialized for fast temporal information processing (Buonomano, 2000).

Thalamocortical Oscillations

The recurrent connections in the thalamocortical system participate in a prominent feature of thalamocortical circuitry: oscillations across a wide range of frequencies, from as slow as 0.2 Hz to as fast as 80 Hz (Timofeev and Bazhenov, 2005; Buzsáki and Draguhn, 2004). (Some slower, "infra-slow" and faster, "ultra-fast" oscillations seem to not integrally involve thalamocortical connections.) These oscillations can be generated via intrinsic currents in thalamus or cortex, via peripheral input, or via network mechanisms, and are typically synchronized via network mechanisms. Thalamocortical oscillations are thought to play a number of important functional roles, both in sleeping and in waking states, though much about these roles remains unknown.

Slow Oscillations
Figure 2: Cortical slow sleep oscillation in vivo (modified from Timofeev and Bazhenov 2005).

Oscillations in the range of .2-1 Hz are the dominant form of thalamocortical activity seen during slow-wave sleep (Steriade, 2003). Slow waves are generated cortically (Steriade et al., 1993), but thalamocortical neurons are synchronized (Contreras and Steriade, 1995), inhibiting transmission of incoming sensory messages to the cortex (Steriade, 2003).

Less is known about the mechanisms underlying slow waves than about delta waves and spindle waves. Proposed mechanisms for slow wave generation include spontaneous miniature synaptic activities, or "minis" (Fatt and Katz, 1952; Timofeev et al., 2000; Bazhenov et al., 2002), and spontaneous activity of layer V cortical neurons (Sanchez-Vives and McCormick, 2000; Compte et al., 2003). Both processes would induce a transition to the active or UP cortical state.

Two recent large-scale numerical simulations of the thalamocortical system have reproduced many experimentally observed properties of slow waves, including transitions to and from slow-wave sleep.

The simulation of Bazhenov et al. (2002) used a four-layer model (thalamus, thalamic reticular nucleus, inhibitory cortex, pyramidal cortex) with 225 Hodgkin-Huxley cells (single-compartment in thalamus, multiple-compartment in cortex). In this simulation the reexcitation of the cortical network on each cycle is driven by coincidences of minis.

Hill and Tononi (2005) simulated two multi-layer visual cortical areas and their associated thalamic and reticular sectors, with over 65,000 neurons. This model is the first to integrate intrinsic neuronal properties with detailed thalamocortical anatomy and reproduce neural activity patterns in both wakefulness and sleep. In this model an UP state can be initiated by a variety of means, including minis, synaptic input from other cortical and thalamic areas, or intrinsic hyperpolarization-activated \(I_h\) currents.

Evidence exists that Delta oscillations (1-4 Hz) are generated intrinsically by thalamic relay neurons as a result of the interplay between their low-threshold Ca++ current and hyperpolarization-activated cation current (Amzica and Steriade 1998; McCormick and Pape, 1990).

Both slow and delta oscillations are thought to participate in consolidation of memories acquired during wakefulness (Gais et al., 2000; Stickgold et al., 2000; Maquet 2001; Huber et al., 2004; Steriade and Timofeev, 2003).

Based on analyses of multiple extracellular recordings of slow oscillations during natural sleep, it has been suggested that fast oscillations during active states of slow-wave sleep could reflect recalled events experienced previously, directly "imprinting" these memories in the network via synchronized events that are observable as slow-wave components in the EEG (Destexhe et al., 1997).


Spindle oscillations (7-14 Hz) consist of waxing and waning field potentials at 7-14 Hz, typically lasting 1-3 seconds and recurring roughly every 5-15 seconds. In vivo, spindle oscillations are typically observed during early stages of sleep or during active phases of slow-wave sleep oscillations. They are generated thalamically (Morison and Bassett 1945), critically involving the thalamic reticular nucleus (Steriade et al., 1985; Steriade and Deschênes, 1984, von Krosigk et al., 1993); burst firing of thalamocortical neurons in turn excite reticular neurons, maintaining the cycle, with corticothalamic feedback involved in synchronizing these oscillations (Destexhe et al., 1998; Destexhe et al., 1999). Modeling studies (Bazhenov et al. 1998; Destexhe et al., 1996; Destexhe and Sejnowski, 1997) have reproduced these features.

As with slow and delta oscillations, spindle oscillations are implicated in memory consolidation and demonstrate short- and intermediate-term synaptic plasticity (Gais et al., 2000; Steriade and Timofeev, 2003).

Fast Oscillations

Waking states of the brain can be characterized by a predominance of relatively high-frequency oscillations, notably Beta (15-30 Hz) and Gamma (30-80 Hz) oscillations. Gamma activity is associated with attentiveness, focused arousal, sensory perception, movement, and prediction. (See Beta-gamma oscillation for references.) Gamma-range synchronous activity has been proposed to be related to cognitive processing, and may transiently synchronize cells with disparate receptive fields. That synchrony has been hypothesized to allow multiple features of a cue to be assembled into a coherent representation. (See Singer 1998 for review.)

It is worth noting that beta oscillations have been selectively induced in hippocampal slices (Boddeke et al., 1997; Shimono et al., 2000), and beta and gamma oscillations are reported to have different synchronization properties (Kopell et al., 2000).

A large-scale model by Traub et al. (2005), incorporating 3,560 multi-compartment thalamic, reticular, and cortical neurons, replicates persistent gamma oscillations. In this model electrical coupling between axons is necessary for persistent gamma.

Spike-and-Wave Oscillations

During epileptic seizures, another pattern of oscillation, spike-and-wave oscillation, is observed. These oscillations occur at a frequency of about 3 Hz in humans. The "spike" in the EEG pattern is known to be related to cortical cell firing, the "wave" to cortical cell silence. One key element in thalamocortical models of spike-and-wave oscillations is the switching of the thalamus to a slow 3 Hz oscillation by excessive corticothalamic feedback.

Recurrent Computation

In principle, corticothalamic feedback could serve to modulate response properties of thalamus, or to induce and synchronize large-scale oscillations, or to relay specific information back to thalamus for more complex computational processing. Evidence suggests that all three processes occur.

Corticothalamic Modulation

Corticothalamic feedback from layer VI provides modulatory input to thalamic relay cells (Jones, 2007, p. 289), as does brainstem cholinergic input (and, more locally, thalamic reticular nucleus GABAergic input), affecting the structure of their receptive fields. Sherman and Guillery have proposed that, in awake sensory processing, burst transmission can serve as a "wake up call" to cortex, activating a set of cortical columns which would then provide feedback switching the relay mode to tonic, which provides more linear signal transfer (Sherman, 2001; Sherman and Guillery, 2006).

The corticothalamic collateral input to the thalamic reticular nucleus is stronger than the direct input to relay cells (Golshani et al, 2001), emphasizing the modulatory aspect of the feedback. There is evidence that projections from some areas of prefrontal cortex terminate widely in the TRN, rather than only in directly associated reticular and thalamic territories (Zikopoulos and Barbas, 2006). This suggests an additional role for corticothalamic modulation: prefrontal areas may participate in attentional regulation of relevant sensory signals, by gating thalamic output back to cortex.

Oscillation-Assisted Processing

Closed-loop neuronal computations occur throughout the nervous system, including the thalamocortical system. Such loops may be viewed from either a homeostatic or a computational point of view. Feedback loops provide a mechanism for neural ensembles to maintain a set of variables within a given range; this can be seen as a homeostatic control process. From a computational perspective, the sequence of changes to the variable state values can be viewed as an encoding of the sensory input driving the homeostatic corrections (Ahissar and Kleinfeld, 2003).

Several studies have investigated correlations between whisking behavior and thalamocortical oscillations in the rat barrel cortex system. This mode of sensory processing is used during "active discrimination", as a function of overall behavioral state (Nicolelis, 2005). Thus, the entire feedback loop is under external modulatory control. Models suggest that a computational function of this loop is to transform temporally encoded vibrissal information into a rate code, by means of phase-locked loops (Ahissar et al., 2000; Ahissar et al., 1997). There is evidence that this fundamental pattern may be a more general property of sensory systems. In human speech perception, comprehension is enhanced when the temporal envelope frequency of the speech signal is similar to cortical activity frequency, and when there is phase locking between the two temporal envelopes (Ahissar et al., 2001).

A broader perspective of all sensory stimuli serving to modulate ongoing, self-generated brain activity has been put forward as an alternative to the traditional "feedforward" view (Llinás and Paré, 1991). This view is supported by several similarities in paradoxical or REM sleep and wakefulness, vs. other sleep states. On this view dreaming is what happens when the intrinsic functional realm of wakefulness is deprived of modulatory sensory input (and brainstem-mediated muscular atonia is present).

Computational Interpretations of Activity Patterns

Candidate computational roles have been proposed for the integrative action of thalalmocortical loops, with regard to the patterns of cortical activity that may occur over time in response to natural afferent stimulation. Earliest cortical activity occurs in middle and superficial layers in response to peripheral input via direct or core thalamic nuclei. Lateral inhibition in superficial layers generates IPSPs that are substantially longer than EPSPs; thus initial excitatory responses are rapidly inhibited, and only those excitatory cells that are most activated by an input pattern can respond at all before lateral inhibition quiets them. With synaptic potentiation of the kind described above (see Neocortical synaptic potentiation), superficial cells that initially respond to a particular thalamic input pattern become increasingly responsive not only to that input but also to a range of similar inputs, such that similar but distinguishable inputs will come to elicit identical patterns of output from layer II-III cells. Results of this kind have been obtained in a number of different models with related characteristics (von der Malsburg, 1973; Grossberg, 1976; Rumelhart, 1985; Coultrip et al., 1992).

Superficial layer responses activate deep layers, and output from layer VI initiates feedback activation of thalamic reticular nucleus (TRN), which in turn inhibits the portions of the core nucleus corresponding topographically to those portions of layer II-III that were active. On the next cycle of thalamocortical activity, the input will arrive at the core nucleus against the background of the inhibitory feedback from TRN, which has been shown to last for hundreds of milliseconds (Huguenard and Prince, 1994; Cox et al., 1997; Zhang et al., 1997).

Thus, in a series of modeling experiments, the predominant component of the next input to cortex is just the uninhibited remainder of the input, whereupon the same operations as before are performed. The result is that the second cortical response will consist of a quite distinct set of neurons from the initial response, since most of the input components giving rise to that first response are now inhibited. Analysis of the second (and ensuing) responses in computational models has shown successive sub-clustering of an input: the first cycle of response identifies the input’s membership in a general category of similar stimuli, the next response (a fraction of a second later) identifies its membership in a particular subcluster, then sub-sub-cluster, etc. (Rodriguez et al., 2004).

In contrast to the topography-preserving projections between core thalamic cells and cortex, the nontopographic projections from layer V to matrix cells bypass the thalamic reticular nucleus (see, e.g., Bourassa and Deschenes, 1995; Deschenes et al., 1998). These non-topography-preserving projections have been interpreted as orthogonalizing their inputs rather than clustering them: i.e., any structural relationships that may obtain among inputs are not retained in the resulting projections. Thus even cortical response patterns that are similar to each other may generate very different patterns in their projections to thalamic matrix cells.

If the thalamic input is changing over time, then the activation of layer V in rapid sequence via superficial layer inputs (in response to an element of a sequence) and via thalamic matrix inputs (corresponding to feedback from the previous element in a sequence) selects responding cells sparsely from the most activated cells in the layer (Coultrip et al., 1992) and selects synapses on those cells sparsely as a function of the sequential pattern of arriving inputs. Thus synapses potentiated at a given time in layer V correspond to the input occurring at that time together with orthogonalized feedback arising from input just prior to that time (Aleksandrovsky et al., 1996; Granger et al., 1994). In modeling experiments, the overall effect is "chaining" of elements in the input sequence, via the "links" created due to layer V activity from coincident inputs corresponding to current and prior input elements (Rodriguez et al., 2004; Granger 2006).

Large-Scale Simulation

Three very large-scale simulation efforts are noteworthy. The current state of the art in thalamocortical simulation is a model by Traub et al. (2005), comprising 3,650 multi-compartment (~100 compartments) neurons of several types from all cortical layers, thalamus, and thalamic reticular nucleus. This model exhibits persistent gamma oscillations, sleep spindles, synchronized population bursts resembling seizures, and ripples.

An even more ambitious project, the Blue Brain Project, is under way (Markram, 2006). This project will simulate an entire two-week-old rat neocortical somatosensory column, comprising some 10,000 neurons, with thousands of compartments each, and more than a dozen Hodgkin-Huxley ion channels per compartment. Blue Brain will use the IBM Blue Gene/L supercomputer architecture, and a large (currently private) database of columnar connectivity data gathered by the Markram laboratory over the past decade.

Izhikevich (2005) has simulated a network of \(10^{11}\) neurons and \(10^{15}\) synapses -- the size of the entire human brain. The simulation used microcircuitry based on quantitative anatomical studies of cat visual cortex (Binzegger et al., 2004) and of thalamic circuitry, employing a very efficient approximation to more detailed traditional spiking neuron models (Izhikevich, 2003). One second of simulated time required 50 days on a cluster of 27 3GHz processors.

Cortico-basal ganglia-thalamocortical Loops

A discussion of the function of the thalamocortical system would not be complete without consideration of those thalamic nuclei which integrally involve the basal ganglia in their thalamocortical loop (primarily ventral lateral anterior, ventral anterior, medial dorsal, and centré median). Often the basal ganglia inputs to thalamus are treated as just one more subcortical input, on a par with any other. But several features set them apart. First, the basal ganglia inputs (from globus pallidus, pars interna (GPi), and substantia nigra, pars reticulata (SNr)) are GABAergic, and thus inhibitory, unlike other inputs (apart from the thalamic reticular nucleus). Second, there are several closed cortico-striato-pallido-thalamocortical loops (Alexander et al. 1986, Middleton et al., 1996). These may coarsely be divided into sensorimotor, associational, and limbic loops (Parent et al., 1995). Thus, cortex is integrally involved in the basal ganglia inputs to thalamus, so these should not be viewed as directly analogous to primary sensory inputs.

Essentially all of cortex, both allocortex and isocortex (except, possibly, primary visual cortex), projects to striatum, via collaterals from the layer V projection to brainstem and motor targets (Swanson, 2000). (These same fibers are the source of the thalamic matrix innervation, in the core vs. matrix perspective, and the higher-order thalamic relay innervation, in the first-order vs. higher-order relays perspective, above.) The basal ganglia-thalamo-cortical pathway does not reciprocally innervate the entire cortex, however. Traditionally the motor output is emphasized, but more broadly basal ganglia is seen as affecting much of anterior cortex (e.g. Alexander et al. 1986). There do, however, seem to be some posterior cortical regions targeted as well (Middleton et al., 1996, Clower et al. 2005).

Reinforcement Learning

A popular view of basal ganglia function is that it performs reinforcement learning. The basal ganglia input to thalamus affects cortical states and subsequent behavior, which leads to increased or decreased reward signals via the dopaminergic nigrostriatal pathway. Striatal plasticity then adjusts the response of basal ganglia to cortical input, so that rewarding actions are reinforced, and nonrewarding actions are inhibited (Schultz 1997, 1998).

Comparisons have been drawn between the basal ganglia and computational models of reinforcement learning, specifically temporal-difference actor-critic models (Barto, 1995; Houk et al., 1995; Sutton and Barto, 1998; Suri and Schultz, 1998, 1999). Such models perhaps best represent the common current view on basal ganglia function. However, these models have been criticized on grounds that they either fail to reproduce the observed dopamine signal in some cases, that they have not been linked with brain structures supporting the required temporal difference signals, or both (Brown et al. 1999; Joel et al. 2002); see also Reinforcement Learning and Models of Basal Ganglia.

More fundamentally, it is unclear how exactly actions might be represented in the basal ganglia: a model in which individual behavioral actions are targeted topographically is highly problematic outside of primary motor cortex. If a notion of action in the formal reinforcement learning sense is to be applied to the basal ganglia, it seems that the picture that must be adopted is that the environment affected by the actions is not the organism's musculature, as is traditionally assumed in such models, but rather the thalamocortical system itself.

Basal Ganglia as Modulator of Cortical Functioning

Consistent with the above considerations, Sherman and Guillery (2006) propose on anatomic and physiological grounds that basal ganglia input is modulatory. Specifically, they propose that basal ganglia inputs modulate relay properties of higher-order relays. For example, the GPi inputs to the ventral anterior and ventral lateral nuclei would effectively serve to gate cortico-thalamo-cortical signals from layer V of motor cortex to input layers of premotor cortex. For an evolutionary point of view, see Granger, Behav & Brain Sci., (2006)

A related modulatory or gating view of basal ganglia output is that expressed by Houk (1995): basal ganglia inputs to thalamus serve to register or negate sensory contexts into working memory, by pushing a bistable layer VI feedback loop into one of two dynamical regimes.

References

  • Abramson, B. P. and L. M. Chalupa (1985). The laminar distribution of cortical connections with the tecto- and cortico-recipient zones in the cat's lateral posterior nucleus. Neuroscience 15(1):81-95.
  • Ahissar, E., S. Haidarliu, and M. Zacksenhouse (1997). Decoding temporally encoded sensory input by cortical oscillations and thalamic phase comparators. Proc. Natl. Acad. Sci. USA 94:11633–11638.
  • Ahissar, E., R. Sosnik R, and S. Haidarliu (2000). Transformation from temporal to rate coding in a somatosensory thalamocortical pathway. Nature 406(6793): 302-306.
  • Ahissar, E., S. Nagarajan, M. Ahissar., A. Protopapas, H. Mahncke, and M. M. Merzenich (2001). Speech comprehension is correlated with temporal response patterns recorded from auditory cortex. Proc. Natl. Acad. Sci. USA 98(23): 13367–13372
  • Ahissar, E. and David Kleinfeld (2003). Closed-loop Neuronal Computations: Focus on Vibrissa Somatosensation in Rat. Cerebral Cortex 13: 53–62.
  • Ahmed, B., J. C. Anderson, K. A. C. Martin, and J. C. Nelson (1997). Map of the synapses onto layer 4 basket cells of the primary visual cortex of the cat. J Comp Neurol 380: 230-242.
  • Alonso, J-M., W. M. Usrey, and R. C Reid (2001). Rules of Connectivity between Geniculate Cells and Simple Cells in Cat Primary Visual Cortex. The Journal of Neuroscience, 21(11): 4002-4015.
  • Alexander, G., M. DeLong, and P. Strick (1986). Parallel Organization of Functionally Segregated Circuits Linking Basal Ganglia and Cortex. Annual Review of Neuroscience 9: 357-381.
  • Aleksandrovsky, B., J. Whitson, A. Garzotto, G. Lynch, and R. Granger (1996). An algorithm derived from thalamocortical circuitry stores & retrieves temporal sequences. IEEE Int'l Conf. Patt. Rec., 550-554.
  • Ambros-Ingerson, J,. R. Granger, and G. Lynch (1990). Simulation of paleocortex performs hierarchical clustering. Science 247: 1344-1348.
  • Amzica, F. and M. Steriade (1998). Electrophysiological correlates of sleep delta waves. Electroencephalogy Clin. Neurophysiol. 107: 69-83.
  • Attneave, F. (1954). Some informational aspects of visual perception. Psychol. Rev.61(3): 183-93.
  • Barlow, H. B. (1961). Possible principles underlying the transformation of sensory messages. In Sensory Communication. Cambridge, MA: MIT Press.
  • Barto, A. G. (1995). Adaptive Critics and the Basal Ganglia. In Models of Information Processing in the Basal Ganglia. Cambridge, MA: MIT Press.
  • Bazhenov, M., I. Timofeev, M. Steriade, and T. J. Sejnowski (1998). Computational Models of Thalamocortical Augmenting Responses. The Journal of Neuroscience, 18(16): 6444–6465.
  • Bazhenov, M., I. Timofeev, M. Steriade, and T. J. Sejnowski (2002). Model of thalamocortical slow-wave sleep oscillations and transitions to activated states. J. Neurosci. 22: 8691-8704.
  • Bear, M. F. and A. Kirkwood (1993). Neocortical long-term potentiation. Curr. Opin. Neurobiol. 3:.197-202.
  • Bekkers, J. and C. Stevens (1989). NMDA and non-NMDA receptors are co-localized at individual excitatory synapses in cultured rat hippocampus. Nature 341: 230-233.
  • Bell, A. J. and T. J. Sejnowski (1997). The Independent Components of Natural Scenes are Edge Filters. Vision Research 37(23): 3327-3338.
  • Bender, D. (1983). Visual activation of neurons in the primate pulvinar depends on cortex but not colliculus. Brain Res 279: 258-261.
  • Binzegger, T., R. J. Douglas, and K. A. C. Martin (2004). A Quantitative Map of the Circuit of Cat Primary Visual Cortex. The Journal of Neuroscience 24(39): 8441– 8453
  • Blasco-Ibanez, J. M., F. J. Martinez-Guijarro, and T. F. Freund (1998) Enkephalin-containing interneurons are specialized to innervate other interneurons in the hippocampal CA1 region of the rat and guinea-pig. Eur J Neurosci 10: 1784-1795.
  • Boddeke, H. W. G. M., R. Best, and P. H. Boeijinga (1997). Synchronous 20 Hz rhythmic activity in hippocampal networks induced by activation of metabotropic glutamate receptors in vitro. Neuroscience 76: 653– 658.
  • Boloori, A.-R. and G. B. Stanley (2006). The Dynamics of Spatiotemporal Response Integration in the Somatosensory Cortex of the Vibrissa System. The Journal of Neuroscience, 26(14): 3767–3782.
  • Bourassa, J. and M. Deschenes (1995). Corticothalamic projections from the primary visual cortex in rats: a single fiber study using biocytin as an anterograde tracer. Neuroscience 66: 253-263.
  • Brown, J., D. Bullock, and S. Grossberg. (1999). How the Basal Ganglia Use Parallel Excitatory and Inhibitory Learning Pathways to Selectively Respond to Unexpected Rewarding Cues. The Journal of Neuroscience 19(23):10502-10511.
  • Braitenberg, V. and A. Schüz (1998). Cortex: statistics and geometry of neuronal connectivity, NY: Springer.
  • Buonomano, D. V. (2000). Decoding Temporal Information: A Model Based on Short-Term Synaptic Plasticity. The Journal of Neuroscience, 20(3): 1129-1141.
  • Buonomano, D. and M. Merzenich (1998) Cortical plasticity: synapses to maps. Ann. Rev. Neurosci. 21: 149-186.
  • Bush, P., and T. Sejnowski (1996). Inhibition synchronizes sparsely connected cortical neurons within and between columns in realistic network models. J. Comput. Neurosci. 3: 91-110.
  • Buzsáki, G. and A. Draguhn (2004). Neuronal Oscillations in Cortical Networks. Science, 304(5679): 1926-1929.
  • Canales, J., C. Capper-Loup, D. Hu, E. Choe, U. Upadhyay, and A. Graybiel (2002). Shifts in striatal responsivity evoked by chronic stimulation of dopamine and glutamate systems. Brain 125: 2353-2363.
  • Castro-Alamancos, M. and B. Connors (1997). Thalamocortical synapses. Prog Neurobiol 51: 581-606.
  • Castro-Alamancos, M., J. Donoghue, and B. Connors (1995). Different forms of synaptic plasticity in somatosensory and motor areas of neocortex. J. Neurosci. 15: 5324-5333.
  • Chrobak, J. and G. Buzsaki (1998). Gamma oscillations in the entorhinal cortex of the freely behaving rat. J Neurosci 18: 388-398.
  • Clarke, P. B. (2004). Nicotinic modulation of thalamocortical neurotransmission. Prog. Brain Res. 145: 253–260.
  • Clower, D., R. Dum, and P. Strick (2005). Basal Ganglia and Cerebellar Inputs to ‘AIP’. Cerebral Cortex 15(7): 913-920.
  • Compte, A., M. V. Sanchez-Vives, D. A. McCormick, and X. J. Wang (2003). Cellular and network mechanisms of slow oscillatory activity (<1 Hz) and wave propagations in a cortical network model. J. Neurophysiol. 89: 2707-2725.
  • Conley, M. and I. T. Diamond (1990). Organization of the Visual Sector of the Thalamic Reticular Nucleus in Galago. Eur. J. Neurosci. 2: 211-226.
  • Contreras, D. and M. Steriade. (1995) Cellular basis of EEG slow rhythms: a study of dynamic corticothalamic relationships. J. Neurosci. 15(1 Pt 2): 604-22.
  • Coultrip, R., R. Granger, and G. Lynch (1992). A cortical model of winner-take-all competition via lateral inhibition. Neural Networks 5(1): 47-54.
  • Cox, C. L., J. R. Huguenard, and D. A. Prince (1997). Nucleus reticularis neurons mediate diverse inhibitory effects in thalamus. Proc. Natl. Acad. Sci. 94: 8854-8859.
  • Cragg, B. (1967). The density of synapses and neurones in the motor and visual areas of the cerebral cortex. J. Anat. 101: 639-654.
  • Creutzfeldt, O. and H. Nothdurft (1978). Representation of complex visual stimuli in the brain. Naturwissenschaften 65: 307-318.
  • Das, S., Y. Sasaki, T. Rothe, L. Premkumar, M. Takasu, J. Crandall, P. Dikkes, D. Conner, P. Rayudu, W. Cheung, H. Chen, S. Lipton, and N. Nakanishi (1998). Increased NMDA current and spine density in mice lacking the NMDA receptor subunit NR3A. Nature 393: 377-381.
  • DeFelipe, J. and E. Jones (1991). Parvalbumin immunoreactivity reveals layer IV of monkey cerebral cortex as a mosaic of microzones of thalamic afferent terminations. Brain Res 562: 39-47.
  • Deschênes, M., P. Veinante, and Z-W. Zhang (1998). The organization of corticothalamic projections: reciprocity versus parity. The Journal of Comparative Neurology 424(2): 197-204.
  • Destexhe, A., D. Contreras, M. Steriade,T. J. Sejnowski, and J. R. Huguenard (1996). In Vivo, ln Vitro, and Computational Analysis of Dendritic Calcium Currents in Thalamic Reticular Neurons. The Journal of Neuroscience, 16(1): 169-185.
  • Destexhe, A. and T. J. Sejnowski (1997). Synchronized Oscillations in Thalamic Networks: Insights from Modeling Studies. In Thalamus, Edited by M. Steriade, E. G. Jones, and D. A. McCormick. Elsevier.
  • Destexhe, A., D. Contreras and M. Steriade (1998). Mechanisms underlying the synchronizing action of corticothalamic feedback through inhibition of thalamic relay cells. J. Neurophysiol. 79: 999-1016.
  • Destexhe, A., D. Contreras, and M. Steriade (1999). Cortically-induced coherence of a thalamic-generated oscillation. Neuroscience 92: 427-443.
  • Di, S., C. Baumgartner, and D. S. Barth (1990). Laminar analysis of extracellular field potentials in rat vibrissa/barrel cortex. J. Neurophysiol. 63: 832-840.
  • Diamond, I. T., E. G. Jones, and T. P. S. Powell (1969). The projection of the auditory cortex upon the diencephalon and brain stem in the cat. Brain Res. 15: 305-340.
  • Diamond, M. E., M. Armstrong-James, F. F. Ebner (1992a). 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.
  • Diamond, M., M. Armstrong-James, M. Budway, and F. Ebner (1992b). Somatic sensory responses in the rostral sector of the posterior group (POm) and the ventral posterior medial nucleus (VPM) of the rat thalamus: dependence on the barrel field cortex. J Comp Neurol 319: 66-84.
  • Fatt, P. and B. Katz (1952). Spontaneous sub-threshold activity at motor-nerve endings. Journal of Physiology 117: 109-128.
  • Felch A, Granger R (2008). The hypergeometric connectivity hypothesis: Divergent performance of brain circuits with different synaptic connectivity distributions. Brain Research, 1202: 3-13.
  • Felleman, D. and D. Van Essen (1991). Distributed Hierarchical Processing in the Primate Cerebral Cortex. Cerebral Cortex 1(1):1-47.
  • Field, D. J. (1987). Relations between the statistics of natural images and the response properties of cortical cells. J. Opt. Soc. Am. A 4(12): 2379-2394.
  • Fitzpatrick, D., J. Lund, D. Schmechel, and A. Towles (1987). Distribution of gabaergic neurons and axon terminals in the macaque striate cortex. J Comp Neurol 264: 73-91.
  • Freund, T. F., K. A. Martin, and D. Whitteridge (1985). Innervation of cat visual areas 17 and 18 by physiologically identified X- and Y-type thalamic afferents. I. Arborization patterns and quantitative distribution of postsynaptic elements. J Comp Neurol 242: 263-274.
  • Freund, T. F., K. A. Martin, I. Soltesz, P. Somogyi, and D. Whitteridge (1989). Arborisation pattern and postsynaptic targets of physiologically identified thalamocortical afferents in striate cortex of the macaque monkey. J Comp Neurol 289: 315-336.
  • Freund, T. F. and V. Meskenaite (1992). gamma-Aminobutyric acid-containing basal forebrain neurons innervate inhibitory interneurons in the neocortex. Proc Natl Acad Sci USA 89: 738-742.
  • Fries P., S. Neuenschwander, A. Engel, R. Goebel, and W. Singer (2001). Rapid feature selective neuronal synchronization through correlated latency shifting. Nat Neurosci 4: 194-200.
  • Gais, S., W. Plihal, U. Wagner, and J. Born (2000). Early sleep triggers memory for early visual discrimination skills. Nat. Neurosci. 3: 1335-1339.
  • Galuske, R. A., W. Schlote, H. Bratzke, and W. Singer (2000). Interhemispheric asymmetries of the modular structure in human temporal cortex. Science 289: 1946-1949.
  • Gazzaniga, M. S. (2000). Regional differences in cortical organization. Science 289: 1887-1888.
  • Gil, Z., B. Connors, and Y. Amitai. Differential modulation of neocortical synapses by neuromodulators and activity. Neuron 19: 679 – 686.
  • Golshani, P., X.-B. Liu, and E. G. Jones (2001). Differences in quantal amplitude reflect GluR4- subunit number at corticothalamic synapses on two populations of thalamic neurons. Proc Natl Acad Sci USA 98(7):4172-4177.
  • Granger, R., J. Whitson, J. Larson, and G. Lynch (1994). Non-Hebbian properties of long-term potentiation enable high-capacity encoding of temporal sequences. Proc. Natl. Acad. Sci. 91: 10104-10108.
  • Granger R (2006). The evolution of computation in brain circuitry. Behav Brain Sci 29: 17-18.
  • Granger R (2006). Engines of the brain: The computational instruction set of human cognition. AI Magazine 27: 15-32.
  • Grossberg, S. (1976). Adaptive pattern classification and universalrecoding. Biol Cybern 23:121-134.
  • Guido, W., S. Lu, and S. Sherman (1992). Relative contributions of burst and tonic responses to receptive field properties of lateral geniculate neurons in the cat. J. Neurophysiol. 68: 2199- 2211.
  • Guido, W., S. M. Lu, J. W. Vaughan, D. W. Godwin, and S. M. Sherman (1995). Receiver operating characteristic (ROC) analysis of neurons in the cat's lateral geniculate nucleus during tonic and burst response mode. Vis. Neurosci. 12: 723-741.
  • Guillery, R. W. and K. Herrup (1997). Quantification without pontification: choosing a method for counting objects in sectioned tissues. J. Comp. Neurol. 386: 2-7.
  • Gulyas, A. I., N. Hajos, and T. F. Freund (1996). Interneurons containing calretinin are specialized to control other interneurons in the rat hippocampus. J Neurosci 16: 3397-3411.
  • Gulyas, A., M. Megias, Z. Emri, and T. Freund (1999). Total number and ratio of excitatory and inhibitory synapses converging onto single interneurons of different types in the CA1 area of the rat hippocampus. J. Neurosci. 19: 10082-10097.
  • Hallanger, A. E., A. I. Levey, H. J. Lee, D. B. Rye, and B. H. Wainer (1987). The origins of cholinergic and other subcortical afferents to the thalamus in the rat. J. Comp. Neurol. 262(1):105-124.
  • Hasselmo, M. (1995). Neuromodulation and cortical function: modeling the physiological basis of behavior. Behav. Brain Res. 67: 1–27.
  • Hendry, S., H. Schwark, E. Jones, and J. Yan (1987). Numbers and proportions of GABA-immunoreactive neurons in different areas of monkey cerebral cortex. J Neurosci 7: 1503-1519.
  • Herkenham, M. (1986). New perspecties on the organization and evolution of nonspecific thalamocortical projections. In Cerebral Cortex (E. Jones, A. Peters, ed). NY: Plenum.
  • Hess, G., C. Aizenman, and J. Donoghue (1996). Conditions for the induction of long-term potentiation in layer II/III horizontal connections of the rat motor cortex. J. Neurophysiol. 75: 1765-1778.
  • Heynen, A. J. and M. F. Bear (2001). Long-term potentiation of thalamocortical transmission in the adult visual cortex in vivo. J. Neurosci. 21: 9801-9813.
  • Hilgetag, C. C., M. O'Neill, and M. Young (1996). Indeterminate Organization of the Visual System. Science 271: 776.
  • Hill, S. and G. Tononi (2005). Modeling Sleep and Wakefulness in the Thalamocortical System. J. Neurophysiol. 93: 1671-1698.
  • Hirsch, J. C. and F. Crepel (1990). Use-dependent changes in synaptic efficacy in rat prefrontal neurons in vitro. J. Physiol. 427:31-49.
  • Houk, J. C. (1995). Information Processing in Modular Circuits Linking Basal Ganglia and Cerebral Cortex. In Models of Information Processing in the Basal Ganglia. Cambridge, MA: MIT Press.
  • Houk, J. C, J. L. Adams, and A. G. Barto (1995). A Model of How the Basal Ganglia Generate and Use Neural Signals that Predict Reinforcement. In Models of Information Processing in the Basal Ganglia. Cambridge, MA: MIT Press.
  • Hubel, D. H. and T. N. Wiesel (1977). Functional architecture of macaque monkey visual cortex. Proc. R. Soc. Lond. B Biol. Sci. 198: 1-59.
  • Huber, R., M. F. Ghilardi, M. Massimini, and G. Tononi (2004). Local sleep and learning. Nature 430: 78-81.
  • Huguenard, J. and D. Prince (1994). Clonazepam suppresses GABAB-mediated inhibition in thalamic relay neurons through effects in nucleus reticularis. J Neurophysiol 71: 2576-2581.
  • Hull, C., and M. Scanziani (2007). It's about time for thalamocortical circuits. Nature Neuroscience 10(4): 400-402.
  • Iriki, A., C. Pavlides, A. Keller, and H. Asanuma (1991). Long-term potentiation of thalamic input to the motor cortex induced by coactivation of thalamocortical and corticocortical afferents. J. Neurophysiol. 65: 1435-1441.
  • Izhikevich, E. M. (2003). Simple Model of Spiking Neurons. IEEE Trans. on Neural Networks 14(6): 1569-1572.
  • Izhikevich, E. M. (2005). Large-Scale Simulation of the Human Brain. http://vesicle.nsi.edu/users/izhikevich/human_brain_simulation/Blue_Brain.htm .
  • Joel, D., Y. Niv, and E. Ruppin (2002). Actor – critic models of the basal ganglia: new anatomical and computational perspectives. Neural Networks 15: 535–547.
  • Jones, E. G. (1981). Functional subdivision and synaptic organization of the mammalian thalamus. Int. Rev. Physiol. 25: 173-245.
  • Jones, E. G. (1998). A new view of specific and nonspecific thalamocortical connections. Adv Neurol 77: 49-71.
  • Jones, E. G. (2001). The thalamic matrix and thalamocortical synchrony. Trends Neurosci. 24(10): 595-601.
  • Jones, E. G. (2007). The Thalamus. 2nd edition. Cambridge, U.K.: Cambridge University Press.
  • Jones, E. G. and S. H. C. Hendry (1989). Differential calcium binding protein immunoreactivity distinguishes classes of relay neurons in monkey thalamic nuclei. Eur.J. Neurosci. 1: 222-246.
  • Keller, A. and E. White E (1989). Triads: a synaptic network component in cerebral cortex. Brain Res 496: 105-112.
  • Kenan-Vaknin, G. and T. Teyler (1994). Laminar pattern of synaptic activity in rat primary visual cortex: comparison of in vivo and in vitro studies employing current source density analysis. Brain Res. 635: 37-48.
  • Killackey, H. and F. Ebner (1973). Convergent projection of three separate thalamic nuclei on to a single cortical area. Science 179: 283-285.
  • Killackey, H. P. and F. F. Ebner (1972). Two different types of thalamocortical projections to a single cortical area in mammals. Brain Behav Evol 6: 141-169.
  • Kim, H., K. Fox, and B. Connors (1995) Properties of excitatory synaptic events in neurons of primary somatosensory cortex of neonatal rats. Cereb Cortex 5: 148-157.
  • Kimura, F. (2000). Cholinergic modulation of cortical function: a hypothetical role in shifting the dynamics in cortical network. Neurosci. Res. 38: 19 –26.
  • Kimura, A., A. Caria, F. Melis, H. Asanuma (1994). Long-term potentiation within cat motor cortex. Neuroreport 5: 2372-6.
  • Kirkwood, A., S. Dudek, J. Gold, C. Aizenman, and M. Bear (1993). Common forms of synaptic plasticity in the hippocampus and neocortex in vitro. Science 260: 1518-1521.
  • Knoblauch, A. and G. Palm (2002). Scene segmentation by spike synchronization in reciprocally connected visual areas. I. Local effects of cortical feedback. Biol. Cybern. 87:151-167.
  • Komatsu, Y., K. Fujii, J. Maeda, H. Sakaguchi, and K. Toyama (1988). Long-term potentiation of synaptic transmission in kitten visual cortex. J. Neurophysiol. 59: 124-141.
  • Kopell, N., G. B. Ermentrout, M. A. Whittington, and R. D. Traub (2000). Gamma rhythms and beta rhythms have different synchronization properties. Proc. Natl. Acad. Sci. USA 97: 1867-1872.
  • Krukowski, A. E. and K. D. Miller (2001). Thalamocortical NMDA conductances and intracortical inhibition can explain cortical temporal tuning. Nat. Neurosci. 4:424-430
  • Kudoh, M. and K. Shibuki (1996) Long-term potentiation of supragranular pyramidal outputs in the rat auditory cortex. Exp. Brain Res. 110: 21-27.
  • Lewicki, M. S. (2002). Efficient Coding of Natural Sounds. Nature Neuroscience 5(4):356-363.
  • Linden, J. F. and C. E. Schreiner (2003). Columnar Transformations in Auditory Cortex? A Comparison to Visual and Somatosensory Cortices. Cerebral Cortex 13(1): 83-89.
  • Liu, X. B. and E. G. Jones (1999). Predominance of corticothalamic synaptic inputs to thalamic reticular nucleus neurons in the rat. Journal of Comparative Neurology 414: 67-79.
  • Llinás, R. R. and D. Paré (1991). Of Dreaming and Wakefulness. Neuroscience 44(3):521-535.
  • Lorente de No, R. (1938). Cerebral cortex: Architecture, intracortical connections, motor projections. In Physiology of the nervous system (J. Fulton, ed), pp 291-340. London: Oxford.
  • Magee, J. C. (2000). Dendritic integration of excitatory synaptic input. Nat. Rev. Neurosci. 1: 181-190.
  • Magee, J. C. and E. P. Cook (2000). Somatic EPSP amplitude is independent of synapse location in hippocampal pyramidal neurons. Nat. Neurosci. 3: 895-903.
  • Maquet, P. (2001). The role of sleep in learning and memory. Science 294: 1048-1052.
  • Markram, H. (2006). The Blue Brain Project. Nature Neuroscience 7(2): 153-160.
  • McCarley, R., J. Winkelman, and F. Duffy (1983). Human cerebral potentials associated with REM sleep rapid eye movements: links to PGO waves and waking potentials. Brain Res. 274: 359-364.
  • McCormick, D. and T. Bal (1994). Sensory gating mechanisms of the thalamus. Curr Opin Neuro 4: 550-556.
  • McCormick, D. and H. Feeser (1990). Functional implications of burst firing and single spike activity in lateral geniculate relay neurons. Neuroscience 39: 103-113.
  • McCormick, D. A. and H. C. Pape (1990). Noradrenergic and serotonergic modulation of a hyperpolarization-activated cation current in thalamic relay neurones. J. Physiol. 431: 319-342.
  • Middleton, F. and P. Strick (1996). The temporal lobe is a target of output from the basal ganglia. Proc. Natl. Acad. Sci. 93(16): 8683-8687.
  • Miller, L., M. Escabí, H. Read, and C. Schreiner (2001a). Functional Convergence of Response Properties in the Auditory Thalamocortical System. Neuron 32(1): 151-160.
  • Miller, K. D., D. J. Pinto, and D. J. Simons (2001b). Processing in layer 4 of the neocortical circuit: new insights from visual and somatosensory cortex. Current opinion in neurobiology 11(4): 488-497.
  • Molinari, M., M. Dell'Anna, E. Rausell, M. Leggio, T. Hashikawa, and E. Jones (1995). Auditory thalamocortical pathways defined in monkeys by calcium-binding protein immunoreactivity. J Comp Neur 362: 171-194.
  • Morison, R. S. and D. L. Bassett (1945). Electrical activity of the thalamus and basal ganglia in decorticate cats. J. Neurophysiol. 8: 309-314.
  • Mountcastle, V. B. (1957). Modality and topographic properties of single neurons of cat's somatic sensory cortex. J Neurophysiol 20: 408-434.
  • Mountcastle, V. B. (1978). Brain mechanisms for directed attention. J. R. Soc. Med. 71: 14-28.
  • Mukherjee, P. and E. Kaplan (1995). Dynamics of neurons in the cat lateral geniculate nucleus: in vivo electrophysiology and computational modeling. J. Neurophysiol. 74: 1222-1243.
  • Nicolelis, M. A. L. (1995). Computing with thalamocortical ensembles during different behavioural states. J. Physiol. 566(1): 37–47.
  • O'Kusky, J. and M. Colonnier (1982). A laminar analysis of the number of neurons, glia, and synapses in the adult cortex (area 17) of adult macaque monkeys. J. Comp. Neurol. 210: 278-290.
  • Olshausen, B. A. and D. J Field (2004). Sparse coding of sensory inputs. Current Opinion in Neurobiology 14(4): 481-487.
  • Olshausen, B. A. and D. J Field (2005). How Close Are We to Understanding V1? Neural Computation 17, 1665–1699.
  • Parent, A. and L.-N. Hazrati (1995). Functional anatomy of the basal ganglia. I. The cortico-basal ganglia-thalamo-cortical loop. Brain Res. Rev. 20(1): 91-127.
  • Pesaran, B., J. S. Pezaris, M. Sahani, P. P. Mitra, and R. A. Andersen (2002). Temporal structure in neuronal activity during working memory in macaque parietal cortex. Nat Neurosci 5: 805-811.
  • Peters, A. and B. Payne B (1993). Numerical relationships between geniculocortical afferents and pyramidal cell modules in cat primary visual cortex. Cereb Cortex 3: 69-78.
  • Peters, A., B. Payne, and J. Budd (1994). A numerical analysis of the geniculocortical input to striate cortex in the monkey. Cerebral Cortex 4: 215-229.
  • Reinagel, P., D. Godwin, S. M. Sherman, and C. Koch (1999). Encoding of visual information by LGN bursts. J. Neurophysiol. 81: 2558-2569.
  • Rioult, M., D. Friedman, and J. Donoghue (2000). Learning-induced LTP in neocortex. Science 290: 533-536.
  • Rockel, A. J., R. W. Hiorns, and T. P. Powell (1980). The basic uniformity in structure of the neocortex. Brain 103:221-244.
  • Rozov, A., J. Jerecic, B. Sakmann, and N. Burnashev (2001). AMPA receptor channels with long-lasting desensitization in bipolar interneurons contribute to synaptic depression in a novel feedback circuit in layer 2/3 of rat neocortex. J Neurosci 21: 8062-8071.
  • Rodriguez, A., J. Whitson, and R. Granger (2004). Derivation and Analysis of Basic Computational Operations of Thalamocortical Circuits. Journal of Cognitive Neuroscience 16: 856-877.
  • Rouiller, E. and Welker (1991). Morphology of corticothalamic terminals arising from auditory cortex of rat: a Phaseolus vulgaris-leucoagglutinin (PHA-L) tracing study. Hear. Res. 56: 179-190.
  • Rulkov, N., I. Timofeev, and M. Bazhenov (2004). Oscillations in large-scale cortical networks: map-based model. Journal of Computational Neuroscience 17(2): 203-23.
  • Rumelhart, D. and D. Zipser (1985) Feature discovery by competitive learning. Cognitive Sci 9:75-112.
  • Sanchez-Vives, M. V. and D. A. McCormick (2000). Cellular and network mechanisms of rhythmic recurrent activity in neocortex. Nat. Neurosci. 3: 1027-1034.
  • Sarter, M. and J. P. Brun (2000). Cortical cholinergic inputs mediating arousal, attentional processing and dreaming: differential afferent regulation of the basal forebrain by telencephalic and brainstem afferents. Neuroscience 95: 933-952.
  • Schultz, W. (1997). A Neural Substrate of Prediction and Reward. Science 275: 1593-1599.
  • Schultz, W. (1998). Predictive Reward Signal of Dopamine Neurons. The Journal of Neurophysiology 80(1): 1-27.
  • Seki, K., M. Kudoh, and K. Shibuki (2001). Sequence dependence of post-tetanic potentiation after sequential heterosynaptic stimulation in the rat auditory cortex. J. Physiol. 533: 503-518.
  • Sherman, S. (2001). Tonic and burst firing: dual modes of thalamocortical relay. Tr. Neurosci. 24: 122-6.
  • Sherman, S. M. and R. W. Guillery (2002). The role of the thalamus in the flow of information to the cortex. Philosophical Transactions of the Royal Society B 1428(357): 1695-1708.
  • Sherman, S. M. and R. W. Guillery (2006). Exploring the Thalamus and its Role in Cortical Function. Cambridge, MA: MIT Press.
  • Shimono, K., F. Brucher, R. Granger, G. Lynch, and M. Taketani (2000). Origins and distribution of cholinergically induced beta rhythms in hippocampal slices. Journal of Neuroscience 20: 8462-8473.
  • Silberberg, G., A. Gupta, and H. Markram (2002). Stereotypy in neocortical microcircuits. Trends Neurosci 25: 227-230.
  • Simoncelli, E. P. and B. A. Olshausen (2001). Natural Image Statistics and Neural Representation. Annu. Rev. Neurosci. 24: 1193–216.
  • Simons, D. J. and G. E. Carvell (1989). Thalamocortical response transformation in the rat vibrissa/barrel system. Journal of Neurophysiology, 61(2): 311-330.
  • Singer, W. (1998). Consciousness and structure of neuronal representations. Philos. Trans. R. Soc. Lond. B Biol. Sci. 353: 1829 –1840.
  • Soto, G., N. Kopell, and K. Sen (2006). Network Architecture, Receptive Fields, and Neuromodulation: Computational and Functional Implications of Cholinergic Modulation in Primary Auditory Cortex. J. Neurophysiol. 96: 2972–2983.
  • Steriade, M., and Deschénes (1984). The thalamus as a neuronal oscillator. Brain Research Reviews 8: 1-63.
  • Steriade, M., M. Deschenes, L. Domich, and C. Mulle (1985) Abolition of spindle oscillations in thalamic neurons disconnected from nucleus reticularis thalami. J. Neurophysiol. 54: 1473-1497.
  • Steriade, M. and I. Timofeev (2003). Neuronal plasticity in thalamocortical networks during sleep and waking oscillations. Neuron 37: 563-576.
  • Steriade, M. and R. R. Llinas (1988). The functional states of the thalamus and the associated neuronal interplay. Physiol. Rev. 68: 649-742.
  • Steriade, M., S. Datta, D. Pare, G. Oakson, and R. Curro (1990). Neuronal activities in brainstem cholinergic nuclei related to tonic activation processes in thalamocortical systems. J. Neurosci. 10: 2541-2559.
  • Steriade, M. and D. Contreras (1995). Relations between cortical and thalamic cellular events during transition from sleep patterns to paroxysmal activity. J. Neurosci. 15: 623-642.
  • Steriade, M. (1997). Synchronized activities of coupled oscillators in the cerebral cortex and thalamus at different levels of vigilance. Cereb Cortex 7: 583-604.
  • Steriade, M., A. Nuñez, and F. Amzica (1993). Intracellular analysis of relations between the slow (<1 Hz) neocortical oscillations and other sleep rhythms of electroencephalogram. J. Neurosci. 13: 3266-3283.
  • Stickgold, R., L. James, and J. A. Hobson (2000). Visual discrimination learning requires sleep after training. Nat. Neurosci. 3: 1237-1238.
  • Suri, R. E., and W. Schultz (1998). Learning of sequential movements by neural network model with dopamine-like reinforcement signal. Experimental Brain Research 121: 350 – 354.
  • Suri, R. E., and W. Schultz (1999). A neural network model with dopamine-like reinforcement signal that learns a spatial delayed response task. Neuroscience 91: 871 – 890.
  • Sutton, R. S. and A. G. Barto (1998). Reinforcement Learning. Cambridge, MA: MIT Press.
  • Swanson, L. (2000). Cerebral hemisphere regulation of motivated behavior. Brain Research 886(2000): 113–164.
  • Symonds, L. L., A. C. Rosenquist, S. B. Edwards, and L. A. Palmer (1981). Projections of the pulvinar-lateral posterior complex to visual cortical areas in the cat. Neuroscience 6(10): 1995-2020.
  • Szentagothai, J. (1975). The 'module-concept' in cerebral cortex architecture. Brain Res. 95: 475-496.
  • Timofeev, I., F. Grenier, M. Bazhenov, T. J. Sejnowski, and M. Steriade (2000). Origin of slow cortical oscillations in deafferented cortical slabs. Cer. Cortex 10: 1185-1199.
  • Timofeev, I. and M Bazhenov (2005). Mechanisms and Biological Role of Thalamocortical Oscillations. In Trends in Chronobiology Research. (F.Columbus, Ed.) Hauppauge, NY: Nova Science Publishers, pp.1-47.
  • Traub, R. D., D. Contreras, M. O. Cunningham, H. Murray, F. E. N. LeBeau, A. Roopun, A. Bibbig, W. B. Wilent, M. J. Higley, and M A. Whittington (2005). Single-Column Thalamocortical Network Model Exhibiting Gamma Oscillations, Sleep Spindles, and Epileptogenic Bursts. Neurophysiol. 93: 2194-2232.
  • Troyer, T. W., A. E. Krukowski, N. J. Priebe, and K. D. Miller (1998). Contrast-Invariant Orientation Tuning in Cat Visual Cortex: Thalamocortical Input Tuning and Correlation-Based Intracortical Connectivity. Journal of Neuroscience, 18(15): 5908-5927.
  • Valverde, F. (2002). Structure of the cerebral cortex. Intrinsic Organization and comparative analysis of the neocortex. Rev. Neurol. 34: 758-780.
  • von Bartheld, C. (1999). Systematic bias in an "unbiased" neuron counting technique. Anat. Rec. 257: 119-120.
  • von Bartheld, C. (2001) Comparison of 2-D and 3-D counting: the need for calibration and common sense. Trends. Neurosci. 24: 504-506.
  • von der Malsburg, C. (1973). Self-organization of orientation sensitive cells in the striate cortex. Kybernetik 14: 85-100.
  • von Krosigk, M., T. Bal, and D. A. McCormick (1993). Cellular mechanisms of a synchronized oscillation in the thalamus. Science 261: 361-364
  • Webber, R. M. and G. B. Stanley (2004). Nonlinear Encoding of Tactile Patterns in the Barrel Cortex. J. Neurophysiol. 91: 2010 –2022.
  • White, E. and A. Peters (1993). Cortical modules in the posteromedial barrel subfield (Sml) of the mouse. J Comp Neurol 334: 86-96.
  • Wong, H., X. B. Liu, M. Matos, S. Chan, I. Perez-Otano, M. Boysen, J. Cui, N. Nakanishi, J. Trimmer, E. Jones, S. A. Lipton, and N. Sucher (2002). Temporal and regional expression of NMDA receptor subunit NR3A in the mammalian brain. J. Comp. Neurol. 450: 303-317.
  • Zhang, S., J. Huguenard, and D. Prince (1997). GABAa receptor mediated Cl-currents in rat thalamic reticular and relay neurons. J. Neurophysiology 78: 2280-2286.
  • Zhu, J., and B. Connors (1999). Intrinsic firing patterns and whisker-evoked synaptic responses of neurons in the rat barrel cortex. J. Neurophysiol. 81: 1171-1183.
  • Zikopoulos, B. and H. Barbas (2006). Prefrontal Projections to the Thalamic Reticular Nucleus form a Unique Circuit for Attentional Mechanisms. The Journal of Neuroscience 26(28): 7348-7361.

Internal references

See Also

Basal Ganglia, Brainstem, Cortex, Models of Visual Cortex, Neuroanatomy,Thalamocortical Circuit, Thalamocortical Oscillations, Thalamus, Visual Cortex

Personal tools
Namespaces

Variants
Actions
Navigation
Focal areas
Activity
Tools