Next we checked whether the suppression occurs at the end of the

Next we checked whether the suppression occurs at the end of the cascade, at the level of AMPA receptor trafficking in and out of the synapse. To that end we exploited the facts that LTP and LTD can be both reversed by activity. The

reversal of LTP (termed de-potentiation) and LTD (termed de-depression) share common downstream mechanism of expression with LTD and LTP, as they involve changes in AMPA receptor function; yet they differ in induction mechanisms, as they involve different kinase and phosphatase pathways (Hardingham et al., 2008 and Lee and Huganir, 2008). We reasoned that if the GPCR-mediated suppression occurs at the expression level (AMPAR trafficking), de-potentiation and de-depression should also be affected. The experiments were carried out in a two independent inputs setting, to allow internal controls, and using pairing PD-1/PD-L1 inhibitor conditioning (to 0mV or –40mV) to induce LTP and LTD as well as to reverse them (Figure 5). First LTD was induced in both inputs, and 20 min later one input was de-depressed by pairing with 0mV while the other input was not stimulated. The second pairing effectively reversed LTD in either control Wnt inhibitor conditions (de-depressed versus nonstimulated; paired t test: p = 0.0086) (Figure 5A), and in the presence of methoxamine (paired

t test: p = 0.0368. Figure 5B), indicating that α1 adrenergic receptors do not suppress de-depression. A similar strategy was used to test the role of β-adrenergic receptors on de-potentiation: LTP induction in both pathways, followed by pairing with –40mV in one input (Figures 5E and 5F). The second pairing reversed LTP either in control conditions (p = 0.0343. Figure 5E) or in the presence

of isoproterenol (p = 0.0007) (Figure 5F). Next we compared the effects of methoxamine Unoprostone on LTD and de-potentiation simultaneously by first inducing LTD in one input and then applying the 0mV pairing to both inputs. In control experiments (Figure 5C) the second pairing potentiated both the depressed input (p = 0.0008) and the naive (p = 0.0038); in the presence of methoxamine (Figure 5D) the depressed inputs potentiated (p = 0.0236), but not the naive inputs (p = 0.2054), confirming that α1-adrenergic receptors prevent LTP but they do not affect de-potentiation. The effects of β-adrenergic receptors on LTD and depotentiation were compared with a similar strategy: first LTP induction of one input, followed by simultaneous pairing with −40mV of both potentiated and naive inputs. Under normal conditions both inputs became depressed (potentiated inputs: p = 0.001; naive inputs: p = 0.0006) (Figure 5G). In contrast, in the presence of isoproterenol only the previously potentiated input became depressed (potentiated inputs: p = 0.048; naive inputs: p = 0.604) (Figure 5H). These results confirmed that β-adrenergic receptors prevent LTD but do not affect de-depression.

, 2007) Critically, however, the human SFEBq cultures were not r

, 2007). Critically, however, the human SFEBq cultures were not reported to produce any late neurons with markers of upper cortical layers, despite some being cultured for as long as 106 days (Eiraku et al., 2008). More recently, similar results with hESCs and hiPSCs were obtained through a simpler embryoid body (EB)-based method, with a high efficiency of dorsal telencephalic specification (Li et al., 2009 and Zeng et al., 2010). EBs were cultured without growth factors for 2 weeks until

neural CB-839 rosettes formed. Gene expression analysis showed that certain Wnt morphogens (dorsalizing signals) were strongly induced during the second week, and nearly all the neural rosette cells were Foxg1+/Pax6+ by the third week. The cells exhibited the same responsiveness to dorsoventral patterning cues (Wnt versus Sonic hedgehog [SHH]) that Sasai’s group originally described (Watanabe et al., 2005). The progenitor cells generated Tbr1+ and Ctip2+ glutamatergic neurons but again, the production of late cortical neurons with markers typical of upper layers was not reported. A remarkably simple protocol for producing cortical neurons from mESCs was reported

by Vanderhaeghen’s group (Gaspard et al., 2008). In this method, mESCs were plated at low density in default differentiation medium. The cells naturally adopted a telencephalic identity, but in contrast to aggregate cultures, a majority of telencephalic cells expressed ventral progenitor cell markers within 2 weeks and differentiated

into GABAergic neurons. Noting that SHH expression was induced during the period of neural conversion, the authors treated the cells with Gefitinib mw a SHH antagonist, resulting in nearly complete suppression of ventral markers and yielding glutamatergic neurons with pyramidal morphology, indicating a dorsal fate shift. These cells also exhibited the known sequence of neuronal subtype production, with Reelin+ and Tbr1+ neurogenesis Dichloromethane dehalogenase peaking first, followed by production of Ctip2+ and then Cux1+ and Satb2+ neurons. However, the authors also noted a large underrepresentation of Cux1+ and Satb2+ neurons when they analyzed the expected proportions of each subtype, suggesting that in vivo cues are important for the full generation of late neurons destined for upper cortical layers. Surprisingly, the cortical cells derived by Gaspard et al. (2008) displayed specific areal identity upon transplantation into the frontal cortex of neonatal mice, extending axonal projections to a repertoire of subcortical targets that would be expected from neurons in the visual/occipital cortex. Prior to grafting, most of the mESC-derived neurons expressed Coup-TF1, which is expressed in the caudal but not rostral cortex. This suggested that the cells have an innate differentiation program that requires neither intracortical (e.g., FGF, Wnt, BMP gradients) nor extracortical (e.g., thalamocortical afferents) patterning cues to acquire area-specific neuronal properties.

We first determined whether there was a disruption in the develop

We first determined whether there was a disruption in the developmental switch from NR2B to NR2A in layer 2/3. We made whole-cell patch-clamp

recordings from layer 2/3 pyramidal neurons in slices of primary visual cortex and found that NMDA EPSCs elicited by layer 4 stimulation exhibited longer decay times and greater ifenprodil sensitivity in mGluR5 knockouts compared to wild-type (Figures 6E–6H). This indicates a deficiency in the development switch from NR2B to NR2A-containing receptors. Visual experience in dark-reared rodents causes a rapid switch from NR2B- to NR2A-containing NMDARs at layer 4 inputs onto layer 2/3 pyramidal neurons in primary visual cortex that depends upon NMDAR activation (Philpot Selleckchem Hydroxychloroquine et al., 2001 and Quinlan et al., 1999). Therefore, we next tested whether this experience-dependent plasticity is disrupted in mGluR5 knockout mice. We dark reared wild-type mice and mGluR5 knockout littermates from P6 until P17–P19, exposed some of these animals to 2.5 hr of light, and then investigated

the effects on NMDA EPSCs at layer 4 inputs onto layer 2/3 pyramidal cells. In wild-type mice NMDA EPSCs in animals exposed to light (+LE) exhibited faster kinetics and reduced ifenprodil sensitivity compared to mice that did not receive light exposure (Figures 7A–7E). The degree GSK1349572 research buy of change in these parameters was very similar to that previously reported (Philpot et al., 2001 and Quinlan et al., 1999) and confirms that 2-C-methyl-D-erythritol 2,4-cyclodiphosphate synthase even brief exposure to light can drive the switch from NR2B to NR2A in visual cortex. In mGluR5 knockout

mice light exposure failed to produce any significant change in NMDA EPSC kinetics or ifenprodil sensitivity (Figures 7A–7E). It was also noticeable that the dark-reared wild-type and knockout mice (that were not exposed to light) exhibited very similar NMDA EPSC kinetics and ifenprodil sensitivity, indicating that visual experience and mGluR5 are necessary for the developmental change from NR2B to NR2A-containing NMDARs in visual cortex during the first few postnatal weeks. During the first postnatal week, most cortical synapses express NR2B-containing receptors, whereas later in development (>P14), many of these receptors are replaced with NR2A-containing NMDARs. Synaptic activity is involved in regulating this switch, and a role for sensory experience in primary sensory cortex has also been demonstrated; however, the molecular mechanisms driving this ubiquitous NMDAR subtype switch have hitherto been largely unexplored. Here, we find that activation of both mGluR5 and NMDARs is required for this switch to occur at synapses on hippocampal CA1 pyramidal neurons. Furthermore, we define a downstream signaling pathway involving PLC activation, release of Ca2+ from IP3R-dependent stores, and activation of PKC (see Figure 8 for model).

Here, we provide a direct link between retrieval-mediated encodin

Here, we provide a direct link between retrieval-mediated encoding processes and flexible memory expression in the human brain. Using multivoxel pattern analysis, we demonstrate that prior experience is reactivated during encoding of related events, and that such online reactivation of memories is predictive of individuals’ ability to infer novel relationships between two discrete, overlapping episodes. We further show that reactivation within content-sensitive higher-order visual areas is related to anterior MTL cortex activation, suggesting that responses in this region may influence the extent and specificity of retrieved memories. Extensive controversy exists as to whether encoding activation relating to

subsequent inference reflects ZD6474 order memory integration during encoding or strengthening of individual learn more directly learned associations, leading to improved “on-the-fly” inference at retrieval (for a review, see Zeithamova et al., 2012). Here, decreasing hippocampal and increasing VMPFC engagement across repetitions of overlapping events were associated with superior inference even when controlling for memory of premise associations, providing a strong evidence for online integration of related memories as they are encoded. Furthermore, we observed increased connectivity between hippocampus and VMPFC across

interleaved presentations of overlapping events. These findings illustrate how a functionally coupled hippocampal-VMPFC circuit supports binding of reactivated memories with current experience, forming integrated memories that relate overlapping experiences. These relational memory networks enable the predictive application of memory by grouping related elements from multiple experiences in support of future inferential judgments. The present study organically builds upon and significantly extends prior studies examining the neural mechanisms supporting retrieval-mediated

learning. Prior rodent RANTES research has shown that the existence of a well-learned spatial schema speeds acquisition of new object-place associations (Tse et al., 2007 and Tse et al., 2011). Another recent report demonstrated that blocking hippocampal plasticity during contextual fear conditioning prevents the transfer of a newly acquired fear response to a previously experienced, overlapping spatial context (Iordanova et al., 2011). The presumption in each of these studies is that existing memories are reactivated during new learning and updated with new information, resulting in facilitated encoding and generalization. However, without an empirical measure of memory reactivation, such a presumption is only speculative. The methods employed in the current study enabled us to directly observe memory reactivation during encoding of related events, providing a key index of a process critical to retrieval-mediated learning. Furthermore, in contrast to prior studies, our findings emphasize the beneficial function of retrieval-mediated learning.

Other considerations are worth noting,

however Modeling

Other considerations are worth noting,

however. Modeling exploration is not trivial, because it requires predicting that participants make a response that counters their general propensity to exploit the option with highest value, and therefore any model of exploration requires knowing when this will occur. Because exploited options are sampled more often, their outcome uncertainties are generally lower than those of the alternative options. Thus, when the subject exploits, they are selecting the least uncertain option, making it more difficult to estimate the positive influence of uncertainty on exploration. As noted above, this problem is exacerbated by “sticky choice,” whereby participants’ choices in a given trial are often autocorrelated with those of previous trials independent of value. Finally, studies failing to report an effect of uncertainty on exploration

Osimertinib cell line have all used n-armed bandit tasks with dynamic reward contingencies across trials (Daw et al., 2006, Jepma et al., 2010 and Payzan-LeNestour and Bossaerts, 2011), and participants responded as if only the very last trial was informative about value (Daw et al., 2006 and Jepma et al., 2010). It may be more difficult to estimate uncertainty-driven exploration in this context, given that participants would be similarly uncertain about all alternative options that had not been selected in the most recent trial. In our behavioral paradigms and model fits, we have attempted to confront these issues allowing us to estimate uncertainty, its effects on exploration, and the neural correlates ZD6474 of this

relationship. First, it is helpful to note the ways that the current paradigm is atypical in comparison to more traditional n-armed bandit tasks. Initially, the task was designed not to study exploration, but rather as a means of studying incremental learning in Parkinson’s patients and Ibrutinib as a function of dopamine manipulation (Moustafa et al., 2008). However, in the Frank et al. (2009) large-sample genetics study, it was observed that trial-by-trial RT swings appeared to occur strategically and attempts to model these swings found that they were correlated with relative uncertainty. Importantly, this is not just a recapitulation of the finding that the model fits better when relative uncertainty is incorporated (i.e., ε is nonzero); much of this improvement in fit was accounted for by directional changes in RT from one trial to the next (RT swings). This distinction is important: in principle a fitted nonzero ε could capture an overall tendency to respond to an action that is more or less certain, e.g., if a subject exploits most of the time, ε would be negative (assuming the exploitation part of the model is imperfect in capturing all exploitative choices).

Performance and payout were only related to how close subjects’ b

Performance and payout were only related to how close subjects’ behavior matched the normative optimal solution (thereby incentivizing an accurate correlation representation) but was independent of the actual amount or variance of

the produced energy mix. Importantly, during the experiment subjects never received direct feedback on their performance at minimizing energy fluctuations (i.e., only saw trial-by-trial outcomes) and the bonus and optimal weights were only revealed after the experiment. We omitted feedback during the task to prevent subjects www.selleckchem.com/products/ch5424802.html from using a strategy that is based on optimizing the performance feedback instead of learning the correlation of the individual outcomes. Although the portfolio value is shown on every trial, and the deviance of this value from its mean gives some hints to performance, this is only a crude measure of whether the current weights are good because even with optimal weights the amount of portfolio fluctuation depends on the current correlation. Because the optimal mixing weights (portfolio weights) in our task depend on individual variance from solar and wind power plants and their correlation strength, the best strategy is to learn the variances and correlations by observation of individual outcomes and then translate these estimates into an optimal ABT-888 molecular weight resource allocation (i.e., weightings). Although subjects

could learn the statistical properties underlying outcome generation by observation, the outcomes of individual trials were unpredictable. Their task was then to continuously mix the two resources into an energy portfolio and thereby minimize the fluctuation of the portfolio value from trial to trial. Both resources fluctuated around a common mean, with outcomes drawn from a rectangular distribution with a specific variance. In our task the standard deviation of one resource was always twice that of the other because this maximized the influence of the correlation on the portfolio weights (see Figure S1 for details). The sequence of correlated random numbers for the two resources

were generated by the Cholesky decomposition method (Gentle, 1998). This was realized by first drawing random numbers xA and xB for resources A, B from a rectangular distribution. Temozolomide The outcome of the second resource xB was then modified as xB = xA∗ r + xB∗ sqrt(1 − r2), whereby r is the generative correlation coefficient. Finally, xA and xB were normalized to their desired standard deviations (in the three blocks: 20/10, 15/30, 10/20) and common means (30, 50, 40). We chose a rectangular distribution to increase the sensitivity of our fMRI experiment in finding neural correlates of covariance and covariance prediction errors as the linear regression against BOLD activity is most sensitive if the values of the parametric modulators are distributed along their entire range. This is not true for normal distributed outcomes, which have proportionally the largest amounts of data close to the mean.

Combined spatial analysis of genetically and functionally defined

Combined spatial analysis of genetically and functionally defined

interneuron populations with an assessment of quantitative contributions to the synaptic regulation of different motor neuron pools will provide answers to these questions. As has become apparent, spinal interneurons cannot be considered to be simply a limited group of local neurons shaping and modulating motor circuit function in recurrent modules. Spinal interneuron diversification is evident at the developmental level by progenitor domain origin, time of neurogenesis, migratory path, and acquisition of distinct transcriptional profiles. These early features translate to diversification in the mature spinal cord, in which neuronal subpopulations exhibit differential spatial distribution patterns, neurotransmitter http://www.selleckchem.com/products/ly2157299.html profiles, GDC0199 connectivity matrices including

synaptic in- and output, and functional properties (Figure 5C). Although interneuron populations are often loosely categorized along a single dimension (e.g., transcriptional, neurotransmitter, or spatial profile), these same interneurons may in fact be functionally multifaceted (Edgley, 2001 and Jankowska, 2008), which complicates classification criteria. Analysis of connectivity profiles provides ample evidence that many spinal interneurons establish connections over many segments, and individual motor neuron pools receive direct input from segmentally widely distributed interneuron populations (Stepien et al., 2010 and Tripodi et al., 2011). Consequently, many spinal “interneurons” exhibit properties analogous to long-distance projection neurons not unlike pyramidal neurons in the cerebral cortex and therefore cannot be strictly considered to function as local interneurons. Neurons in the spinal cord of this category exhibit fundamentally different connectivity profiles and functions, including excitatory and inhibitory subtypes. On the other end of Lacidipine the spectrum, Renshaw cells or spinal interneuron populations in the substantia gelatinosa (Brown, 1981 and Todd, 2010) can be considered more similar to locally projecting cortical interneurons

such as fast-spiking Parvalbumin interneurons (Isaacson and Scanziani, 2011), both contributing exclusively to local circuit computations. In the cortex, one defining arbiter for the use of the term “interneuron” is based on the fact that these neurons migrate into the cortex from distant sites (i.e., ganglionic eminence) and many of them project locally (Fishell and Rudy, 2011, Gelman and Marín, 2010 and Klausberger and Somogyi, 2008). In contrast, spinal neurons are generated locally, eliminating this distinguishing parameter. For future reference, it will be important to consider that the commonly used terminology “spinal interneurons” embraces a bewildering array of functionally distinct neuronal subtypes in sum charged with local as well as long-distance computations in the spinal cord (Figure 5C).

Stem cells thus undergo both asymmetric and symmetric divisions w

Stem cells thus undergo both asymmetric and symmetric divisions within their niches, depending on tissue this website and developmental

context (reviewed in Morrison and Kimble, 2006). Mammalian tissues also have specialized niches that secrete short-range factors that promote stem cell maintenance (Morrison and Spradling, 2008). As in the niches characterized in Drosophila and C. elegans, Notch ligands, BMPs, and Wnt proteins have been implicated in the regulation of stem cell maintenance in multiple mammalian tissues, including in the CNS ( Doetsch, 2003) and in hair follicles ( Blanpain and Fuchs, 2006). These factors are presumed to be locally secreted by supporting cells that create the niches, though the identities of these supporting cells are not yet well characterized in most mammalian tissues. Stem cells are also extrinsically regulated by long-range signals, including an evolutionarily conserved role for insulin pathway regulation (Figure 1C).

Circulating insulin-like peptide is required for the maintenance of Drosophila germline stem cells and intestinal stem cells, and quantitative changes in nutritional status lead to changes in stem cell function as a result of changing insulin-like peptide levels ( LaFever and Drummond-Barbosa, 2005 and McLeod et al., 2010). Mammalian stem cells are also positively regulated by insulin signaling as fetal forebrain stem cells adjacent to the lateral ventricle are regulated by IGF2 in cerebral spinal fluid ( Lehtinen et al., 2011). Selleckchem RAD001 Nonetheless, additional work will be required to determine whether mammalian stem cells are regulated by systemic nutritional status. Aging is associated with reduced regenerative capacity and stem cell function in multiple tissues, including the CNS (Figure 2C) (Maslov et al., 2004). Stem cell function Dipeptidyl peptidase decreases with age in many tissues in an evolutionarily conserved manner. Fly spermatogonial

stem cell function declines during aging as a consequence of both cell-intrinsic (Cheng et al., 2008) and niche changes (Boyle et al., 2007). In aging mammalian tissues, stem cells exhibit reduced self-renewal potential and accumulation of damage to DNA, mitochondria, and other macromolecules (Rossi et al., 2008 and Sharpless and DePinho, 2007). The declines in stem cell function during aging are also associated with increasing tumor suppressor expression (Figure 2B). The p16Ink4a cyclin-dependent kinase inhibitor, a negative regulator of cell-cycle progression that sometimes causes cellular senescence, is generally not detectable in young adult tissues, but expression increases during aging (Krishnamurthy et al., 2004). This increase in p16Ink4a expression contributes to the age-related decline in stem cell function in the hematopoietic and nervous systems, as well as the decline in β cell proliferation in the pancreas. Deficiency for p16Ink4a partially rescues the age-related declines in stem cell frequency, mitotic activity, and neurogenesis in the forebrain ( Molofsky et al.

Single-fluorophore blinking events were detected at the end of th

Single-fluorophore blinking events were detected at the end of the movie (typically GDC-0199 molecular weight in frames 5,000–10,000), and their mean intensity, I, was measured for each cluster. The total fluorophore number, N, of the cluster was then calculated according to the formula: N = A / (I × τw). For dual-color quantification, decay recordings were acquired first for mRFP followed by Dendra2, since excitation at 561 nm did not affect the nonconverted form of Dendra2. The calculated fluorophore numbers of individual gephyrin clusters (from the pulsed photoconversion or the fluorescence

decay method) were equated to the fluorescence intensity of the same clusters in images taken with the mercury lamp (background-corrected integrated cluster intensity). This resulted in a conversion factor ϕ (fluorescence intensity/molecule) that could be applied to any structure

visualized in conventional fluorescence images, provided that identical imaging conditions were maintained. The authors thank Alain Bessis, Yasmine Cantaut-Belarif, and Andréa Dumoulin (Institut de Biologie de l’Ecole Normale Supérieure) as well as Christophe Zimmer and Mickaël Lelek (Institut Pasteur) for technical help. This project was funded by the Fondation Pierre-Gilles de Gennes through a research contract with Nikon France, the Institut pour la Recherche sur la Moelle Épinière et l’Encéphale, and by grants TRIDIMIC buy Dabrafenib and MorphoSynDiff from the Agence Nationale pour la Recherche. C.G.S. acknowledges grant Lamonica, and I.I. acknowledges the Netherlands Organisation for Scientific Research for financial support. P.C.R. was supported by a Marie Curie International

Incoming Fellowship within the 7th European Community Framework Programme. C.G.S., I.I., M.D., and A.T. designed the experiments; C.G.S., isometheptene I.I., P.C.R., P.R., and M.E.B. conducted the experiments and analyzed the data; C.G.S. and I.I. wrote the manuscript. “
“Spontaneous neuronal activity pervades the developing nervous system and correlations contained in its patterns guide the synaptic refinement of many immature circuits (Blankenship and Feller, 2010 and Katz and Shatz, 1996). This has best been studied in the developing visual system, where waves of spontaneous activity originate in the retina (Meister et al., 1991) and dictate firing patterns up to primary visual cortex (V1) (Ackman et al., 2012 and Mooney et al., 1996). Across many species, retinal waves mature in three stereotypic stages (I–III) (Blankenship and Feller, 2010 and Wong, 1999). In each stage, distinct mechanisms give rise to unique activity patterns that serve specific functions in organizing visual circuits. During stage III (postnatal day 10–14, P10–P14 mice), the firing patterns of different RGC types diverge (Lee et al., 2002, Liets et al., 2003 and Wong and Oakley, 1996).

The occurrence of read-through transcripts from exon 1a through t

The occurrence of read-through transcripts from exon 1a through the constant region of TRIP8b was verified through RT-PCR (data not shown). Lentivirus was stereotactically injected into the dorsal hippocampi of 5-week-old mice. Transverse hippocampal slices (400 μm) were prepared two http://www.selleckchem.com/products/Bortezomib.html weeks after viral injection. Animals were sacrificed in accordance to institutional IACUC standards. For solution composition and detailed methods see Supplemental Experimental Procedures. Virally infected neurons

were identified by EGFP fluorescence. Following recordings of infected neurons, slices were fixed for 30 min in PFA and imaged to ensure that all dye-filled cells were also EGFP+ and that the dendritic structure of infected cells was normal. Series resistance was less than 15 MΩ; capacitance and series resistance were monitored and compensated throughout

the experiments. Recordings were performed at 32°C. All data was aquired with Pclamp software (Molecular Devices) and analyzed with IGORPro Selleckchem SCH727965 (wavemetrics). This work was supported by grants NS36658 and MH80745 from NIH, by fellowships from NIH F32NS064732 (R.P.), the Italian Academy for Advanced Studies in America (B.S.) and a Research Grant from the Epilepsy Foundation (B.S.). We gratefully acknowledge Terunaga Nakagawa and Morgan Sheng for providing the pLLhS and Pavel Osten for providing the pFCK(0.4)GW lentiviral vectors, and Frank Müller for generously providing the rat 7C3 monoclonal HCN1 antibody. We thank Ming-Kuei Jang and Thomas Yocum for preliminary experiments, Haiying Liu for technical assistance and Vivien Chevaleyre for experimental advice. “
“The majority of synaptic inputs onto neurons in the neocortex originate from nearby neurons within the same cortical area, producing local microcircuits that are ubiquitous in neocortex (Braitenberg and Schüz, 1998, Douglas and Martin, 2004 and White, 2007). Local excitatory

connections provide the major excitatory input to neocortical principal neurons, are highly recurrent, and are critically important for information processing, particularly in sensory neocortex (Douglas et al., 1995, Buonomano and Maass, 2009 and Rigas and Castro-Alamancos, 2009). Despite the ubiquitous PDK4 nature of local excitatory circuits, very little is known about their organization and almost nothing known about development at the level of connections between individual neurons, thus leading to a poor understanding of mature network architecture. Layer 4 of the rodent barrel cortex is the primary input layer for ascending sensory information arriving via thalamocortical fibers (Petersen, 2003). Layer 4 contains clusters of neurons, named barrels, each of which receives topographically mapped input from a corresponding whisker; this provides an anatomical correlate for whisker receptive fields.