), prepatent period, site of colonization, amongst other aspects

), prepatent period, site of colonization, amongst other aspects ( Coudert et al., 1995 and Pakandl, 2009). Even though oocyst morphology allows a good differentiation of rabbit Eimeria, it is practically impossible to carefully evaluate thousands of oocysts just to assess if a given strain is really pure. Moreover, a small portion of oocysts may differ from the typically expected morphology, posing a challenge Selleckchem Hydroxychloroquine to a correct diagnosis.

In addition, all these biological features may present a variable level of overlap, hampering in some cases an accurate Eimeria species identification ( Long and Joyner, 1984). In the case of chicken Eimeria, various molecular differentiation assays have been proposed, using distinct targets such as ITS1 ( Schnitzler et al., 1998 and Schnitzler et al., 1999), ITS2 ( Woods et al., 2000 and Gasser et al., 2001), and SCARs ( Fernandez et al., 2003a and Fernandez et al., 2003b). For a comprehensive review www.selleckchem.com/products/byl719.html on the molecular diagnosis of avian coccidiosis, the reader is referred to Morris and Gasser (2006). Finally, real-time PCR assays have also been described ( Blake et al., 2008 and Morgan et al., 2009), permitting quantitative studies of the distinct Eimeria species.

Conversely, no PCR-based method has been developed for the unequivocal differentiation of the 11 Eimeria species that infect rabbits. Ceré et al. (1995) described a RAPD assay for the differentiation of nine Eimeria species, but this method presents

low reproducibility. Also, the typical multiband character makes it unsuitable for testing mixed field samples. Ceré et al. (1996) developed a specific and sensitive PCR assay for E. media, but this approach has not been extended to the remaining species. The precise diagnosis of all species of rabbit coccidia is very important for several reasons. First, individual Eimeria species present dramatic differences in their pathogenicity ( Coudert et al., 1995). Second, an accurate Eimeria species differentiation method would represent an essential tool for monitoring the purity of strains in the production of a future multivalent vaccine. Finally, a correct species identification would allow research laboratories to assess however the purity of the strains under study. In this work, we report the development of molecular diagnostic assays for the detection and discrimination of the 11 Eimeria species that infect the domestic rabbit. Isolates of the 11 Eimeria species that infect the domestic rabbit (Oryctolagus cuniculus f. domesticus) were used throughout the work. E. coecicola, E. flavescens, E. irresidua, E. piriformis, E. stiedai and E. vejdovskyi isolates were originally obtained from field samples in the Czech Republic. E. exigua, E. intestinalis, E. magna, E. media and E. perforans were isolated in France.

Studies using high-[K+] depolarization and Ca2+ ionophores to eva

Studies using high-[K+] depolarization and Ca2+ ionophores to evaluate the effect of Ca2+-loading on the pH of presynaptic terminals isolated from brain (synaptosomes) have reported conflicting results: lack of effect (Richards et al., 1984 and Nachshen and Drapeau, 1988), acidification (Martinez-Serrano et al., 1992), or alkalinization

(Sánchez-Armass et al., 1994). To date, we know of no direct studies of pH changes in intact presynaptic terminals where Ca2+ influx is activated by physiological MDV3100 supplier action potential stimulation. We report here such measurements, made in motor nerve terminals of mice that transgenically express Yellow Fluorescent Protein (YFP) in neurons (Thy-1 promoter; Ormö et al., 1996 and Feng et al., 2000). These measurements are based on the fact that YFP fluorescence is pH sensitive over the physiological range: reversible protonation of the YFP chromophore domain decreases its fluorescence as pH acidifies from 8 to 5.5 (reviewed by Bizzarri et al., 2009). Using this pH indicator we found that the earliest effect of stimulation on the pH of motor terminals is (as expected from

studies of neuronal somata and dendrites) a Ca2+-dependent acidification whose magnitude is reduced by both the HCO3−/CO2 buffer system and an amiloride-sensitive Na+/H+ exchanger (NHE). This early acidification is followed by a pronounced, prolonged alkalinization not previously reported in neurons. We present evidence that this alkalinization is due to H+ extrusion via vesicular H+-ATPase (vATPase) transiently Epigenetics Compound Library price inserted into the plasma membrane during exocytosis. If this hypothesis is true, then the rate of decay of this alkalinization offers a method for measuring the time course with which certain vesicular components are endocytosed. We also find that inhibition of vATPase activity reduces vesicular endocytosis. This result, combined with previous reports that acidification inhibits one or more most components of clathrin-mediated endocytosis (see Discussion), suggests that the prolonged poststimulation alkalinization facilitates endocytosis. Figure 1A

shows a motor nerve terminal in the levator auris longus muscle of a mouse that transgenically expressed YFP in motor neurons. This preparation allows measurement of pH changes in motor terminal cytosol with no interference from changes that might also occur in muscle or Schwann cells. Figure 1B illustrates how YFP fluorescence in this terminal changed during and after the motor nerve was stimulated with trains of action potentials (50 Hz, 20 s). Figure 1C plots the magnitude of the average YFP fluorescence change in this terminal normalized to resting fluorescence (F/Frest), and Figure 1D shows the averaged response for 18 terminals. Changes in F/Frest were converted to cytosolic pH and [H+] assuming a resting pH of 6.

, 2007, Orban et al , 2004 and Passingham,

, 2007, Orban et al., 2004 and Passingham,

Romidepsin research buy 2009). One striking parallel in the memory signals seen across the monkey and human medial temporal lobe is significantly stronger responses to novel relative to familiar visual stimuli in the perirhinal cortex (Brown et al., 1987, Brozinsky et al., 2005, Fahy et al., 1993, Gonsalves et al., 2005, Henson et al., 2003, Köhler et al., 1998, Li et al., 1993 and Montaldi et al., 2006). Beyond this signal of relative stimulus novelty, however, the parallels between memory-related physiological signals in monkeys and humans are less striking. For example, in the monkey perirhinal cortex, Miller and Desimone (1994) reported stimulus-selective enhancement to a behaviorally relevant matching stimuli (match

enhancement) as well as stimulus-selective suppression to nonrelevant matching stimuli (match suppression) during a delayed match to sample task. Reports of match enhancement in the human perirhinal cortex, however, have been mixed (Dudukovic et al., 2011 and Duncan et al., 2009) although the tasks used in humans differed in numerous respects from the task used in monkeys. In the hippocampus, several human fMRI studies (Dudukovic et al., 2011 and Duncan et al., 2009) as well as a human single unit study in epileptic patients (Fried et al., 1997) FG 4592 reported strong match enhancement signals. By contrast, in the monkey hippocampus, several recent reports have described decrements but not enhancements in neural responses associated with repeated stimulus presentations (Jutras and Buffalo, 2010 and Yanike et al., 2009). Although many previous studies have mapped early visual areas in monkeys and humans performing the same perceptual task, few studies have compared medial

temporal lobe activity Ribonucleotide reductase across species as subjects perform the same memory task. One exception is Law et al. (2005), who developed a conditional motor associative learning task for humans based on one used in a previously published monkey physiology study (Wirth et al., 2003). Law et al. (2005) reported clear increases in the BOLD fMRI signals across the medial temporal lobe structures as human subjects learned new conditional motor associations. These findings appeared to parallel the single unit findings in the monkey hippocampus described by Wirth et al. (2003) that showed either increases or decreases in hippocampal single unit activity that were correlated with the animal’s learning curve. However, it remained unclear how the increases and decreases in single unit activity seen in individual monkey hippocampal cells corresponded to the global pattern of increased BOLD activity seen in humans.

g , Ojima et al ,

g., Ojima et al., learn more 1984) and that odors generally evoke widely distributed PCx activity (Illig and Haberly, 2003). Recent experiments have shown that different odorants activate unique subpopulations of neurons distributed across the PCx without spatial preference (Stettler and Axel, 2009) and that the projections of individual glomeruli (Nagayama et al., 2010 and Sosulski et al., 2011) and of single mitral cells (Gosh et al., 2011) are broadly

distributed in the PCx. A second set of questions concerns how the odor features represented by MOB glomeruli are recombined in the cortex. How many different mitral cell inputs synapse on an individual PCx neuron? What are the numbers and distribution of glomeruli from which the inputs to a single neuron arise? Up until recently, there has been little direct evidence concerning MOB to PCx convergence. It is known that PCx neurons can respond to dissimilar odorants that are likely to activate nonoverlapping glomeruli (Rennaker et al., 2007 and Poo and Isaacson, 2009). Some PCx neurons respond to electrical stimulation only when more than one glomerulus is coactivated (Apicella et al., 2010), and odor mixtures can activate PCx neurons that are not activated by the components alone (Stettler and Axel, 2009). Although intracortical excitation

could contribute to some of these observations, it was recently shown that individual PCx neurons indeed receive anatomical connections from multiple broadly distributed mitral cells (Miyamichi et al., 2011). Further critical INCB024360 questions relate to more detailed features

of the integrative properties of individual PCx neurons. How strong are the individual functional inputs from each glomerulus? How many glomeruli connect to each PCx cell and how many inputs must be coactive for a PCx neuron to respond? Do inputs combine linearly or nonlinearly? In this issue of Neuron, Davison and Ehlers (2011) provide important new insight into these issues by using in vivo laser scanning glutamate uncaging to create distributed artificial patterns of glomerular activation in the MOB while recording in the PCx in anesthetized mice. With this approach, the authors were able to independently activate targeted locations within Parvulin the glomerular layer of the MOB with near single glomerular spatial resolution. Davison and Ehlers’s data address several aspects of the convergence and integration of MOB inputs by individual PCx cells. First, PCx neurons did not generally respond to single-site stimulation; PCx firing was only triggered reliably by joint activation of at least three uncaging sites. Moreover, for a given number of sites, PCx activation was specific to the pattern of those sites: each cell responded differentially to different spatial patterns and different PCx cells responded differentially to a particular pattern.

While the 5% deviant responses were essentially as likely to be l

While the 5% deviant responses were essentially as likely to be larger or smaller in the Random compared

to the Periodic condition (66/138, 48%), the majority of the responses to 20% deviants were larger in the Random compared to the Periodic condition (103/156, 66%); furthermore, the average response to the 20% deviants was significantly larger in the Random than in the Periodic condition. The responses to standards followed the reverse tendencies: the differences between the responses in the Periodic and Random conditions became less prominent with increasing deviant probability (and decreasing standard probability). Thus, while the LFP responses to Periodic standards were overwhelmingly smaller than the responses to Random standards for 5% deviant probability (99/124, 80%), the imbalance in the standard response was substantially smaller when deviant probability Apoptosis inhibitor MG-132 cell line was 20% (85/147, 58%). It has been previously shown that SSA has several timescales, from hundreds of milliseconds to tens of seconds (Ulanovsky et al., 2004). In order to examine the time course of the effects shown above, we calculated the average responses

to the standards with different time resolutions along the sequence. Figure 5 shows the average LFP responses to standards (Figure 5A) and deviants (Figure 5B), as a function of the sequential position of the stimulus within the sequence for the 5% (left) and 20% (right) deviant probabilities. In Figure 5A, the blue and green bars represent the average response to the standard stimuli at four ranges of trials along the sequence (1–19, 20–80, 81–278, 279–475 for the 5% conditions; 1–4, 5–19, 20–59, 60–100 for the 20% conditions) in the Random and Periodic conditions, respectively. In Figure 5B, the red and yellow bars represent

the average response to the deviant stimuli in four ranges of trials (1:3, 4:6, 8:16, 17:25) in the Random and Periodic conditions, Rutecarpine respectively. We analyzed the data with a three-way ANOVA on time bin and sequence type, with recording site as a random factor. The main effects of time bin were significant for all conditions [5%: standards F(3,2032) = 46.01, p < < 0.01; deviants F(3,2508) = 3.22, p = 0.022; 20%: standards F(3,3076) = 47.57, p < < 0.01; deviants F(3,3172) = 4.85 p = 2.3∗10−3]. The main effect of sequence type (Periodic versus Random) was significant for the standards in the 5% conditions [F(1,2032) = 52.75, p < < 0.01] but not for the deviants [F(1,2508) = 0.16 p = 0.69]. In contrast, in the 20% conditions the main effect of sequence type was significant for the deviants [F(1,3172) = 14.5 p = 1∗10−4] but not for the standards [F(1,3076) = 0.29 p = 0.59].

The firing patterns of hippocampal interneurons are highly depend

The firing patterns of hippocampal interneurons are highly dependent on the network state, such as theta oscillations during movement or large-amplitude irregular network activity during sleep (Buzsáki, 2006, Ego-Stengel and Wilson, 2007, O’Keefe and Conway, 1978 and Ranck, 1973). Drug-free behavior-dependent firing patterns of some identified cell types have been determined recently in freely moving rats (ivy cells, PV+ basket cells; Lapray et al., 2012) and in head-fixed mice (O-LM cells, PV+ basket cells; Varga et al., 2012), although for O-LM cells this did not include sleep.

The firing patterns of identified bistratified cells in drug-free animals are unknown. We have recorded the firing of two distinct types of dendrite-targeting neuron in freely moving rats to test the hypothesis that differences in the axonal terminations of SOM-expressing cells are associated with different firing patterns under natural awake behavior and Smad inhibitor sleep. This required the recording and labeling of SOM-expressing interneurons in freely moving rats using

the juxtacellular labeling technique to identify NVP-BKM120 the cells and enabled us to quantitatively dissect the firing dynamics of these cells and compare them to PV+ basket cells (Lapray et al., 2012), which target a different subcellular domain of pyramidal cells. We have recorded the firing patterns of single interneurons using a glass electrode during periods of sleep, movement, and quiet wakefulness. Then, we either moved the electrode into a juxtacellular position or sometimes the cells spontaneously drifted close to the electrode, which made it possible to attempt labeling the cells with neurobiotin for identification of cell types. The

labeled Oxymatrine cells were assessed by immunofluorescence microscopy and tested for the presence of various molecules, including SOM and NPY. Nine identified interneurons (n = 9 rats, one cell each) were immunopositive for SOM, NPY, or both when tested by immunofluorescence microscopy and showed dendritic and axonal arborizations similar to previously described bistratified and O-LM cells (Buhl et al., 1994 and McBain et al., 1994). The recording sites were distributed over an area of 1.7 × 1.4 mm along the rostrocaudal and mediolateral axes (Figure S1A available online). Somata of bistratified cells (n = 4/5 recovered) were located in the vicinity of pyramidal cell somata (Figures 1A and S1A and S1C), had mainly radially oriented dendritic trees (n = 3/5 recovered; see exception Figure S1C), and axon collaterals distributed in strata oriens and radiatum (n = 3/5 recovered). The axonal extent of a well-labeled cell was large (Figure 1A), reaching 2.4 mm mediolaterally and 1.7 mm rostrocaudally, confirming previous results obtained in vivo (Klausberger et al., 2004). Somata (n = 4/4 tested) were immunopositive for NPY (Figures 1B and S1E) and parvalbumin (PV), the latter also expressed in dendrites and axon (Figures 1C and S1D; Table 1).

Patch-clamp recordings were made from SACs and DSGCs

Patch-clamp recordings were made from SACs and DSGCs Androgen Receptor Antagonist solubility dmso in flatmount retina in Ames medium (Sigma-Aldrich, Saint Louis, MO), equilibrated with 95% O2 and 5% CO2) at ∼35°C as described previously (Lee and

Zhou, 2006, Zheng et al., 2004 and Zhou and Fain, 1995). Spike responses of DSGCs were recorded with an on-cell loose-patch electrode (3–5 MΩ, filled with Ames medium). Whole-cell patch clamp was made from SACs using a pipette solution containing (in mM) 110 CsMeSO4, 15 CsOH, 5 NaCl, 0.5 CaCl2, 2 MgCl2, 5 ethylene glycol-bis(2-aminoethylether)-N,N,N′,N′-tetraacetic acid (EGTA), 2 adenosine 5′-triphosphate (disodium salt), 0.5 guanosine 5′-triphosphate (trisodium salt), 10 HEPES, and 2 ascorbate (pH 7.2). The same

pipette solution, supplemented with 10 mM lidocaine N-ethyl bromide (QX314), was used Tyrosine Kinase Inhibitor Library in vitro for whole-cell recording from DSGCs. Lucifer yellow (0.1%–0.3%, w/v) was included in the whole-cell pipette solution for morphological confirmation of cell types. The effects of Ca2+ channel blockers, ω-conotoxin GVIA, and ω-agatoxin IVA, which had a slow onset, were tested under perforated patch-clamp recording from SACs using electrodes (6-8 MΩ) filled with (in mM) 145 glutamic acid, 10 HEPES, 10 NaCl, and 500 μg/ml amphotericin B (titrated to pH 7.2 with 147 mM CsOH). Lysozyme (1.5 mg/ml) was added to the superfusate to prevent nonspecific binding of calcium channel toxins to the perfusion system. Flash photolysis of caged Ca2+ (DM-nitorphen) in SACs was done under dual whole-cell voltage-clamp recordings from pairs of SACs and DSGCs,

with the pipette solution for SAC recording containing (in mM) 95 CsMeSO4, 30 CsOH, 5 NaCl, 2 CaCl2, 6 Ca(NO3)2, 2 adenosine 5′-triphosphate (disodium salt), 0.5 guanosine TCL 5′-triphosphate (trisodium salt), 40 HEPES, 2 ascorbate, and 10 DM-nitrophen (pH 7.2). DM-nitrophen (1-(4,5-dimethoxy-2-nitro-phenyl)-1,2-diaminoethane-N,N,N′,N′-tetraacetic acid) was allowed to diffuse into the SAC for 5–7 min before an ultraviolet (UV) light (355 nm in wavelength, 25 ms in duration) was flashed onto the SAC dendrites. The UV light was generated by a UV laser (30 mW average output power, 3 μJ/pulse at 10kHz, Model DPSL355/30, Rapp Opto Electronik, Heidelberg, Germany), which was guided to the epifluorescence port of the microscope via a spot illumination adaptor (OSI BX, Rapp Opto Electronik) by a quartz optic fiber (940 μm in diameter) and focused in a 50 μm diameter spot on the SAC dendrites via a 60× objective lens (NA/09). The uncaging area was positioned by moving the microscope stage on which the recording chamber and micromanipulators were mounted.

Surprisingly, however, they were also eliminated between P15 and

Surprisingly, however, they were also eliminated between P15 and P25 (Figure 3G). The rate of inactive DG axon elimination was not affected by the percentage of Epacadostat ic50 mature axons that were inactivated (Figure 3H). This elimination was not just a developmental retraction but was induced by synaptic inactivity, because in DG-A::tau-lacZ mice (no TeTxLC), DG axons were maintained at P25 (Figure 3I; quantification at P15 and P23 is shown in Figure 3J). To further characterize the inactive axon elimination in DG-A::TeTxLC-tau-lacZ mice, we visualized

TeTxLC/tau-lacZ-expressing axons by staining with the antibody to β-gal. Confocal microscopy revealed that at P15, TeTxLC-expressing axons were intact at both the proximal and distal regions (Figure 4A). At P20, the proximal region of TeTxLC-expressing axons appeared still intact (Figure 4B, proximal). However, OTX015 order at the distal region of TeTxLC-expressing axons, there were clear signs of axon retraction: swollen axonal tips (retraction bulbs) and remnants of axons (axosomes) (Figure 4B, distal). These features resemble axon retraction observed during synapse elimination at the neuromuscular junction (Bishop et al., 2004). Without TeTxLC (DG-A::tau-lacZ), the distal region of DG axons remained intact at P23 (Figure 4C). We further examined the morphology of active and inactive DG axons in

DG-A::TeTxLC-tau-lacZ mice by sparsely labeling DG axons with a lipophilic tracer, DiI. Inactive (TeTxLC/tau-lacZ-expressing) axons were identified by staining for β-gal. At P18, active (β-gal−/DiI+) DG axons formed many large mossy fiber boutons with long extensions in CA3 (Figure 4D) (Amaral and Dent, 1981). In

contrast, inactive (β-gal+/DiI+) DG axons did not have any large mossy fiber boutons (Figure 4D). These results indicate that synaptic transmission plays an important role in the refinement of mossy fiber synapses and that inactive axons are retracted during development. Is the elimination of inactive DG axons initiated by the death of DG neurons? To address this question, we injected 5-bromo-2-deoxyuridine (BrdU) into DG-A::TeTxLC-tau-lacZ from mice at P7–8 to label newborn DGCs (Kee et al., 2002 and Kokoeva et al., 2005) and examined the number of surviving BrdU-positive cells at P15, P20, and P25 (Figure 4E). All of the BrdU-labeled neurons became NeuN positive mature neurons by P15 (Figure 4F), consistent with earlier reports showing that DG neurons differentiate very quickly during development (Amaral and Dent, 1981 and Hastings and Gould, 1999; also see Figure 7). The total number of BrdU-positive cells in DG-A::TeTxLC-tau-lacZ mice was similar to that in wild-type mice at P15 and P20 (Figure 4E). In addition, at P20 there was no apparent increase in the number of DG cells that are positive for activated caspase 3, a marker for apoptosis (Figure S2).

During ozone and heat treatments, juice samples were mixed at 150

During ozone and heat treatments, juice samples were mixed at 150 rpm for the entire treatment time in a shaking water bath to ensure even distribution of inoculum and dispersal of ozone. All experiments were conducted in a fume hood. Excess gaseous ozone was passed into an ozone decomposer. For enumeration of pathogens, sample aliquots (1 ml) were transferred into test tubes containing 9 ml of D/E neutralizing broth (Difco, Becton Dickinson, Sparks, MD, USA) after each treatment

and homogenized using a vortex mixer (VM-10, Daihan Scientific Co., Ltd, RAD001 clinical trial Korea). After homogenization, samples were 10-fold serially diluted with 9 ml of 0.2% sterile buffered peptone water and 0.1 ml of the samples was spread plated onto selective media (E. coli O157:H7: Sorbitol MacConkey Agar (SMAC), Difco; S. Typhimurium: Xylose Lysine Desoxycholate Agar (XLD), Difco; and L. monocytogenes: Oxford Agar Base with antimicrobial supplement Osimertinib in vitro MB Cell (MOX), MB Cell). All plates were incubated at 37 °C for 24 h before counting.

Color was measured by using a Minolta colorimeter (model CR400; Minolta Co., Osaka, Japan). The values for L*, a* and b* were recorded to evaluate the color changes of apple juice after each treatment with heat or combination of ozone and heat. Untreated apple juice was used as the control. A 2 ml of sample was poured into the bottom half of the measurement equipment. The measuring head of the colorimeter was placed on top of the measurement Methisazone equipment. Before measurement, the treated juice was cooled to about 15 °C by dipping the bottle in crushed ice. The parameter L* is a measure of lightness, a* is an indicator of redness, and the parameter b* is a measure of yellowness. All measurements were performed in triplicate. To measure ozone concentration, distilled water was substituted for apple juice, as will

be explained in the Discussion section. The ozone concentration of distilled water treated with ozone in the same way as apple juice was measured by the indigo method (Bader and Hoigni, 1981). An indigo stock reagent was prepared by the following method (Gordon and Bubnis, 2002). Indigo stock solution was prepared by dissolving 770 mg of potassium indigotrisulfonate (Sigma Aldrich Co., LLC) into a 1 l flask containing 500 ml of distilled water and 1 ml of phosphoric acid (85%) and diluting to volume (1 l) with distilled water. Indigo reagent II was prepared in a 1 l flask by adding 100 ml of indigo stock solution, 10 g of sodium dihydrogen phosphate (NaH2PO4) and 7 ml of phosphoric acid (85%). This was stirred and diluted to volume (1 l) with distilled water. The solution was stored in the dark. After each treatment, the treated sample was cooled to about 15 °C and then 90 ml of the sample was transferred to a flask containing 10 ml of indigo reagent II. The absorbance was measured by a spectrophotometer at 600 nm.

We next investigated whether there was a topographical organizati

We next investigated whether there was a topographical organization of unisensory responses in RL, and, if so, whether the two sensory maps were aligned. We used 3 × 4 grids of extracellular electrodes covering all of RL (see Experimental Procedures). For unimodal sensory input, we

separately stimulated the upper/lower and medial/lateral halves of both visual space and the whisker pad (Figure 1F). For each neuron, we computed a relative preference index as (U − L)/(U + L), where U and L are the neuronal responses to the upper or lower visual or tactile field stimulations, respectively. Then we averaged the relative spatial preference indexes of Selleck CP673451 all neurons along a given position of the grid. We finally mapped the spatial preference within RL for the upper versus lower and medial versus lateral stimulation for both modalities (Figure 1G). In line with (Marshel et al., 2011 and Wang and Burkhalter, 2007) we

found a retinotopic map within RL, but we also found a spatial segregation between responses elicited by the upper or lower aspects of the whisker pad. For example, the rostral part of RL preferentially responded to the upper visual field and to the upper whiskers. To quantify the degree of alignment between the two maps, we computed the correlation between the retinotopic and tactile maps along the same spatial direction (upper-lower and medial-lateral axes separately; Figure 1H). GSK1120212 order because We found a significant degree of spatial alignment of the somatic and visual spatial

preference maps along the upper-lower axis of the sensory space (black circles; r = 0.58, p < 0.001) but a weaker and nonsignificant alignment along its mediolateral axis (gray squares; r = 0.11, p = 0.44). We next compared bimodality and MI at the level of PSPs and of APs by IOI-targeted whole-cell recordings from layer 2/3 pyramids (n = 46 from 12 mice; Figure 2A). First, bimodal neurons were more abundant for PSPs (56% of responsive cells) compared to APs (39% of them; Figure 2B). Indeed, many neurons that were bimodal for PSPs were unimodal for APs (see example of Figure 2C): out of 24 neurons that were bimodal for PSPs, only 11 (46%) were bimodal for APs, 4 (17%) were unimodal and 9 did not respond with APs. Also, some cells with bimodal PSPs responded with APs only when V and T stimuli were simultaneously presented (multisensory—M stimulation; see example of Figure 2D). Second, we compared MI for PSPs and APs. Multisensory neurons display a response to a cross-modal combination of stimuli that is enhanced compared to the preferred unisensory stimulus (Stein and Stanford, 2008). To quantify such multisensory enhancement (ME), we computed an ME index, defined as (M−max(V,T))/max(V,T)(M−max(V,T))/max(V,T), where M is the response to M stimulation, and max(V,T) is the highest unimodal response.