Finally, SM’s residual recognition ability appeared to be consist

Finally, SM’s residual recognition ability appeared to be consistent with the response properties of right LOC and hV4. We localized SM’s structural lesion relative to retinotopically and functionally defined cortical areas. The lesion was situated within LOC, anterior to hV4 and dorsolateral to VO1/2, and was confined to a circumscribed region in the posterior part of the lateral fusiform gyrus in the RH. Typically, this region responds more to intact objects than scrambled objects (Malach et al., 1995) and damage to this circumscribed area is likely the principle etiology Alectinib nmr of SM’s object agnosia. The precise relationship between lesion localization and agnosia

has been difficult to establish to date. For example, although the lesion site of patient DF, a well-known agnosic patient who suffered a hypoxic episode (James et al., 2003), has been well documented in anatomical terms, the lesion was not sited relative to retinotopic cortex. Moreover, DF’s lesion is much more distributed than SM’s, implicating bilateral damage of ventral occipitotemporal cortex. A similar profile has been reported for agnosic patient JS, whose etiology is one of ischemic stroke; like DF, the extent of the brain

damage was extensive and bilateral (Karnath et al., 2009) making it difficult to pinpoint the critical area underlying object recognition. Fludarabine solubility dmso Our results suggest a resolution to the ongoing controversy regarding whether a unilateral or bilateral lesion is necessary for agnosia (De Renzi, 2000). As we show, a structural unilateral RH may suffice for object agnosia but because of the detrimental functional effect on the LH, the outcome essentially mimics a bilateral lesion. This finding raises important issues about Edoxaban whether the focal lesion per se serves as the underpinning of the disorder or whether a reconceptualization in terms of a more distributed neural system might be a better formulation. The first functional finding concerns the normal retinotopy obtained in SM. Although retinotopic maps can be altered extensively

in individuals post-stroke (Dilks et al., 2007), this is not so in SM. Critically, the intact retinotopy in SM precludes the ascription of any altered functionality to a foundational problem such as altered topographical organization. In addition, SM’s visual responses were relatively unperturbed, although object-related responses were reduced in temporal and parietal regions. Consistent with this, there was a reduction in the AIs across the range of object types not only in the region of the lesion, but also in other sectors of the rectangular grid. There is growing recognition that visuoperceptual impairments may arise from lesions to nodes of a distributed ventral occipitotemporal circuit, but also from a disconnection between more posterior and more anterior cortical regions.

, 2006, Giannakopoulos et al , 2003 and Näslund et al , 2000) Th

, 2006, Giannakopoulos et al., 2003 and Näslund et al., 2000). The criticism that follows is Aβ deposition itself does not necessarily predict or cause clinical AD. Such observations, however, can be understood in several other ways. First, there may be a threshold effect that involves the MK 2206 density and duration, or even rate of Aβ accumulation that together with the age of onset of the pathological processes determines the onset of the clinical

manifestations of AD. Second, as with other illnesses, there are almost certainly genetic, pathological, epigenetic, and environmental mediators that modulate progression, disease course, and manifestation of illness. For Decitabine chemical structure example, one proposed mediator involves the concept of “cognitive reserve’ that hypothesizes that factors that enhance neuroplasticity and synaptogenesis, may make an individual more resistant to the clinical manifestations of the underlying

neuropathology, thereby delaying onset of the clinical expression of the illness (Cummings et al., 1998 and Stern et al., 1999). Third, it is also possible that early subtle cognitive impairment of AD that we might now refer to as preclinical stage 3 is often not recognized in elderly people who die and come to autopsy. Some evidence for this is from the Religious Order Study where those who died without cognitive impairment and who had intermediate or high likelihood of AD based on neuropathological

examination scored 0.25 standard deviation lower on episodic memory tests than those without pathology (Bennett et al., 2006), worsening of episodic memory being the earliest and most characteristic cognitive phenotype for AD (Dubois and Albert, 2004). Finally, it has become common to explain the results from failed AD therapeutic trials with presumptive anti-Aβ therapies as evidence that the hypothesis is wrong. This is clearly inaccurate, as to date, such trials were not definitive tests of the cascade hypothesis, but rather expedient ways to test potentially disease-modifying AD therapeutics in the current clinical, regulatory, and fiscal environment. None of the putative anti-Aβ agents that have failed in pivotal phase 3 therapeutic Calpain trials were optimal or even optimized agents within their class of anti-Aβ therapeutics: Alzhemed (tramiprosate, homotaurine) was a weak aggregation inhibitor; Flurizan (tarenflurbil, R-flurbiprofen) was a γ-secretase modulator with low potency and poor brain penetration; and semagacestat, a nonselective γ-secretase inhibitor (GSI), had significant mechanism-based toxicity limiting its dosage and efficacy with respect to lowering Aβ production (Golde et al., 2010). None of these drugs showed efficacy against primary endpoints in phase 2 trials but were advanced to phase 3 nonetheless. Other anti-Aβ therapies (e.g.

2 μM ADTN no longer depressed signals through OFF bipolar cells (

2 μM ADTN no longer depressed signals through OFF bipolar cells ( Figures 4A and 4B). Both these observations are consistent with the idea that an olfactory stimulus modulates retinal function by decreasing dopamine release. The

second manipulation was to antagonize the action of endogenous dopamine by injection of 100 nM SCH 23390 (7-chloro-3-methyl-1-phenyl-1,2,4,5-tetrahydro-3-benzazepin-8-ol), a selective antagonist of dopamine D1 receptors (Mora-Ferrer and Neumeyer, 1996 and Bourne, 2001). SCH 23390 injection resulted in a complete impairment of luminance signaling through OFF bipolar cells (Figures 5A and 5B). In contrast, the maximum amplitude of the response in ON bipolar cells was not significantly find more affected, although the light sensitivity (I1/2) was increased by a factor of 3.8 ( Figures 5C and 5D). Antagonizing D1 receptors therefore caused effects qualitatively similar to an olfactory stimulus: a selective decrease in

the gain of signaling through the OFF pathway (cf. Figures 1 and 5). To investigate the role of D2 receptors, we used the antagonist sulpiride at a concentration of ∼2 μM (Lin and Yazulla, 1994 and Mora-Ferrer and Gangluff, 2000). Sulpiride altered the luminance-response function in two ways. First, the sensitivity increased sufficiently to reduce threshold by ∼2 log units, likely reflecting the potentiation of rod inputs (Ribelayga et al., 2008). Second, the luminance-response selleck screening library relation did not simply rise monotonically, but instead passed through a maximum (Figures 5E and 5F). Despite these changes in circuit function, the maximum response to luminance was reduced by 29% ± 5.6% when methionine was applied after injection of sulpiride, an effect similar to that of olfactory stimulation under control conditions (Figures 5E and 5G). Sulpiride also increased sensitivity of signals through ON bipolar cells (0.85 found log units), but methionine had no effect on the amplitude of these responses (Figures 5H–5J). An olfactory

stimulus therefore continued to cause a selective reduction of responses through the OFF pathway when activation of D2 receptors was blocked. Together, the results in Figures 4 and 5 indicate that cross-modal regulation of retinal function depends primarily on the activity of D1 receptors. To test further the idea that activation of the ORC acts by decreasing dopamine level in the retina, we attempted to prevent these changes without interfering with the activity of dopamine receptors. Our strategy was to inject vanoxerine (GBR 12909; 2 μM), a potent and specific blocker of the transporters involved in dopamine reuptake from extracellular space and into secretory vesicles (Reith et al., 1994 and Singh, 2000). Vanoxerine administration has been reported to result in a small but steady increase in extracellular dopamine concentration, followed by a persistent “clamp” once both uptake and release are blocked (Rothman et al., 1991, Lima et al., 1994, Reith et al.

02 in ninth individual, Figure 8B) Thus, despite the probabilist

02 in ninth individual, Figure 8B). Thus, despite the probabilistic nature of activity onsets in slow waves, the underlying process is not entirely stochastic and reflects the cumulative drive of afferent synaptic input. By examining simultaneous depth EEG and single-unit recordings in multiple regions of the human brain, we show that ON and OFF periods of spiking activity and corresponding EEG slow waves typically involve a limited number of brain regions. Alectinib order Similarly, sleep spindles are mostly local. In addition, slow waves

have a tendency to propagate from prefrontal cortex to the MTL and hippocampus. We also found that activity onsets in individual waves reflect the afferent synaptic drive to a given region. The main observation reported here is that most sleep slow waves are in fact confined to local regions (i.e., detected in a minority of brain areas). Specifically, in most cases we found that when some brain regions were in an ON state, neurons in other brain regions were completely silent. The unique dataset used here was instrumental in unequivocally establishing this phenomenon, since (1) the large size of the human brain facilitated simultaneous recordings across multiple distant sites, and (2) selleck compound the simultaneous

recording of depth EEG, MUA, and spiking of individual neurons allowed us to establish that differences between depth EEGs in different regions reflect local activities rather than noise. Typically, less than 30% of monitored brain regions were involved in each slow wave event. Although our sampling was mostly limited to medial brain areas, it represented activity in multiple (8–12) regions across lobes and hemispheres, so it is reasonable to infer that

these results can be generalized to the entire brain. When global ON and OFF states were observed, they were associated with large slow Edoxaban waves in scalp recordings, usually during deep sleep early in the night, or with K-complexes throughout the night. Local slow waves tended to co-occur across homotopic prefrontal regions, but not across homotopic regions in the MTL (Figure 4E), supporting the role of callosal pathways in slow wave synchronization and propagation (Amzica and Steriade, 1995a). Medial prefrontal regions including the anterior cingulate and orbitofrontal cortex were among the most involved in slow wave occurrence. The demonstration that slow waves are mostly local shows that examples of isolated slow waves (Mohajerani et al., 2010 and Sirota and Buzsaki, 2005) constitute the rule rather than the exception. It is also consistent with evidence that the intensity of sleep slow waves can vary across brain regions (Finelli et al., 2001), that sleep can be regulated locally (Huber et al., 2004 and Huber et al., 2006), and that multiple local generators could contribute to EEG SWA (Murphy et al., 2009 and Riedner et al., 2007).

, 2011) The way in which perceptual learning is represented in t

, 2011). The way in which perceptual learning is represented in the cortex may be dependent on the nature of the Dasatinib datasheet discrimination task. It is important, for example,

to distinguish between learning on lower order properties, such as those associated with inputs to the cortex (somatosensory vibration or acoustic frequency), feedforward properties such as orientation tuning, and the higher-order properties that are dependent on context, such as three-line bisection, vernier discrimination, or contour detection and shape discrimination. The cortical changes associated with contextually dependent perceptual learning have to account for its specificity. In fact, the way learning is represented in these tasks is to influence contextual interactions that are relevant to that task. This is exemplified by changes

in contour integration accompanying learning in a contour detection task (Figure 7; Li et al., 2008) and changes in modulation of responses by changing the distance between parallel lines in a three-line bisection task (Crist et al., 2001; Li et al., 2004). By enhancing the modulation in neuronal tuning to stimulus components that are relevant to the task, learning increases the task relevant information conveyed by neurons. As subjects learn a task, there is a change in the functional properties of neurons encoding the information involved in the task. Remarkably, one can see this occur even in V1. As shown in Figure 7, the ability to detect a contour composed of collinear line segments embedded in a complex background selleck chemical improves with practice. Longer contours made of a larger number of line segments are easier to detect than those made of fewer line segments, and the number of segments required to reliably detect the contour decreases with practice. One can see from the black dashed psychometric Sclareol curve the increase in detectability as a function of the number of line elements. This represents the animals’ performance early in the period of training, during the first week. This curve steepens with practice (red dashed curve), showing

the improvement in performance as a result of perceptual learning in the task. If one measures the contour related responses in V1, there is a corresponding steepening of the neurometric curve that tells how well an ideal observer can detect the embedded contours of different lengths simply based on neuronal responses. Perceptual learning can enable neurons to carry information that is required to perform complex visual discrimination tasks, not only for contour detection, as described above, but for discriminating the shapes of contours embedded in complex scenes. For animals trained in a task requiring discriminating a circle, a straight line or a wave shape, neurons take on selectivity for related shapes (Figures 8 and 9).

, 2011) Taken together with our findings here on class 5 semapho

, 2011). Taken together with our findings here on class 5 semaphorins and their role in regulating the establishment of neural connectivity within the retina, these

observations suggest that inner retinal lamination is initially orchestrated through a combination of spatially distinct transmembrane GSK-3 beta phosphorylation semaphorin repellent cues. Sema6A, which is expressed in inner retinal neuron subtypes, directs a small subset of select laminar stratification events within the IPL; in contrast, Sema5A and Sema5B guidance cues present in outer retinal neurons control inner retinal lamination by constraining a major portion of inner retinal neurites to the IPL. It seems likely that a combination of transmembrane semaphorin short-range

repulsive interactions and attractive interactions mediated by CAMs facilitates laminar stratification by regulating synapse formation among select neuronal subtypes that together participate in the formation of specific neural circuits in the retina. Outer retinal lamination events within the OPL are controlled by as yet unidentified repellents or attractants. Our observations showing how transmembrane R428 class 5 semaphorins and their PlexA1 and PlexA3 receptors function during retinal development provide an example of repulsive guidance cue regulation of mammalian IPL lamination and segregation of inner retinal neurites from the outer retina. Correct development of inner retinal lamination is critical for appropriate physiological retinal responses, demonstrating that establishment of retinal laminar organization and retinal circuit function are intimately related. It will be of interest to determine whether similar molecular mechanisms facilitate the elaboration of laminar organization in other regions of the CNS, including the spinal cord and the cerebral cortex, and to understand how laminar organization in these regions of the CNS is related to function. Defining this relationship will advance our understanding of lamination as an organizing however principle throughout the nervous system. The day of vaginal plug observation was designated

as E0.5 and the day of birth as P0. Genetically modified mouse lines and targeting strategies for the generation of Sema5A−/− and Sema5B−/− mice are described in Supplemental Experimental Procedures. Immunohistochemistry was performed as previously described (Matsuoka et al., 2011). The primary antibodies used in this study are listed in Supplemental Experimental Procedures. In situ hybridization was performed on either fresh-frozen or PFA-fixed retina sections (20 μm thickness) as described previously (Matsuoka et al., 2011). Digoxigenin-labeled antisense riboprobes specific for the coding sequences of Sema5A (3353–3860 bp), Sema5B (2808–3366 bp), PlexA1 (381–1314 bp), and PlexA3 (4977–5616 bp) were used for in situ hybridization.

, 2006 and Ishizuka et al , 2006) Moreover, while retinoids were

, 2006 and Ishizuka et al., 2006). Moreover, while retinoids were already well known to be present find more in large quantities in embryonic tissues and in the retina, it was soon found that mature mammalian brains ( Deisseroth et al., 2006 and Zhang et al., 2006), and indeed all vertebrate tissues thus far examined (e.g., Douglass et al., 2008) contain sufficient all-trans retinal for microbial opsin genes to define a single-component strategy. By 2010 the major classes of ion-conducting microbial opsins (including bacteriorhodopsin, channelrhodopsin, and halorhodopsin) had all proven to function as optogenetic control tools in mammalian neurons, as described

below. Since earlier, multicomponent efforts for photosensitization of cells (for example, involving cascades of multiple genes or combinations of genes and custom organic

chemicals (Zemelman et al., 2002, Zemelman et al., 2003, Banghart et al., 2004, Lima and Miesenböck, 2005, Kramer et al., 2005 and Volgraf et al., 2006) have been recently reviewed (Gorostiza and Isacoff, 2008 and Miesenböck, 2009), here we provide a primer focusing on single-component Erastin optogenetics, delineating guiding principles for scientific investigation and summarizing the enabling technologies for neuroscience application. However, most of the techniques developed for this approach (ranging from genetic targeting methods, to addressing experimental confounds, to intact-system light delivery methods) will be relevant to any biological system or optogenetic strategy. We do not attempt to review in any form the very large number of papers and results that have emerged in this field, nor to address every technique, reagent, and device linked to optogenetics. Rather, here we highlight limitations, challenges, and obstacles in the field and outline general principles for designing, conducting, and reporting optogenetic experiments. Optogenetics is not simply photoexcitation or photoinhibition of targeted cells; rather, optogenetics must deliver gain or loss of function of precise events—just as in genetics, where

single-gene manipulations are the core currency of the field. This means that in neuroscience, millisecond-scale precision is essential to true optogenetics, to keep pace with the known over dynamics of the targeted neural events such as action potentials and synaptic currents. Moreover, this level of precision must be operative within intact systems including freely moving mammals. All strategies to achieve optical control, including those involving microbial opsin genes, initially displayed serious limitations in meeting this goal. The multicomponent character, longer-timescale temporal properties, and/or requirement for high-intensity UV light characteristic of the earlier strategies (Zemelman et al., 2002, Banghart et al., 2004, Lima and Miesenböck, 2005 and Kramer et al.

In vertebrates, H2S drastically increases under hypoxic condition

In vertebrates, H2S drastically increases under hypoxic conditions to levels that are inversely correlated

with tissue O2 levels (Olson, 2011, Olson et al., 2006 and Peng et al., 2010). H2S is endogenously produced by multiple types of enzymes in animals and is constantly oxidized, so its increase might be directly find more regulated by local O2 levels to mediate effects of hypoxia (Chen et al., 2004, Kimura, 2010, Olson, 2011, Peng et al., 2010 and Singh et al., 2009). In both C. elegans and mammalian cells, H2S has been shown to promote HIF-1 activity and upregulate HIF-1 target genes ( Budde and Roth, 2010 and Liu et al., 2010). However, the mechanism by which H2S elicits its effects on HIF-1 has been unknown. Our findings demonstrate an essential role of CYSL-1 in mediating H2S upregulation of HIF-1 target genes through CYSL-1 interaction with www.selleckchem.com/products/Adriamycin.html the EGL-9 C terminus. A recent study found that cysl-1 mutants are sensitive to H2S and hypothesized that CYSL-1 might act in a pathway downstream of HIF-1 to enzymatically assimilate H2S ( Budde and Roth, 2011). Unexpectedly, our studies show that CYSL-1 acts upstream of HIF-1 by directly inhibiting EGL-9 in a manner that is modulated by H2S accumulation.

Interestingly, both H2S and RHY-1 appear to regulate HIF-1 activity in a VHL-1-independent manner ( Budde and Roth, 2010 and Shen et al., 2006), consistent with the notion that CYSL-1 inhibits EGL-9 and mediates H2S activation of HIF-1 independently of EGL-9 hydroxylase activity. Bisulfide is known to bind to an allosteric regulatory site of Salmonella OASS proteins, which are highly similar to CYSL-1 in C. elegans, and can stabilize the interaction between OASS and the SAT C terminus ( Salsi et al., 2010). H2S inhibits mitochondrial

cytochrome-C oxidase and can also directly modify target proteins via sulfhydration ( Mustafa et al., 2009). Although CYSL-1 has only weak intrinsic sulfhydrylase activity in vitro, it remains possible that H2S might modify EGL-9 via CYSL-1-modulated sulfhydration to facilitate sequestration of EGL-9 by CYSL-1. The detailed mechanism by which H2S and its in vivo derivatives modulate CYSL-1 and EGL-9 to regulate HIF-1 remains to be investigated. PD184352 (CI-1040) CYSL-1-homologous CBS proteins in mammals are known to be major H2S-biosynthetic enzymes (Chen et al., 2004 and Singh et al., 2009), and we suggest that the pathway we identified is fundamentally similar in nematodes and mammals (Figure 7A). In nematodes, H2S and CYSL-1 regulate HIF-1 through EGL-9. In mammals, H2S also regulates HIF proteins (Li et al., 2011 and Liu et al., 2010), and we propose that CYSL-1-like CBS proteins generate endogenous H2S to modulate HIF. In mammals, HIF activation protects tissues from reperfusion injury (Loor and Schumacker, 2008); we propose that the C.

Yet, little is known about how the time-varying firing rate of se

Yet, little is known about how the time-varying firing rate of sensory neurons control the specific motor sequences Epigenetic pathway inhibitors underlying ongoing, complex motor behaviors. Collision avoidance and

escape behaviors provide a favorable model to study this question. They are critical for survival and are implemented by specialized neural circuits in several species (Wang and Frost, 1992, Graziano et al., 1994, Wicklein and Strausfeld, 2000, Yamamoto et al., 2003, Preuss et al., 2006, Oliva et al., 2007 and Fotowat et al., 2009). In locusts, the third neuropil in each of the two optic lobes contains an identified neuron, the lobula giant movement detector (LGMD) that responds specifically to objects approaching on a collision course in its associated visual hemifield, or their 2D projection: looming stimuli (Hatsopoulos et al., 1995, Schlotterer, 1977, Rind and Simmons, this website 1992, Judge and

Rind, 1997 and Peron and Gabbiani, 2009). Each LGMD synapses in the brain onto the descending contralateral movement detector (DCMD) neuron, such that their spikes are in one-to-one correspondence (Rind, 1984 and Killmann and Schurmann, 1985). In response to looming stimuli, the firing rate of these neurons gradually increases, peaks, and rapidly decreases before expected collision (Gabbiani et al., 1999). Similar response profiles have now been described in neurons of wide-ranging species (pigeon: Sun and Frost, 1998; frog: Nakagawa and Hongjian,

2010; fish: Preuss et al., 2006; fruit fly: Fotowat et al., 2009). In locusts, this response profile is robust to a broad spectrum of stimulus changes, suggesting that it may play an important role in the generation of escape behaviors (Gabbiani et al., 2001). From the brain, each DCMD axon projects through the contralateral nerve cord to motor centers involved in jump and flight steering (O’Shea et al., 1974 and Simmons, 1980). In particular, the DCMDs make both direct and indirect synaptic contacts with the fast extensor tibia (FETi) motoneuron of the hindleg and indirect connections to the flexor tibia motoneurons (Burrows and Rowell, 1973, Pearson et al., 1980 and Pearson and Robertson, 1981). The involvement of DCMD activity in jump escape behaviors has been studied, but its role remains Etomidate unclear (Fotowat and Gabbiani, 2007, Burrows, 1996 and Santer et al., 2005). Up to now, it was impossible to record simultaneously from the DCMD and motoneurons during freely executed, visually guided jump escape behaviors. Consequently, it was not possible to observe how sensory and motor activities are related on a trial-by-trial basis. To achieve this goal, we built a telemetry system capable of wireless transmission of neural and muscle recordings. This system was sufficiently small that locusts could carry it as a backpack and still respond to looming stimuli by jumping.

APV alone had no effect on EPSCs (n = 3, p = 0 21), and CNQX alon

APV alone had no effect on EPSCs (n = 3, p = 0.21), and CNQX alone decreased EPSCs by 70.3% ± 30.3% (n = 3; p < 0.01). IPSCs, measured at 0mV, ranged from 19.2 to 1061.1 pA and averaged 316.3 ± 284.4 pA (n = 16). IPSCs were detected in 16 out of 16 cells (100%) recorded at 0mV, and EPSCs were detected in14 out of 15 cells (93.33%) recorded at −70mV. When synaptic currents were recorded at −40mV, 11 out of 12 cells (91.67%) exhibited both IPSCs and EPSCs. IPSCs were delayed relative

JQ1 purchase to EPSCs, with latencies of onset of EPSCs and IPSCs averaging 3.7 ± 0.8 ms (n = 15) and 10.5 ± 1.2 (n = 16), respectively (Figure S2). This delay difference indicates that inhibition from AON axons to MCs is disynaptic and excitation is most probably elicited by direct AON-to-MC synaptic connections. The relative timing and contribution of the two components was clearly evident when responses were recorded at a holding potential of −40mV (Figure 2E).

As a consequence of the different delays, membrane potential recordings Alectinib price from MCs showed brief depolarization upon light stimulation, followed by hyperpolarization (Figure 2F). When recorded at resting potential (Vm ∼−55mV), the average amplitudes of EPSPs and IPSPs were 0.57mV ± 0.25mV and 2.2mV ± 1.8mV, respectively (n = 13). We investigated the source of the synaptic currents elicited by stimulation of AON axons, starting with excitation. We considered two possibilities: direct excitation of MCs and indirect excitation through excitatory local neurons. Light-evoked excitation is DNA ligase blocked by ionotropic glutamate receptor blockers as described above (Figure 2) and shows an amplitude dependence with a reversal potential of 5.8mV ± 11.6mV (n = 4 cells). We next tested monosynaptic excitation directly using a previously described method (Petreanu et al., 2007; Gire et al., 2012; Hagiwara et al., 2012), in which transmitter release is evoked directly from ChR2-expressing axons in the presence of 1 μM tetrodotoxin (TTX) to remove polysynaptic

excitation, 100 μM 4-AP to enhance axonal depolarization, and 10 μM gabazine. Four out of six recorded cells showed an excitatory current under these conditions, with an average amplitude of 10.5 ± 9 pA (range of 4.5–24 pA). Because the excitatory response persisted under these conditions, it is at least partly due to direct glutamate release from AON neurons without the involvement of intermediary neurons. We also tested if the response was due to extrasynaptic spillover of glutamate by using the weak competitive glutamate antagonist γ-DGG (Gire et al., 2012). Application of 500 μM γ-DGG, which is known to significantly attenuate spillover-mediated transmission between ETCs and MCs (Gire et al., 2012), did not significantly affect light-evoked EPSC amplitudes (percent block 0.2% ± 30%, n = 3 MCs).