Despite this caveat, the study provides an important challenge to

Despite this caveat, the study provides an important challenge to our understanding of the role of

gain fields in spatial representation and computation. A number of outstanding questions remain. First, are these findings robust across different cortical areas known to contain eye-position signals, or are they specific to LIP? Another recent study of gain field dynamics (Morris et al., 2012) shows similar lags for eye-position signals in LIP, such that most LIP neurons do not provide reliable information about eye position until around 200 ms after an eye movement. Interestingly, while this result is consistent with Xu et al. (2012), these results were not reproduced in nearby dorsal visual areas VIP, MT, and MST. Instead, eye-position signals in these areas appear to learn more update much more rapidly, right around the time of the saccade and in some cases even slightly before the movement begins. These apparent inconsistencies in the temporal dynamics of gain fields across cortical areas produce a tension that requires resolution. Nevertheless, caution must be exercised in PD-1/PD-L1 inhibitor drugs drawing too strong a conclusion, since the paradigms differ in substantial ways: Morris et al. (2012) investigate eye-position modulation during static

fixation, whereas Xu et al. (2012) examine modulation in response to a visual target. A second outstanding question is whether the findings about the dynamics of eye-position gain fields in LIP apply to other motor systems or are specific to the oculomotor system. The authors imply that their findings have wide application, but this remains to be seen. Unique features of the oculomotor system could weigh against the extensibility of Xu et al.’s reported results. Most prominently, the oculomotor system—unlike many other motor systems—does not generally require an explicit computation of target during location in supraretinal (e.g., head-centered) coordinates, since typically only the retinal difference vector (the difference between the fovea and the retinal position of the target) is required for saccade programming. Consequently, the use or disuse of eye-position gain fields

for computations related to saccade programming might not accurately reflect how other motor systems use them, especially where reference frame transformations are required (Pouget and Snyder, 2000). Finally, Xu et al.’s results should lead researchers in the field to reflect more broadly about what other roles (if any) gain fields might play in motor planning and sensorimotor transformations. Given their widespread presence throughout the brain, it is incumbent upon the field to embrace the purely negative answer that they play no functional role only as a last resort. Xu et al. (2012) hypothesize that the temporal properties of these eye-position signals, while unsuited for use in real-time saccade programming, might be deployed in a more ancillary way as a kind of feedback to calibrate motor efference copy signals.

3 This is supported by studies showing that foot positioning at c

3 This is supported by studies showing that foot positioning at contact in runners wearing

minimal footwear is more similar to the barefoot condition than to the conventional shoe condition even if they continue to contact first on the heel.11 and 13 It is thus possible that running form varies between footwear conditions in subtle ways that were not measured here, and future studies that attempt to undertake finer scale measurement of kinematic variables in the field are needed. The results of this study provide insight into the role of footwear in determining foot Antidiabetic Compound Library clinical trial strike pattern. They indicate that the majority of barefoot runners tend to contact the ground on the midfoot or forefoot when running on an asphalt road. This contrasts with the typical rearfoot striking pattern observed in conventionally shod runners on hard surfaces. Results also show that a minimally cushioned running shoe may not perfectly simulate barefoot running, with frequency of midfoot and forefoot striking being approximately equal to rearfoot striking. “
“Barefoot (BF) running has recently increased in popularity among runners with a perception that it is more natural and may result in fewer injuries. In fact, the top reason runners report for choosing to transition to BF or minimal

running is the notion of injury prevention.1 The potential for a lower risk of injury Stem Cells inhibitor is postulated based on strengthening of the foot,2 and changes in loading parameters due to alterations in running pattern associated with BF running.3 It has been documented that up to 89% of traditionally shod runners land on their heels or with a rearfoot strike (RFS).4 and 5 This strike pattern is associated

with an impact transient in the vertical ground reaction force (VGRF), followed by a propulsive peak. The impact transient appears as a distinct change in the positive slope of the VGRF trace, sometimes characterized by a local maximum or impact peak (VIP). The rate of development of the VGRF is referred to as the loading rate (Fig. 1A). High loading rates and impact transients have been associated with a number of common running-related injuries such as tibial stress not fractures,6 patellofemoral pain,7 and plantar fasciitis.8 Most habitual BF runners land on the ball of their foot, referred to as a forefoot strike (FFS) with their foot in a relatively flat orientation.9 This pattern typically has a single propulsive peak in the VGRF, lacking a distinct vertical impact transient.10 Elimination of this impact transient is accomplished by reducing vertical stiffness of the body. Vertical stiffness can be assessed using a simple mass spring model11 and 12 which works well for an FFS pattern. However, when impact transients are present, a dual stiffness model, such as described by Hunter,13 should be used. The influence of strike pattern on the medial and lateral components of the ground reaction force (GRF) is not well established.

, 2011) Roesch et al (2009) reported that nucleus accumbens neu

, 2011). Roesch et al. (2009) reported that nucleus accumbens neurons integrate information about the value of an expected reward with features of the motor output (i.e., response speed or choice) that occur during decision making. DA release may set a threshold for worthwhile cost expenditures, and under some Selleck Alectinib circumstances may provide an opportunistic drive for exploitation of resources (Fields et al., 2007; Gan et al., 2010; Beeler et al., 2012). This suggestion is consistent with the proposed involvement of accumbens

DA in the behavioral economics of instrumental behavior, particularly in terms of cost/benefit decision making (Salamone et al., 2007, 2009). As stated above, organisms typically are separated from primary motivational stimuli or goals by obstacles VE-821 in vitro or constraints. Another way of saying this is that the process of engaging in motivated behavior requires that organisms overcome the “psychological distance” between themselves and motivationally relevant

stimuli. The concept of psychological distance is an old idea in psychology (e.g., Lewin, 1935; Shepard, 1957; Liberman and Forster, 2008) and has taken on many different theoretical connotations in different areas of psychology (e.g., experimental, social, personality, etc.). In the present context, it is simply used as a general reference to the idea that objects or events are often not directly present or experienced, and therefore organisms are separated along multiple dimensions

many (e.g., physical distance, time, probability, instrumental requirements) from these objects or events. In various ways, mesolimbic DA serves as a bridge that enables animals to traverse the psychological distance that separates them from goal objects or events. Multiple investigators have phrased this in diverse ways or emphasized different aspects of the process (Everitt and Robbins, 2005; Kelley et al., 2005; Salamone et al., 2005, 2007, 2009; Phillips et al., 2007; Nicola, 2010; Lex and Hauber, 2010; Panksepp, 2011; Beeler et al., 2012; see Figure 2), but many of the functions in which accumbens DA has been implicated, including behavioral activation, exertion of effort during instrumental behavior, Pavlovian to instrumental transfer, responsiveness to conditioned stimuli, event prediction, flexible approach behavior, seeking, and energy expenditure and regulation, are all important for facilitating the ability of animals to overcome obstacles and, in a sense, transcend psychological distance. Overall, nucleus accumbens DA is important for performing active instrumental responses that are elicited or maintained by conditioned stimuli (Salamone, 1992), for maintaining effort in instrumental responding over time in the absence of primary reinforcement (Salamone et al.

Perhaps the ICA underestimated this spatial segregation, causing

Perhaps the ICA underestimated this spatial segregation, causing voxels from one network to distort the task-component loadings from the other and masking the contribution of a diffuse

higher-order “g” factor. This objection is highly unlikely for several reasons. First, LY2157299 in vitro while ICA seeks to maximize independence, it does not necessarily derive completely independent components. For example, in the current study, the MDwm and MDr components did show the expected negative correlation across voxels (r = −0.19). Second, such a close conformity between the second-order correlations from the simulated and behavioral models would have been highly unlikely to occur by chance alone if the ICA had failed. Furthermore, if the networks are spatially separable, then it should be possible to take relatively unmixed measures of their task-related activations by examining the centers of each cluster, where there is minimal network overlap. For example, when mean task activation levels were extracted from 5 mm spherical ROIs centered on peak IFO and IFS coordinates within the MDwm and MDr networks bilaterally, a marked double dissociation was evident across tasks. Specifically, there was either strong coactivation of regions or strong activation in one region and virtually no activation in the

other dependent on the task context (Table S5). This is clearly the pattern of results that would be expected if the ROIs were placed exclusively within functionally dissociable and spatially separable networks. Nonetheless, when the 2F simulations see more were rerun based on these IFS and IFO activation levels, the second-order

correlation between the estimated oblique components was not diminished but, rather, formed a precise match to the Internet behavioral data (r = 0.47, SD ± 0.02). Thus, while the contribution of diffuse factors should not be entirely discounted, the results accord particularly closely with the view that the higher-order “g” component is primarily accounted for by cognitive tasks recruiting multiple functionally dissociable brain networks. Indeed, from a phenomenological to perspective, the idea that tasks tend to corecruit multiple functional brain networks makes intuitive sense, as generating a task that depends on any single cognitive process is likely to be rather intractable. Consider a simple working memory task, in which the spatial locations of a sequence of flashes must be observed, maintained, and repeated (spatial span). Even in this simple context, the participant must comprehend the written instructions, otherwise, they may report the correct locations but in the incorrect sequence. More importantly, people often apply chunking strategies when encoding information in short-term memory in order to generate a more efficient memory trace.

Briefly, a 25 mm glass coverslip (thickness, 0 08 mm) was glued o

Briefly, a 25 mm glass coverslip (thickness, 0.08 mm) was glued over a 22 mm hole in the bottom of a 35 mm tissue-culture dish using silicone sealant. Dissociated neuronal cultures from rat hippocampi at P1 and P2 were prepared. Neurons were plated onto prepared 35 mm tissue culture dishes at a density of 1 × 106 cells per dish. The age of cultured neurons

was counted from the day of plating (1 DIV). Neurons at 7–10 DIV were transfected using a standard calcium phosphate precipitation method and allowed to grow to maturity (>3 weeks) to be imaged. Neurons 5 FU were transfected with equivalent concentrations of plasmids encoding one of six different GFP-tagged htau constructs (WT, P301L, AP, AP/P301L, E14, or E14/P301L htau). Neurons transfected with htau constructs were cotransfected with DsRed to visualize dendritic spines. Some neurons were transfected with GFP alone to visualize dendritic spines. The culture dishes fit tightly in a homemade holding chamber on a fixed platform BI 2536 solubility dmso above an inverted epifluorescent microscope sitting on an X–Y translation stage (Burleigh Instruments, Fishers, NY). Following the protocol described in Strasser et al. (2004), hippocampal and glial cultures were prepared from E16 and P1–2 mice, respectively. Glial cultures prepared from P1–2 mouse cortices were plated on tissue culture dishes in glial plating media (minimal essential medium [MEM] with Earle’s salts, 10%

fetal bovine serum [FBS] or newborn calf serum, 2 mM glutamine, 10 mM sodium pyruvate, 10 mM HEPES, 0.6% glucose, 100 U/ml penicillin and 100 μg/ml streptomyocin). For primary hippocampal neuron cultures, approximately 1.5 × 104 cells were plated on sets of 5 × 12 mm coverslips that had been previously coated with Poly-D-Lysine (100 μg/ml) + laminin (4 μg/ml) in neuronal plating media (MEM with Earle’s salts, 10 mM HEPES, 10 mM sodium pyruvate, GPX6 0.5 mM glutamine, 12.5 μM glutamate, 10% FBS, and 0.6% glucose). Each set of 5 coverslips was maintained in a 35 mm dish and each dish corresponded to 1 mouse. Approximately 4 hr after plating, the media was replaced with either neuronal growth medium

(Neurobasal media with B27 supplement, 0.5 mM glutamine) that had been conditioned on glia for 24–48 hr immediately prior to use. Mice were genotyped by PCR analysis of tail snip lysates using transgene-specific primers. Miniature EPSCs were recorded from cultured dissociated rat and mouse hippocampal neurons at 22–30 DIV with a glass pipette (resistance of ∼5 MΩ) at holding potentials of −55 mV and filtered at 1 kHz as previously described (Liao et al., 2005). Input and series resistances were checked before and after the recording of mEPSCs, which lasted 5–20 min. There were no significant difference in the series resistances and input resistance among various groups of experiments. One recording sweep lasting 200 ms was sampled for every 1 s.

They allow the same ganglion or amacrine cell to be visually targ

They allow the same ganglion or amacrine cell to be visually targeted for recording. Even if several cell types express the fluorescent marker, one can use the anatomy of the cells to separate them, so that a single type can repeatedly be patched or imaged. An example where the expression is almost “pure” is the Jam-B cell, a ganglion cell type with a curious, wedge-of-pie shape and its own version of direction selectivity (Kim et al., BI-2536 2008). This cell had been reported in anatomical surveys, but no particular attention had been paid (indeed, one study—by the author of this review—mistakenly classified them as developmental accidents) until a mouse in which

they were selectively labeled was available. These mice will also be useful for validating the retina neurome, because they provide an additional criterion for what constitutes a cell type, but they have a limitation when it comes to accounting for the retinal cell populations. The creation of these mouse strains

is still a highly inexact science. This compromises the endgame—the attempt to learn when the census of cell types is complete. Most of the strains that exist so far show mixed expression of the marker in several cell types, or expression in only parts of a true cell population. And there is no way to know anything about the cells that are NOT labeled—no way to know where the labeled cell stands in the whole population of ganglion cells and how many unlabeled cell types remain. How many cells remain for which no one has yet hit upon an CYTH4 effective promoter strategy? What is the true mosaic of genetically marked cells, ON-1910 when one cannot count on reporter expression to mark all of the cell type’s members? Sooner or later, when the molecular fundamentals of gene expression are under better experimental control, a precise algorithm for the creation of cell type-specific lines will be devised and these obstacles will be overcome. In the meantime, other methods will also be required. An approach that avoids the sampling problem is provided by high throughput electron microscopy,

also known as connectomics (Anderson et al., 2009; Briggman and Denk, 2006; Denk et al., 2012; Denk and Horstmann, 2004; Kleinfeld et al., 2011; Lichtman and Sanes, 2008; Seung, 2009). The method, a descendent of early, hand-implemented, serial sectioning (Cohen and Sterling, 1991), requires still-developing computational methods, but even now it is extremely powerful. A small area of retina is serially sectioned and high-resolution images of every cell are reconstructed. In these images, the synapses between the cells can be identified and connections traced. Furthermore, the reconstruction can be made to include a cell of known physiological function, so that synaptic contributions to that particular cell’s response are identified (Briggman et al., 2011).

Emerging clues to selectivity at the subunit level, especially in

Emerging clues to selectivity at the subunit level, especially in the context of nicotine dependence, concern the α5 subunit. Figure 1 shows that this subunit never participates at the agonist binding interface between α and β subunit but occupies a fifth or “auxiliary” position. In rodent brain, most α5∗ nAChRs are thought to be (α4)2(β2)2α5 pentamers (Gotti et al., 2006 and Albuquerque et al., 2009). In all known animals, the α5, α3, β4 genes form a cluster. Indeed, (α3)2(β4)2α5 pentamers are widespread in the peripheral nervous system, in the medial habenula, and in some nonneuronal cell

types. NU7441 mw [We do not emphasize (α3)2(β4)2α5 nAChRs or α3β4 nAChRs, because such nAChRs have relatively low nicotine sensitivity and relatively low susceptibility to upregulation.] Single-nucleotide polymorphisms found in the human α5, α3, β4 gene cluster are associated with nicotine dependence and

its age-dependent onset; number of cigarettes smoked mTOR inhibitor per day and “pleasurable buzz” elicited by smoking; alcoholism, sensitivity to the depressant effects of alcohol, and age of alcohol initiation; cocaine dependence; opioid dependence; lung cancer; and cognitive flexibility (Erlich et al., 2010, Hansen et al., 2010, Improgo et al., 2010, Saccone et al., 2010 and Zhang et al., 2010). A major “risk allele” is in a noncoding region of α5 and is associated with decreased expression of α5 subunit mRNA (Wang et al., 2009). A second “risk allele” occurs in the coding region, within the M3-M4 loop, and also produces decreased function of (α4)2(β2)2α5 nAChRs (Wang et al., 2009 and Kuryatov

et al., 2011). Furthermore in experiments using chronic nicotine exposure in rats, (α4)2(β2)2α5 nAChRs are not upregulated, but (presumptive) (α4)2(β2)3 nAChRs in the same brain region are (Mao et al., 2008). Summarizing the available data, the “risk alleles” may decrease the fraction of (α4)2(β2)2α5, increasing that of α4β2 nAChRs. Because α4β2 nAChRs are the most susceptible to nicotine-induced upregulation, the data again seem consistent with the idea that selective upregulation of α4β2 nAChRs underlies whatever nicotine dependence. The potential power of α4β2 upregulation to explain the initial events of nicotine dependence thus derives from its selectivity, displayed at every level of organization: regional, neuronal, cellular, and stoichiometric. Selective upregulation would directly result in modified neuronal excitability and neuronal interactions. As noted above, in the context of nicotine dependence, selective upregulation presently has been studied in detail only in midbrain and in the perforant path. Thus it remains an audacious hypothesis that the initial stages of nicotine dependence can be explained solely by “selective upregulation,” with no additional mechanisms of regulation, adaptation, neuroadaptation, homeostasis, or plasticity.

CSF generated neurospheres from adult SVZ precursors as well (Fig

CSF generated neurospheres from adult SVZ precursors as well (Figure 4I). Consistent with these observations and our explant studies,

the Igf1R inhibitor picropodophyllin blocked the formation of spheres in the presence of E17 CSF Anticancer Compound Library research buy (data not shown). Our data suggest that the choroid plexus is the most prominent source of Igf2 in CSF (Figures 3 and S3A). Accordingly, media conditioned with E17 choroid plexus provided enhanced support for neurosphere formation compared to media conditioned with embryonic cortex, adult choroid plexus, or adult brain (Table S3), demonstrating that one or more factors actively secreted from the embryonic choroid plexus, including potentially Igf2, is sufficient for stem cell growth and maintenance. Thus, distinct factors secreted by the choroid plexus into the embryonic Roxadustat concentration CSF, including Igf2, confer E17 CSF with an age-associated advantage to stimulate and maintain

neural stem cell proliferation, and Igf signaling is likely one pathway that promotes this process. Mouse explant experiments confirmed a requirement for Igf signaling in the proliferation of progenitor cells. Mouse embryonic CSF supported the survival and proliferation of mouse cortical progenitors (C57BL/6 explants: 20% ACSF in NBM mean, 7.4 ± 0.2; 20% E16.5 CSF in NBM mean, 14.1 ± 1.4; Mann-Whitney; p < 0.01; n = 3), and purified Igf2 in 20% ACSF in NBM stimulated cortical progenitor proliferation (Figure 5A). When the Igf1R was genetically inactivated in cortical progenitors (Igf1RloxP/loxP/NestinCre+/−) ( Liu et al., 2009), wild-type CSF no longer stimulated cortical progenitor proliferation (ACSF, 17.6 ± 2.9; E16.5 CSF, 16.4 ± 3.0; Mann-Whitney; N.S.; n = 3). Importantly, CSF obtained from Igf2−/− mice failed to stimulate progenitor proliferation in wild-type second explants compared to control ( Figure 5B), suggesting that Igf2 in its native CSF environment stimulates proliferation of progenitor cells during cerebral cortical development. As expected for the roles we have shown for Igf2 in regulating proliferation, we found that Igf2-deficiency reduced brain size ( Figure 5C).

Igf2−/− brain weight decreased by 24% at P8 compared to controls ( Figure 5D). Accordingly, the overall cortical perimeter and surface area were reduced in Igf2−/− brains compared to controls as well ( Figures 5E–5G). Profound defects in somatic size couple to brain size ( Purves, 1988). As previously reported ( DeChiara et al., 1991 and Baker et al., 1993), Igf2−/− body weight was reduced compared to control (mean body weight (g) at P8: Igf2+/+, 5.6 ± 0.01; Igf2−/−, 2.8 ± 0.1; Mann-Whitney; p < 0.0001; n = 11), suggesting that Igf2 may be a secreted factor that scales brain size to body size. Consistent with the mouse CSF Igf2 expression pattern that is significantly increased during later embryonic development ( Figure S3B), blunting Igf2 expression markedly reduced the proliferating progenitor cells at E16.

, 2011) To measure the relative expression of ghsr1a and drd2 mR

, 2011). To measure the relative expression of ghsr1a and drd2 mRNA was isolated from different regions of the mouse brain. RT-PCR shows ghsr1a expression is most abundant in hypothalamus compared to striatum and

hippocampus and that drd2 is expressed mainly in the striatum with lesser amounts in the hypothalamus ( Figure 1A). Immunofluorescence on brain slices from Ghsr-IRES-tauGFP mice ( Jiang et al., 2006) show colocalization of DRD2 and GFP in subsets of neurons with the most abundant coexpression in the hypothalamus ( Figure 1B). The specificity of the DRD2 monoclonal antibody used for immunofluorescence studies was rigorously tested ( Figures S1A–S1D available online). Importantly, DRD2 immunofluorescence was observed in brain slices from drd2+/+, but not drd2−/− Forskolin clinical trial CHIR-99021 manufacturer mice. To investigate whether neuronal cells that coexpress GHSR1a and DRD2 are characterized by modification of signal transduction, we selected the SH-SY5Y neuroblastoma cell line that expresses DRD2 endogenously and generated SH-SY5Y cells that stably express GHSR1a (SH-GHSR1a). In SH-SY5Y parental cells, DRD2 activation by the selective DRD2 agonist, quinpirole, causes coupling to Gαi without inducing release of intracellular Ca2+, whereas quinpirole treatment of SH-GHSR1a cells produces dose dependent

rapid transient Ca2+ signals reaching a maximum by 20 s (Figures 2A and 2B EC50 = 32.76 ± 3.4 nM). Attenuation of the Ca2+ signal by the DRD2 antagonist raclopride confirms DRD2 specificity (Figure 2C) and attenuation by the GHSR1a antagonist/inverse agonist L-765,867, Subst P derivative (Holst et al., 2004 and Smith et al., 1996). Since GHSR1a and DRD2 colocalize in the hypothalamus (Figure 1B), we prepared primary cultures of hypothalamic neurons. Treatment of the cultured neurons induces rapid Ca2+ transients (Figure 2D, upper panel). After washing to remove quinporole,

ghrelin treatment produces an immediate Ca2+ response (Figure 2D, lower panel). These results are consistent with coexpression of GHSR1a and DRD2 in hypothalamic neurons. To study GHSR1a and DRD2 interactions in a system where we could control the relative concentrations of GHSR1a and DRD2, we performed Ca2+ mobilization assays in HEK293 cells Phosphoprotein phosphatase stably expressing the bioluminescent calcium sensor aequorin (Button and Brownstein, 1993). When DRD2 is expressed alone dopamine does not induce Ca2+ mobilization, but when GHSR1a is coexpressed dopamine induces dose-dependent rapid Ca2+ transients with a maximal response at 15–20 s (Figures 3A and 3B, EC50 = 41.88 ± 1.12 nM). To determine GHSR1a specificity, the closely related motilin receptor that also couples to Gαq/11 (Feighner et al., 1999) was coexpressed with DRD2; in this context, dopamine treatment does not induce a Ca2+ response (Figure 3B).

After a delay period, the central fixation point turned off (go s

After a delay period, the central fixation point turned off (go signal), and the monkeys had to make a saccade to the remembered target position. Within one block of 24 trials, saccades to one position were associated with a large amount of liquid reward (large-reward trials) while check details saccades to the other position were associated with a small amount of reward (small-reward trials). In the next block of 24 trials the position-reward contingency was reversed without external instructions. Animals reliably adjusted their saccade performance along with the position-reward contingency reversals: saccades to the large-reward position had higher velocities and shorter latencies and saccades

to the small-reward position had lower velocities and longer latencies (Figures 5A and 5B). This indicates that the animals’

performance was modulated by the expected reward size. While the monkey was performing the reward-biased memory-guided saccade task, we recorded from single neurons in and around the VP which was defined as the structure below the anterior commissure (AC) and above the substantia innominata (Figures 1C and 1D), following Haber et al. (1993). The structure above the AC was defined as part of the external segment of the globus pallidus (GPe). Posterior to these regions are the main bodies of the GPe and the internal segment of the globus pallidus (GPi) (see Figure 6A). This anatomical see more classification was roughly correlated with the variation of single neuronal activity physiologically and click here functionally. In particular, reward-related neurons were found predominantly in the VP, less frequently in the GPe dorsal to the AC, and rarely in the GPe-GPi

posterior to the AC. In this study, we focused on the single neuronal activity recorded in the VP. We tested 190 neurons in the VP using the reward-biased memory-guided saccade task. Among them, 118 neurons (32 in monkey P and 86 in monkey H) (62%) showed task-related modulations. In addition to the neurons formally tested, we encountered 73 neurons which were judged to be unrelated to the task and thus were not tested formally. The average spontaneous firing rate of the task-related VP neurons was 26.6 ± 14.8 spikes/s. The average spike duration of VP neurons was 0.82 ± 0.12 ms. Figure 2 shows two examples of single VP neurons recorded in the reward-biased saccades. As shown in the raster display, both VP neurons changed their activity completely depending on the expected outcome (large or small reward), for both ipsiversive and contraversive saccades. The first VP neuron increased its activity after the onset of fixation point (Figure 2A). This neuron’s activity further increased after the appearance of the target cue indicating the delivery of large reward (large-reward cue, red), but decreased after the appearance of the cue indicating small reward (small-reward cue, blue).