, 2009), which may disassemble the spine’s actin skeleton through

, 2009), which may disassemble the spine’s actin skeleton through phosphorylation of MARCKS (Calabrese and Halpain, 2005). Lead poisoning, which potently activates PKC signaling (Markovac and Goldstein, 1988), may also cause PFC gray matter

loss through this mechanism (Cecil et al., 2008; Hains et al., 2009). Interestingly, traumatic head injury activates stress signaling pathways in surrounding tissue, suggesting universal detrimental actions (Kobori et al., 2006). Stress-induced architectural changes in PFC neurons are reversible in young rats but not in aged rats DAPT order (Bloss et al., 2011). Thus, environmental or genetic insults that disinhibit stress signaling pathways in aging or in mental illness can readily disrupt the precise regulation needed for the integrity of PFC circuits and healthy cognitive function. PFC cognitive functions decline with advancing age, beginning in middle age, in both humans (e.g., Davis et al., 1990; Gazzaley and D’Esposito, 2007) and monkeys (Moore et al., 2006; Rapp and Amaral, 1989). Impaired dlPFC function appears to arise in part from dysregulation RO4929097 cost of DNC signaling with advancing age (Figure 7). It is important to understand these changes, as loss of PFC function is particularly problematic in the Information Age when top-down executive abilities are essential to maintain challenging careers and to manage even basic activities, such as health

care and finances. Age-related vulnerabilities in the association most cortices may also contribute to vulnerability to neurodegeneration, as these are the neurons that are afflicted earliest and most severely in AD (Bussière

et al., 2003). Neurobiological studies of aged rhesus monkeys have illuminated much of the normal aging process, as these animals do not have incipient AD, yet have well-developed association cortices. Ultrastructural studies of the dlPFC have shown large reductions in the numbers of layer III synapses with advancing age, and the loss of synapses correlates with cognitive deficits (Peters et al., 2008). Spine loss particularly afflicts the long, thin spines (Dumitriu et al., 2010; Figure 7A), which are the spines enriched in Ca+2-cAMP signaling proteins (Paspalas et al., 2012). Recent physiological studies have shown marked, age-related reductions in the persistent firing of delay cells, with reductions already evident in middle age (Figure 7B; Wang et al., 2011). In contrast, the firing patterns of sensory neurons (e.g., cue cells) remain intact with advancing age (Wang et al., 2011). Although some of the loss of persistent neuronal firing during working memory likely arises from synapse loss in the recurrent excitatory microcircuits needed to maintain firing throughout the delay period, some of the physiological vulnerability arises from a dysregulated neurochemical environment in remaining spines (Figure 7D). Thus, firing is restored by inhibiting cAMP signaling (e.g.

Other organs, like the liver, heart, and kidney, show similar-mag

Other organs, like the liver, heart, and kidney, show similar-magnitude differences between the sexes, though they are much less studied than the brain. Beyond these global differences, sex differences in specific brain structures have been more difficult to verify. One widely publicized notion is that the corpus

callosum is proportionally larger in female brains. It began with a tiny postmortem study (DeLacoste-Utamsing and Holloway, 1982) showing a statistically marginal effect, which was nonetheless published in Science and made famous by TIME Magazine, Newsweek, and other popular media. Though thoroughly challenged by a meta-analysis of 49 Tyrosine Kinase Inhibitor Library cell assay studies, which collectively showed no significant sex difference in corpus callosum volume or splenial shape ( Bishop and Wahlsten, 1997), the claim lives on among sex difference entrepreneurs like Michael Gurian (see also http://www.girlslearndifferently.com), often as an explanation for females’ mythically superior “multitasking” abilities. Similarly, the planum temporale, a structure involved in receptive language, is often claimed to be more Selumetinib cell line symmetrical between left and right sides of the brain in females

as compared to males, when in fact, meta-analysis of 13 studies found no significant sex difference in its symmetry ( Sommer et al., 2008). Moving on to more reliable differences, sexual dimorphism in the third interstitial nucleus of the anterior hypothalamus (INAH3) has now been confirmed by four different laboratories (Garcia-Falgueras and Swaab, 2008), although the function of this tiny (0.1 mm3) structure, visible only in postmortem tissue, remains unclear. Much more data are available for structures clearly visible by MRI, but surprisingly few findings have been convincingly replicated thus far. Structures that do seem to exhibit reliable volumetric sex

differences (at least during certain developmental ages) include the amygdala, caudate, and portions of the orbitofrontal cortex, although a full review of these complex findings is beyond the scope of this article. Data acquired by fMRI are equally voluminous, but very few sex differences in brain function or connectivity have been confirmed through systematic review. An early claim—that in processing language, second men are left lateralized whereas women exhibit more symmetrical activation of left and right hemispheres—has been largely refuted through meta-analysis (Sommer et al., 2008). However, because the early finding received high-profile coverage in The New York Times, Newsweek, and other media, the claim continues to percolate in popular writings, such as a website promoting all-girls boarding schools that states, “Men tend to use only one brain hemisphere at a time, but women employ ‘whole brain’ thinking” (http://www.girlslearndifferently.com).

g , Figure 1B) A sizeable fraction of cells however showed a com

g., Figure 1B). A sizeable fraction of cells however showed a combinatorial coding of both the attended location and the bar release. Some of the cells, like that shown in Figure 4C, Screening Library cell assay responded selectively if the “E” was

in their receptive field and instructed release of the left bar; other cells had the complementary preference, responding best if the “E” was in their receptive field and instructed release of the left bar (not shown). These manual modulations were not free-standing limb motor responses but modulatory effects on visual selection (i.e., the effects were not seen if a distractor appeared in the receptive field; Figure 4C, right), a conclusion consistent with the later finding that reversible inactivation produced visual but not skeletal motor defects ( Balan and Gottlieb, 2009). These findings are difficult to explain in a purely visual framework that casts target selection as a disembodied bias term (Figure 1B). They are also puzzling in an action based framework that asks whether parietal areas are involved in skeletal or ocular actions ( Snyder et al., 2000). However, neural responses with combinatorial

(mixed) properties are hallmarks of goal-directed cognitive control ( Rigotti et al., 2010), and in the context of information selection may embody the bank of knowledge that is necessary for selecting cues. These results therefore raise the important question of how target selection interfaces with frontal processes of executive control Selleckchem BI2536 and with visual learning mechanisms that assign meaning to visual cues ( Albright, 2012; Freedman and Assad, 2011; Carnitine palmitoyltransferase II Mirabella et al., 2007). One important question is what these complex responses imply for the nature of top-down control. Is the attentional feedback from the parietal lobe only carried by neurons with simple spatial responses, consistent with current assumptions that it only carries spatial information (e.g., Figure 1B)? Or, alternatively, does

the top-down feedback carry higher bandwidth information regarding both stimuli and actions, conveyed by neurons with combined responses ( Baluch and Itti, 2011)? A second question concerns the sophistication of the information conveyed by this combinatorial code: does this code reflect only coincidental associations between stimuli and contexts or actions, or do they reflect internal models of multielement tasks? In sum, the preceding discussion has highlighted some of the complexities that can be entailed by a shift of gaze. Far from requiring a mere direct or habitual sensorimotor link, computing an effective scan path for sampling information requires an executive mechanism that infers the relevant steps in an extend task, and uses this inference to determine points of significant uncertainty and sources of information that may reduce that uncertainty.

, 1998) They developed normal baseline receptive field propertie

, 1998). They developed normal baseline receptive field properties in V1, but brief MD had no effect: the critical period of ODP never opened (Hensch et al., 1998). Enhancing inhibition by infusing diazepam

(an agonist of the GABAA receptor that increases inhibitory conductance when GABA binds) into V1 restored ODP. Brief administration Selleckchem Alisertib of diazepam at any age could open a period of susceptibility to the effects of MD in Gad65-knockout mice that was similar in quality and duration to the normal critical period (Fagiolini and Hensch, 2000). Subsequent administrations of diazepam could not open a second critical period. Remarkably, diazepam treatment in wild-type mice at P15, before the normal critical period, could also initiate a single precocious critical period with a similar 2 week duration (Fagiolini and Hensch, 2000). This finding suggests that a transient increase in GABAergic transmission is sufficient to open the critical period using machinery

that is already in place earlier in development. Opening the critical period appears to trigger unknown mechanisms that lead to its permanent closure 2 weeks later. Subsequent studies narrowed the requirement GSI-IX solubility dmso of GABAergic transmission for the opening of the critical period of ODP to the GABAA receptors containing the α1 subunit. Diazepam binds to several GABAA receptor subtypes, including α1, α2, α3, α5, and γ2 (Sieghart, 1995). Using knockin mice with diazepam-insensitive GABAA receptor subunits, Fagiolini et al. (2004) demonstrated that mutant α2 or α3 GABAA receptor subunits, but not α1 subunits, could still produce a precocious critical period, as in wild-type mice, when diazepam was administered. This experiment suggests that inhibitory neurons like the parvalbumin-expressing (PV) basket cells, which make contacts onto GABA receptors containing the α1 subunit, may play a special role in opening MYO10 the critical period,

although it remains possible that inputs onto receptors containing α5 and γ2 subunits may also be necessary. In normal development, the maturation of the underlying inhibitory circuitry appears to be important for opening the critical period. Several molecular factors that regulate the opening of the critical period also regulate the development of inhibitory neurons in V1. Transgenic animals overexpressing brain-derived neurotrophic factor (BDNF) in excitatory neurons had a precocious critical period and accelerated development of high visual acuity; they also had earlier maturation of inhibitory neurons (Hanover et al., 1999 and Huang et al., 1999). Other studies suggest roles for polysialic acid neural cell adhesion molecule (PSA-NCAM), the homeoprotein transcription factor, orthodenticle homolog 2 (Otx2), and IGF-1 in both the opening of the critical period and the maturation of inhibitory innervation, specifically the perisomatic contacts by PV basket cells onto pyramidal cells (Ciucci et al., 2007, Di Cristo et al.

The observed defects in SV endocytosis in syp−/− neurons result i

The observed defects in SV endocytosis in syp−/− neurons result in functional consequences including the pronounced depletion and slower recovery of the recycling

SV pool. The slow time constant of poststimulus endocytosis might Selleck BMN673 seem to be at odds with the rapid divergence of the synaptic depression time course between wild-type and syp−/− neurons during sustained stimulation ( Figures 4A and 4B). Such rapid depression observed in syp−/− neurons prompted us to test the possibility that rapid retrieval, or “kiss and run” endocytosis of vesicles, is affected in the absence of syp. We note that whether kiss-and-run/fast retrieval (within ∼1 s) is a common mode of endocytosis in hippocampal synapses remains the subject of debate ( Balaji et al., 2008, Ertunc et al., 2007, Granseth et al., 2006 and Zhang et al., 2009). We calculated the rate of vesicle retrieval that occurs during stimulation as a fraction of the total recycling pool (as determined by the maximal ΔF values in the Baf traces) ( Figures 2E and 2F). For wild-type neurons, only ∼1.3% of total recycling pool appears to undergo endocytosis within 1 s; this result argues against the notion that the rapid retrieval (i.e., kiss-and-run) predominates during sustained transmission. Hence, these results indicate that the rapid divergence of the synaptic depression time course between wild-type and syp−/− neurons cannot

be attributed to loss of putative rapid endocytosis. An alternative http://www.selleckchem.com/products/Everolimus(RAD001).html explanation for the pronounced synaptic depression in syp−/− neurons is that syp might regulate another relatively rapid step, such as the clearance of vesicle release sites ( Neher, 2010). Interactions between SNARE proteins on vesicular and target membranes need to be disrupted after exocytosis to allow vesicle recycling.

Syp might facilitate second this process by binding to synaptobrevin II and clearing it from active zones. The loss of syp might lead to a “traffic jam” of vesicular components at release sites and thereby contribute to synaptic depression during sustained activity. However, it is not known whether the clearance of release sites is a rate-limiting step in hippocampal synapses. Finally, we note that read-outs from pHluorin imaging experiments and physiological recordings might not be directly comparable with each other due to several technical differences. These include (imaging versus electrophysiology) different methods of stimulating neurons (field stimulation versus local stimulation) and differences in temporal resolution (“s” versus “ms”). Therefore, there are caveats regarding direct comparison of data from these two experimental approaches. We consider the following possibilities regarding how SV endocytosis can be affected in the absence of syp: (1), unitary endocytic events become slower, or (2), number of SVs that can be retrieved at the same time, i.e., “endocytic capacity” is reduced while endocytosis of individual SVs remains unaffected (Balaji et al., 2008).

Detailed information is provided in the Supplemental Information

Detailed information is provided in the Supplemental Information. Detailed information is provided in the Supplemental Information. Detailed information about the methods used for immunoprecipitation, immunoblotting, and subcellular fractionation is provided in the Supplemental Information. Subcellular fractionation was conducted as described previously (Dunah and Standaert, 2001). Total RNA was isolated from mouse hippocampi or cultured hippocampal neurons using an Agilent Total RNA Isolation Mini Kit (Agilent Technologies), and semiquantitative RT-PCR was performed as described in the Supplemental Information. To

determine the stability of RNAs for the NR2B Akt inhibitor and NR2A receptor subunits, hippocampal Vorinostat in vitro neuronal cells at 7 days in vitro (d.i.v.) were incubated with actinomycin D (Act D, 10 μg/ml, Sigma) to inhibit transcription. Samples for RNA analysis were collected

at 0, 6, 12, and 24 hr, respectively. Paraffin sections (5 μm) or microslicer sections (25 μm) through the hippocampus were prepared for immunohistochemistry experiments. Detailed information is presented in the Supplemental Information. Detailed information is provided in the Supplemental Information. To perform time-lapse imaging of NR2B-EGFP and NR2A-EGFP, after 7 days of culture, hippocampal neurons were cotransfected with untagged NR1-1a/NR2B-EGFP or untagged NR1-1a/NR2A-EGFP vectors. Thirty-six to forty-eight hours after transfection, living neurons were observed under an LSM 5 Duo confocal laser-scanning microscope (Carl Zeiss, Germany). Movement of NR2B and NR2A clusters along dendrites was monitored over time, and images were acquired every 4 s. Determination of cluster velocity was performed using high frame-rate acquisition (one frame per second) for

30 s. The path of individual vesicles was traced, Thymidine kinase and distances were evaluated using LSM 5 Duo software. All images were processed using Photoshop 7.0 (Adobe, San Jose, CA) and further edited as a video file using After Effects (Adobe). The analysis and graphical representation were performed using ImageJ and GraphPad Prism (San Diego, CA). To monitor NR2 subunit degradation, hippocampal neurons at 7 d.i.v. were transfected with untagged NR1-1a and NR2A-PA-GFP or NR2B-PA-GFP, and then imaged at 10 d.i.v. Upon photoactivation with 405 nm laser light, fluorescence was observed under 488 nm excitation in NR2A-PA-GFP- or NR2B-PA-GFP-expressing cells. When needed, neurons were pretreated with MG132 (10 μM). Images were captured at 37°C using an LSM5Duo confocal microscope (Carl Zeiss). Each cell was observed for 6 hr. Conditions were kept constant throughout each experiment and between experiments. Protein levels of NR2A or NR2B in neuronal soma and synapses were estimated by normalizing for the intensity of PA-GFP fluorescence. Hippocampal neurons at 6 d.i.v. were cotransfected with untagged NR1-1a and NR2B-EGFP.

We found that like spatial attention, feature attention affects b

We found that like spatial attention, feature attention affects both rates and correlations and that the magnitudes of these effects covary. Although spatial attention increases the firing rates of most neurons (Figure 3A), feature attention can either increase or DAPT molecular weight decrease firing rates (Figure 2). The presence of both positive and negative rate modulations gives us further dynamic range to test the hypothesis that modulations in firing rate correspond to opposite modulations in correlation. In the plot in Figure 3B, we

arbitrarily define positive rate changes as stronger responses when the animal was performing the orientation rather than the spatial frequency change detection task. The plot verifies that, as in spatial attention, pairs of neurons whose firing rates increase with feature attention show decreases in correlation (Figure 3B, top right). Conversely, neurons whose firing rates decreased with feature attention showed increases in correlation (Figure 3B, bottom left). The relationship between modulation of rate and of correlation was quantitatively similar for the two types of attention learn more (Figure 3C). The slopes of the best fit lines relating the change in noise correlation for each pair to their mean modulation of firing rate were statistically indistinguishable for feature attention (slope, −0.0036, 95% confidence interval [CI] −0.0058 to −0.0014;

12,162 same-hemisphere pairs with similar modulation; see Experimental Procedures) and spatial attention (slope −0.0037,

95% CI −0.0049 to −0.0024; 63,656 same-hemisphere pairs with similar modulation). The y-intercepts of the best-fit lines were also indistinguishable from each other and from zero (feature intercept = −0.010 ± 0.025, 95% CI, spatial intercept = −0.001 ± 0.009, 95% CI). In principle, we could have obtained the results in Figures 3A–3C if the Tryptophan synthase rates and correlations of separate populations of cells were modulated by spatial and feature attention. Instead, we found that most cells were modulated to some extent by both types of attention. Figure 3D shows how modulations by spatial and feature were distributed among cells. No separate subpopulations are obvious. Our data suggest that spatial and feature attention affect local populations of cells in similar ways. Both types of attention modulate the firing rates of individual neurons as well as pairwise spike count correlations. The tight link (and inverse relationship) between attentional modulation of rates and correlations suggests that both changes may be mediated by a single mechanism that decreases correlations whenever gains are increased. Like all neuronal and behavioral processes, attention varies from moment to moment. Analyzing attentional fluctuations is revealing for three reasons.

12 M sodium phosphate buffer For

brain sections, mice we

12 M sodium phosphate buffer. For

brain sections, mice were perfused transcardially with 4% paraformaldehyde in 0.1 M phosphate buffer. Preparation of frozen sections and immunofluorescence were performed by standard procedures. Primary cortical BMS-777607 datasheet neurons were fixed with 1.3% glutaraldehyde in 66 mM sodium cacodylate buffer; postfixed in 1% OsO4, 1.5% K4Fe(CN)6, and 0.1 M sodium cacodylate; stained with 0.5% uranyl magnesium acetate; dehydrated; and embedded in EMbed 812 (Electron Microscopy Sciences, Hatfield, PA, USA). Diaphragms and brain tissue of P0 mice were fixed by immersion in 2% paraformaldehyde, 2% glutaraldehyde in 0.1 M cacodylate buffer (pH 7.4). Additional transmission electron microscopy and tomography were performed as previously described (Ferguson et al., 2007 and Hayashi et al., 2008). For electrophysiology,

cortical neurons were plated at a density of 75,000/cm2 and recorded at room temperature (20°C–22°C) between 10 and 14 DIV. For dynamic imaging studies of exo-endocytosis, vGlut1-pHluorin was transfected into cortical neurons 8 days after plating, and imaging was performed 13–25 days after plating. Statistical significance was determined with the use of Student’s t tests or ANOVA, selleck followed by Tukey’s post hoc test; data with p values <0.05 are indicated by an asterisk (∗) in figures. We thank Frank Wilson, Lijuan Liu, Louise Lucast, Livia Tomasini, Kumi Mesaki, and Ricky Kwan for superb technical assistance. We appreciate the contributions of Tim Nottoli (Yale Cancer Center Animal Genomics Shared Resource) toward gene targeting and the support of the Yale Center for Genomics and Proteomics. This

work was supported in part by the G. Harold and Leila Y. Mathers Charitable Foundation, National Institutes of Health grants (R37NS036251, DK45735, and DA018343), the W.M. Keck Foundation, and a NARSAD Distinguished Investigator Award to P.D.C., a pilot grant from the Yale DERC to X.L., grant RR-000592 from the National Center for Research Resources of the National Institutes of Health to J.R. McIntosh, National Institutes of Health grant NS36942 to T.A.R., and a Canadian Institutes of Health Research fellowship to S.M.F. “
“The development of precise patterns of neural connectivity characteristic of the mammalian brain is thought to occur through a combination of molecular Rolziracetam and neuronal activity-dependent mechanisms (Goodman and Shatz, 1993 and Cline, 2003). During late stages of mammalian brain development, sensory-driven neuronal activity profoundly shapes neural circuit structure and function so that manipulating sensory experience (e.g., through monocular deprivation) can produce dramatic shifts in neural response properties and corresponding changes in neural circuits during “critical periods” of development. In contrast, during early stages of brain development, molecular factors directly regulate cell survival, neurite outgrowth, and branch formation.

From 2002 to 2008, we conducted three trials of NVAS VITA I rand

From 2002 to 2008, we conducted three trials of NVAS. VITA I randomized normal birth weight neonates (≥2500 g) 1:1 to 50,000 IU vitamin A or placebo (2002–2004) [1]. VITA II randomized low birth weight neonates (<2500 g) 1:1 to 25,000 IU vitamin A or placebo (2005–2008) [2]. VITA III randomized normal birth weight neonates 1:1:1 to 50,000 IU vitamin Fulvestrant nmr A, 25,000 IU vitamin A or placebo (2004–2007)

[3]. The trials are presented in more detail in Table 1. The Early MV trial enrolled 4.5 months old children from August 2003 to April 2007 as described in detail elsewhere [5]. Children were randomized 1:1:1 to three treatment groups: a standard dose of Edmonston-Zagreb (EZ) MV at 4.5 months of age and at 9 months of age (group A); no vaccine at 4.5 months and EZ MV at 9 months of age (group B); no vaccine at 4.5 months and Schwarz MV at 9 months

of age (group C). All children were enrolled and randomized at 4.5 months of age. It was a condition for entering the trial that the children had received the third dose of DTP (DTP3) at least four weeks before enrollment; HDAC inhibitor hence, children in groups B and C had DTP3 as their most recent vaccination between 4.5 and 8 months of age. Children in groups B and C who received MV at 9 months of age were randomized to an additional MV or no additional MV at 18 months of age. We found no differences between groups B and C, and hence the two groups have been combined [5]. The Ketanserin vitamin A trials had mortality by 12 months of age as main outcome; the early MV trial had mortality by 3 years of age as main outcome. In the present reanalysis we studied the effect of NVAS versus placebo between 4.5 and 8 months of age, when the children had early MV or DTP3 as their most recent vaccine, and from 9 to 17 months, when the children according to the protocol had two doses of MV or one dose of MV as their most recent vaccine. Follow-up was censored at age 18 months when children in the one-dose MV group were randomized to a booster

dose of MV or no booster and many children received booster DTP. The trials were registered at clinicaltrials.gov (VITA I: NCT00168597; VITA II and III: NCT00168610; Early MV trial: NCT00168558). All trials were approved by the Research Coordination and Ethical Committee of the Ministry of Health in Guinea-Bissau and the Danish Central Ethical Committee gave its consultative approval. All analyses were done using Stata 12.1 (StataCorp, College Station, TX). Characteristics at enrollment into the early MV trial were compared using chi-square test (categorical variables), t-test (normally distributed continuous variables), and Kruskall–Wallis test (non-normally distributed continuous variables). We compared mortality rates (MR) between NVAS and placebo recipients within strata of early and no early MV in Cox proportional hazards models with age as the underlying time variable. Hence, age was inherently adjusted for.

Control experiments expressing GFP rather than G-CaMP in PERin ne

Control experiments expressing GFP rather than G-CaMP in PERin neurons showed no fluorescent changes upon movement, showing that responses are not motion artifacts ( Figure 5B). Taken together, these experiments argue that PERin is activated upon movement, likely by mechanosensory inputs from multiple legs. If movement of the legs activates

PERin to inhibit proboscis find more extension, then one prediction would be that removing leg inhibition would promote extension and that this would require PERin. Flies whose legs were either removed (stumps) or immobilized with wax (wax) showed increased spontaneous proboscis extension, demonstrating that leg inputs inhibit extension (Figures 6A and 6B). Extensions were further enhanced in E564-Gal4, UAS-Shits flies, suggesting that tonic activity in PERin or nonleg inputs may also inhibit extension. Importantly, activation of PERin neurons with dTRPA1 in flies with stumps or immobilized legs prevented the increased spontaneous proboscis extension, suggesting that PERin neurons act downstream of leg inputs to inhibit extension ( Figures 6C and 6D). These studies suggest

that PERin neurons function to inhibit extension Vorinostat while the animal is participating in other behaviors, such as locomotion. As PERin promotes behavioral exclusivity by altering the threshold for feeding initiation in response to mechanosensory-driven behaviors, we hypothesized that commitment to one behavior might more generally prevent other behaviors. Because E564-Gal4; UAS-Kir2.1, tub-Gal80ts flies display constitutive proboscis extension, we wondered whether engagement in this behavior might alter the probability of other behaviors. To test this, we monitored the activity of E564-Gal4; UAS-Kir2.1, tub-Gal80ts flies in a closed arena. Control flies, as well as E564-Gal4; UAS-Kir2.1, tub-Gal80ts flies not expressing Kir2.1, showed robust walking activity, whereas flies before expressing Kir2.1 in E564 neurons had greatly reduced activity, with some flies not taking a single step in the 60 s assayed ( Figures 7A and 7B). All flies were able to move when

presented with a startle stimulus. To test whether the movement impairment was a consequence of silencing PERin, we generated mosaic animals in which Kir2.1 and mCD8-GFP were expressed in subsets of E564 neurons, screened for constitutive proboscis extension, and assayed the extenders and nonextenders for movement (Figures 7A and 7B). Flies with extended proboscises displayed impaired locomotion. To ensure that the locomotion defect was a result of inactivating PERin, we screened mosaic animals for locomotor defects and determined the frequency distribution of neural classes in flies with normal locomotion (>250 mm/min traveled) or impaired locomotion (<200 mm/min traveled). PERin was enriched in flies with locomotor defects and no other cell-type correlated with locomotor defects (Figure 7C).