Next, we tested different adaptation protocols to determine the r

Next, we tested different adaptation protocols to determine the role of adaptive gratings in the reversal. Our second adaptation protocol, termed null adaptation protocol, contained 40 s of gratings drifting only in the ND of the cell. This protocol also produced cells whose tuning was either reversed or ambiguous, but more cells remained stable than with the P-N protocol: 22% (4/18 cells) reversed, 22% (4/18 cells) became ambiguous, and 56% (10/18 cells) remained stable (Figures 2D and 2E, left). Grouping data across all cells showed that the null adaptation protocol significantly decreased DSI values (Figure 2E, right; Table S1).

Hence, stimulation in the ND alone suffices in inducing reversal. Our third adaptation protocol, termed preferred-orthogonal (P-O) protocol, contained 40 s of gratings drifting in the PD, followed by 40 s of gratings drifting orthogonal to the P-N axis. This Selleck Volasertib adaptation protocol also caused most cells to lose their original directional preference: 44% (4/9 cells) reversed, 22% (2/9 cells) UMI-77 concentration became ambiguous, and 33% (3/9 cells) remained stable. Once again, the DSI values decreased significantly after this protocol (Figure 2G, right; Table S1). However, surprisingly, the reversed cells exhibited a new PD that was similar to the original ND rather than the direction of the training stimulus (Figures 2F and 2G, left), suggesting that the adaptive stimulus drives reversal but

does not instruct the direction of the reversal. Our fourth protocol, termed counterphase protocol, contained counterphase Metalloexopeptidase gratings in which the gratings did not move but instead switched their colors from black to white in a frequency that was similar to the frequency of the moving gratings (4–8 Hz; Figure 2H). Although the counterphase protocol changed the PD of some DSGCs—25% (3/12 cells) reversed, 17% (2/12 cells) became ambiguous, and 58% (7/12 cells) remained stable (Figure 2I, left)—they did not produce a significant decrease in the DSI across the population (Figure 2I, right;

Table S1). Hence, motion in the adaptive stimuli is not critical for reversal but it increases its probability. As a control for our various protocols, we took a group of cells and performed consecutive DS tests separated by a gray screen that appeared for 5–9 min (comparable to the time between first and second DS tests in the P-N adaptation protocol). The control protocol did not reverse any cell’s PD, but some cells did become ambiguous (36% or 4/11 cells). However, the DSI values in this control group did not change significantly (Figure S2D, right, and Table S1). In addition, we presented the P-N adaptation protocol prior to recording from the cell and found that the majority of the cells (n = 5/8) had a reversed directional preference, indicating that the reversals were not due to the recording itself. We next addressed the issue of why some cells reverse after exposure to a given adaptation protocol while others do not.


“Tobacco smoking, nicotine/tobacco dependence and attentio


“Tobacco smoking, nicotine/tobacco dependence and attention-deficit/hyperactivity disorder (ADHD) frequently co-occur. Persons with ADHD are more likely to become regular smokers (Pomerleau et al., 1995 and Tercyak et al., 2002), begin smoking earlier, smoke more heavily (Kollins et al., 2005 and Lambert and Hartsough, 2000), and may experience greater difficulty when trying to stop smoking (Humfleet et al., 2005 and Covey et al., 2008) compared to persons without ADHD. Nicotine ameliorates inattentiveness and problems in response inhibition (Conners et al., 1996, Levin et al., 1996, Potter and Newhouse, 2004 and Poltavski

and Petros, 2006), which are core symptoms of ADHD. Nicotine can reduce the demonstrated deficits in dopaminergic TGF-beta inhibition function associated with ADHD (Volkow et al., 2007) suggesting LGK-974 a “self-medication” rationale for greater tobacco use among persons with ADHD (Gray and Upadhyaya, 2009). The increased recognition that tobacco use and nicotine dependence are highly prevalent among persons with ADHD (Gray and Upadhyaya, 2009) has spurred investigations into details

of the relationship between those disorders, such as the association between their symptom profiles. The core symptoms of ADHD (inattention, hyperactivity, and impulsiveness; APA, 2000) are conceptually and clinically similar to symptoms of nicotine withdrawal, such as difficulty concentrating, restlessness, and impatience (APA, 2000). A study of adolescent smokers that examined correlations during the non-abstinence phase of a smoking cessation treatment found significant correlations among several of the ADHD and the nicotine withdrawal symptoms (Gray et al., 2010). A 12-day abstinence trial conducted with adult, non-treatment seeking smokers, on the other hand, observed that withdrawal symptoms, which were experienced more severely by smokers with than without ADHD, were unrelated to changes in ADHD symptoms (McClernon et al., 2011). To clarify relationships among smoking-related (i.e., withdrawal symptoms and craving) Megestrol Acetate and ADHD-related symptoms, as well as their relevance to the efficacy of smoking cessation treatment for smokers with ADHD, we conducted secondary

analyses of data from a trial of osmotic-release oral system methylphenidate (OROS-MPH) for smokers with ADHD (Winhusen et al., 2010). The parent trial was a randomized, placebo controlled trial that evaluated if OROS-MPH to treat ADHD, combined with smoking cessation treatment, increases smoking abstinence. The main results showed that OROS-MPH reduced ADHD symptoms but did not improve smoking abstinence rate (Winhusen et al., 2010). Our objectives in the current analysis were: (1) to assess overlap between ADHD symptoms and nicotine withdrawal symptoms and craving; (2) to assess the relationship between craving or withdrawal symptoms and the OROS-MPH effect on ADHD symptoms; (3) to assess the association of ADHD symptoms, craving, and withdrawal symptoms with abstinence.

, 2003; Eichner et al , 2011) L1 and L2 provide inputs to EMDs a

, 2003; Eichner et al., 2011). L1 and L2 provide inputs to EMDs and thus their outputs must represent some of these filtering stages (Rister Docetaxel ic50 et al., 2007; Joesch et al., 2010; Clark et al., 2011). We show that L2 outputs

are strongly shaped by the light distribution across a broad region in space and by contrast polarity. Thus, the kinetics and amplitudes of L2 outputs differ for bright and dark objects of different shapes and sizes. Consequently, probing EMDs with minimal motion cues that differ in contrast and spatial extent could produce different results due to differential input filtering rather than differences in motion detection per se (Hassenstein and Reichardt, 1956; Egelhaaf and Borst, 1992; Eichner et al., 2011; Clark et al., 2011). More generally, spatiotemporal coupling observed in L2 can give rise to speed tuning, differentially regulated for bright and dark objects, and thus affect tuning of downstream EMDs to different speeds or to dark or bright Venetoclax solubility dmso motion cues (Fleet et al., 1985; Fleet and Jepson, 1985; Egelhaaf and Borst, 1989; Srinivasan et al., 1990; Juusola and French, 1997; Zanker et al., 1999). Finally, the surround responses of L2 effectively

convert a contrast increment at one spatial location into depolarizing responses at neighboring locations, providing a route by which increment information could enter a dark edge-detecting pathway, even given downstream half-wave rectification (Clark et al., 2011). Anatomical studies describe a dense network of connections in the lamina (Meinertzhagen and O’Neil, 1991; Rivera-Alba et al., 2011). Here we show how GABAergic circuits within this network shape the functional properties of L2 (Figures 6, 7, and 8). Photoreceptors until receive direct GABAergic input that depends on both GABAARs and GABABRs and shapes the RF surround in L2 (and presumably other LMCs). GABAAR-dependent synapses elsewhere in the circuit relay surround inputs into photoreceptors. A possible surround input is the centrifugal cell, C3, the only cell that is both presynaptic

to photoreceptors and GABAergic (Buchner et al., 1988; Kolodziejczyk et al., 2008; Rivera-Alba et al., 2011). Furthermore, since our genetic manipulation of GABARs affected both L2 cells as well as photoreceptors, we cannot exclude the possibility that receptors on both cells are redundantly required. Thus, the GABAergic centrifugal cell C2, which is presynaptic to L2, could provide these inputs. Additional GABAARs have been identified in L4 and another wide-field tangential cell (Enell et al., 2007; Kolodziejczyk et al., 2008) and could mediate the distal effects of manipulating GABAARs. Modulation of GABAergic signaling in L2 expands the RF center and increases spatial pooling. Such a change in RF shape increases signal-to-noise ratios and occurs under low light level conditions (Dubs et al., 1981; Dubs, 1982).

Tau-mediated synaptic dysfunction is likely

to be a key s

Tau-mediated synaptic dysfunction is likely

to be a key step in the transition from the asymptomatic to symptomatic phase of tauopathies, the former phase characterized by changes in functional and molecular biomarkers, the latter phase characterized by neuronal death and cognitive deficits. Our electrophysiological and immunocytochemical experiments revealed that tau mutation or hyperphosphorylation leads to impaired trafficking or anchoring of AMPARs and NMDARs that precedes spine loss. Understanding the early pathological FG-4592 nmr events preceding synapse loss in neurodegenerative diseases is of critical relevance since synaptic loss correlates more robustly with cognitive deficits in AD than plaque or tangle load (Davies et al., 1987, Terry et al., 1991, Masliah et al., 2001 and Selkoe, 2002). Most excitatory synaptic transmission is mediated through dendritic spines which are principal loci of synaptic plasticity (Hering and Sheng, 2001). Glutamatergic receptors govern the morphological plasticity of spines with NMDARs regulating the formation of new spines and AMPARs stabilizing established spines (Fischer et al., 2000). Interestingly, a recent study demonstrated that tau decreased NMDAR expression in the PSD complex,

though no alterations in NMDAR-mediated currents were detected (Ittner et al., 2010). Our study complements and extends these findings to include tau-induced impairments in synaptic targeting or anchoring of AMPARs and AMPAR-mediated synaptic currents.

check details Dendritic spine stability is compromised by loss of synaptic AMPARs, leading to spine loss (McKinney et al., 1999 and Richards et al., 2005; for review, see McKinney, 2010). Despite the strong correlation between loss of spines and Mannose-binding protein-associated serine protease synaptic AMPARs, the temporal relationship between these two physiological events is not well-defined. It has also been reported that these two cellular processes can be dissociated (Sdrulla and Linden, 2007 and Wang et al., 2007). Although other studies (Shahani et al., 2006 and Tackenberg and Brandt, 2009) have also failed to observe tau-mediated spine loss, the state of synaptic AMPARs was not reported in these papers. Thus, we hypothesize that the absence of spine loss, despite decreased synaptic AMPARs, is due to the very early nature of the tau-induced synaptic impairments. A similar theory was proposed in a study using the APPswe/PS1 dE9 transgenic mouse model of AD (Shemer et al., 2006). The relationship between Aβ-induced early synaptic dysfunction and tau-mediated global disruption of glutamate receptors remains to be determined. A recent report (Zempel et al., 2010) suggests that spines might serve as a site of convergence for amyloid beta and tau.

Integrated genetic-neuroscience studies of interindividual variat

Integrated genetic-neuroscience studies of interindividual variation in the PI3K inhibitor relative abundance and properties of these various cell types

will provide a snapshot of the phenotypic variation within a population and how it relates to the heritable information of the genome. This paper offers a first glance at genes whose patterns of expression vary among individuals. The present study also describes sex differences in the expression of several autosomal genes, as previously observed (Kang et al., 2011). The biology that underlies and results from the gene expression differences between males and females may provide insight into neurodevelopmental disorders that differentially afflict men and women such as autism. Finally, the results call into question the prevailing cytoarchitecture-based hexalaminar nomenclature used for the neocortex. For example, in this study the authors show that what is presently known as layer 4A in primary visual cortex is transcriptionally

far more similar to layer 3 than to other layer 4 sublaminae. Interestingly, www.selleckchem.com/products/Y-27632.html Hassler and Stephan (1966) and subsequently Casagrande and Kaas (1994) arrived to similar conclusions by tracing neuronal connections. If further work demonstrates this clustering is driven by excitatory neurons, a genetically informed reconsideration of laminar nomenclature may be in order. “
“The neocortex is a laminated structure composed of billions of neurons that make synaptic connections with distant and interspersed populations of neurons located both within the neocortex and throughout the central nervous system (CNS). The past two decades have been extremely fruitful in identifying some of the molecular mechanisms regulating the ability of axons to navigate through the CNS and find their target structure. However, less is known about the mechanisms regulating the final choice that neurons have to make within a given target region. There, a daunting all task still awaits the axon: to

make synaptic contacts with a few hundred/thousand neurons among millions of possible postsynaptic targets (Sanes and Zipursky, 2010). This problem of synaptic specificity has received a lot of attention recently and the list of extracellular cues regulating this critical step is rapidly expanding (de Wit et al., 2011 and Shen and Scheiffele, 2010). In this issue of Neuron, the Kriegstein lab expands the portfolio of Shh functions by demonstrating its involvement in the formation of functional synaptic contacts between specific subpopulations of cortical neurons ( Harwell et al., 2012). The Shh pathway plays several critical functions as a patterning cue during early brain development by regulating gene expression, cell-fate specification, as well as neural progenitor proliferation.

For the quantification of migration in MGE explants, the distance

For the quantification of migration in MGE explants, the distance migrated by the 40 furthest BMS-387032 mw cells was measured. For the analysis of interneuron migration in vivo, the number of GFP-expressing cells was quantified in the same region located in the prospective somatosensory cortex for each brain. The area quantified was divided into 10 equal bins and the percentage of cells in each bin was calculated. For GFP and PV analysis at P21, the same region of the somatosensory, motor, and visual areas was quantified in control and mutant brains. Layers were drawn following

nuclear staining. Layers I, II/III, and IV were grouped as supragranular layers, while layers V and VI were grouped as infragranular layers. Cxcr4 fluorescence levels and colocalization of Cxcr4 and WGA was measured using ImageJ software (NIH, http://rsb.info.nih.gov/ij/).

In the first case, stacks of individual cells were taken using a Leica Confocal microscope (MCSII) every 1μm. Fluorescence intensity was measured in every stack of cell and the total fluorescence was calculated as the sum of the fluorescence of all stacks of the cell. For Cxcr4/WGA measurements, a single confocal plane was obtained per cell and the Mander’s coefficient was used to calculate colocalization. For statistical analyses, normality and variance tests were first applied to all experimental data. When data followed a normal find more distribution, paired comparisons were analyzed with t test, while multiple comparisons were analyzed using either ANOVA with post hoc Bonferroni correction (equal variances) or the Welch test with post hoc Games-Howell (different variances). A χ2 test was applied to analyze the distribution of cells in either bins or layers. We thank A. Casillas, T. Gil, M. Pérez, K. Schäfer, A. Sorgenfrei, and H. Stadler for technical assistance; K. Campbell (Dlx5/6-Cre-IRES-Gfp) and N. Kesaris (Lhx6-Cre) for below mouse strains; E. Arenas, F. Arenzana-Seisdedos, F. Guillemot, M. Penfold, M. Thelen, and V. Pachnis for plasmids and reagents; and V. Borrell for critically reading early

versions of this paper. We are also thankful to members of the Marín, Rico, and Borrell labs for helpful discussions and comments. J.A.S-A. was supported by a fellowship from the FPU program of the Spanish Ministry of Science and Innovation (MICINN). This work was supported by grants from Spanish MICINN SAF2008-00770 and CONSOLIDER CSD2007-00023, and the EURYI scheme award (see www.esf.org/euryi) (to O.M), and by Federal State Sachsen-Anhalt with the European Fund For Regional Development (EFRE 2007-2013) and Deutsche Forschungsgemeinschaft (DFG) grant STU295/5-1 (to R.S). “
“Nervous system development and function is dependent upon a variety of soluble and membrane bound trophic stimuli, many of which act through receptor tyrosine kinases (RTKs).

Perhaps the interface offers an opportunity to learn about the lo

Perhaps the interface offers an opportunity to learn about the logic of two behaviorally related

systems for the price of one interrogation. We thank Jay Bikoff, Andrew Fink, Samaher Fageiry, and Mark Churchland for discussions that helped shape the opinions in this essay. T.M.J. thanks Liz Wright and Rob Brownstone for providing a dose of Halifax serenity needed for the completion of this essay. This work was supported by NINDS and ProjectALS. T.M.J. is an HHMI investigator; both A.M. and E.A. are Howard Hughes Medical Institute Fellows of The Helen Hay Whitney Foundation. “
“Although psychological experts tell us to avoid becoming too compartmentalized in our thinking, compartmentalization is a key feature of neurons. Generation of an axonal and LBH589 in vivo a somatodendritic domain is a prerequisite for the directed flow of information in the nervous system. Therefore, the establishment of the complex neuronal morphology with one axon and several dendrites is a critical step during neuronal differentiation. The underlying mechanisms that regulate the formation of neuronal

polarity are currently under intense investigation. In culture, hippocampal neurons start off as round, unpolarized cells that transform into a multipolar cell with several short neurites that all have the potential Ibrutinib molecular weight to become an axon. Only one of these neurites will grow quickly and turn into an axon while the other neurites only start to grow later and become dendrites. Stable microtubules in the axon shaft and a dynamic actin network in the axon growth cone are instructive for axon growth (Stiess and Bradke, 2010). However, so far, it has remained Ribonucleotide reductase unclear how a neuron coordinates intracellular changes that could lead to the growth of the

axon and the simultaneous halt of the other neurites. The reported restriction of growth permissive proteins, including the partitioning-defective (Par) proteins Par3 and Par6 (Shi et al., 2003) and Rap1B (Schwamborn et al., 2007), to the nascent axon may present a hallmark of neuronal polarity. The asymmetric localization of axon determinants can be achieved by transport into one process (Bradke and Dotti, 1997), e.g., along selectively stabilized microtubules in the growing axon (Stiess and Bradke, 2010). In addition, the selective stabilization of the proteins in the future axon might lead to the asymmetric localization of polarizing proteins. Indeed, the small GTPase Rap1B in its inactive form becomes ubiquitinated and thus targeted for proteasomal degradation by the E3 Ligase Smurf2 in the minor neurites (Schwamborn et al., 2007). The resulting axonal localization specifies the future axon and is required for neuronal polarization. In this issue of Neuron, Cheng et al.

The weaker “U”-shaped relationship that appears instead in Figure

The weaker “U”-shaped relationship that appears instead in Figure 5C (open symbols) would not promote spurious MT-pursuit correlations. Therefore, the small eye movements of fixation do not cause the MT-pursuit correlations in our data. The eye speed at the initiation of pursuit shows “endpoint” variance of about 15% of the mean speed (Osborne et al., 2005). From the perspective of sensory processing, the endpoint variance could arise from correlated noise in the responses of MT neurons (Huang and Lisberger, 2009), or from downstream sources including noise added by the population decoders (e.g., Shadlen et al., 1996). These Selleckchem Ulixertinib two potential sources trade

off in a potentially informative way. Larger, structured neuron-neuron correlations in MT cause larger MT-pursuit correlations ( Schoppik et al., 2008) and larger endpoint variance ( Huang and Lisberger, 2009). Larger downstream noise causes smaller MT-pursuit correlations and larger endpoint variance ( Medina and Lisberger, 2007). Thus, we might further our understanding of the source(s) of endpoint variation in pursuit initiation if we could quantify the amount of noise reduction between the responses of MT neurons and the motor output. Given the large number of MT neurons that probably contribute to pursuit, one might expect noise reduction to be excellent. However,

either sensory noise or downstream noise would limit noise reduction. To www.selleckchem.com/products/Fulvestrant.html compare neural to behavioral noise, we transformed eye speed in each behavioral trial into the same units as the firing

rate of the MT neuron recorded at the same time. First, we converted eye speed 100 ms after the onset of pursuit ( E˙i(100)) to an estimate of target speed ( T˙i) as: equation(Equation 11) T˙i=E˙i(100)〈E˙i(100)〉T Equation 11 normalizes the eye velocity from each trial so that the mean normalized eye velocity was equal to the actual target velocity. The dots over the symbols indicate speed, T˙ and E˙ refer to the target and the eye, i indexes the PD184352 (CI-1040) trials, and the denominator is the mean across all trials. We performed the analysis for eye velocity at t = 100 ms because this time marks the end of the open-loop period when pursuit is driven purely by the target motion present before the onset of pursuit. Second, we converted the estimate of target speed for each trial to the units of spikes/s by projecting through the mean speed tuning curve for the neuron under study, as illustrated in Figure 6A. Finally, we characterized noise reduction by expressing the variance of eye velocity in units of spikes/s as a percentage of the variance of actual firing rate and plotted the result as a function of preferred speed normalized to target speed ( Figure 6B). The shape of the mean tuning curves leads to the “M” shaped functions in Figure 6B, for both the data (symbols) and the model MT neurons (red and blue traces).

, 2001) The stable value information in the caudate tail may be

, 2001). The stable value information in the caudate tail may be transmitted to the superior colliculus through the substantia nigra pars reticulata so that monkeys make saccades preferentially to high-valued objects (Yasuda et al., 2012). Although these studies individually provide important data, it has been difficult to reach a unified view on basal ganglia functions. Our recording and inactivation experiments on the primate caudate head and tail provide insights in understanding how the basal ganglia normally control behavior in multiple but integrative ways and how behavior can be disrupted in multiple ways in basal ganglia disorders. Two adult male rhesus monkeys (Macaca mulatta, 8–9 kg) were used for the experiments.

All animal care and experimental procedures were approved BMS354825 by the National Eye Institute Animal Care and Use Committee and complied with the Public Health Service Policy on the humane care and use of laboratory animals. We implanted

a plastic head holder and a recording chamber to the skull under general anesthesia and sterile surgical conditions. The chamber was tilted laterally by 25° and was aimed at the caudate head, body, and tail. Two search coils were surgically implanted under the conjunctiva of the eyes to record eye movements. After the monkeys fully recovered from surgery, we started training them with flexible and stable value procedures. While the monkey was performing a task, we recorded the activity of single neurons in different subregions in the caudate nucleus using learn more conventional methods. The recording sites were determined with 1 mm spacing grid system, with the aid of MR images (4.7T, Bruker) obtained along the direction of the recording chamber. Single-unit recording was performed using glass-coated electrodes (Alpha-Omega). The electrode was inserted into the brain through a stainless-steel about guide tube and advanced by an oil-driven micromanipulator (MO-97A, Narishige). The electric signal from the electrode was amplified with a band-pass filter (2 Hz–10 kHz; BAK) and collected at 1 kHz. Neural spikes were isolated online using a custom voltage-time window discrimination software (MEX, LSR/NEI/NIH).

Behavioral tasks were controlled by a QNX-based real-time experimentation data acquisition system (REX, Laboratory of Sensorimotor Research, National Eye Institute, National Institutes of Health [LSR/NEI/NIH]). The monkey sat in a primate chair, facing a frontoparallel screen 33 cm from the monkey’s eyes in a sound-attenuated and electrically shielded room. Visual stimuli generated by an active matrix liquid crystal display projector (PJ550, ViewSonic) were rear projected on the screen. We created the visual stimuli using fractal geometry. Their sizes were approximately 8° × 8°. This procedure allowed us to examine behavioral and neuronal encoding of flexible object values as they were being updated in blocks of trials (Figure 1A and Figure S1A).

These data are consistent with an important and rapid recruitment

These data are consistent with an important and rapid recruitment of inhibition during active touch, which is likely to impose the hyperpolarized reversal potentials for the touch-evoked PSPs found in excitatory layer 2/3 neurons. Internal cortical dynamics and precontact membrane potential therefore play a major role Talazoparib concentration in governing the trial-by-trial touch-evoked PSP, but one would also expect important contributions to the response variability mediated by differences in kinetics during different whisker-object

contacts. However, in agreement with previous local field potential measurements (Hentschke et al., 2006), in most neurons we found that the amplitude of the touch-evoked PSP was modulated neither by precontact velocity nor contact duration (Table S2). However, we did find a strong influence of the intercontact interval (ICI) upon the touch response. The first whisker-object contact in a touch sequence generally evoked the largest membrane potential depolarization (Figures 6A and 6E).

Subsequent touches on average evoked smaller depolarizations, indicating a dependence upon the recent history of C2 whisker-related touches (Figure 6I). Averaging active touch responses for different intercontact interval ranges revealed a decrease in response Nutlin-3 research buy amplitude as the ICI becomes smaller (Figures 6B and 6F). In order to evaluate further the impact of ICI, we plotted the amplitude of the touch response as Thiamine-diphosphate kinase a function of the preceding intercontact interval (Figures 6C and 6G). Spearman’s rank test revealed a significant modulation of response amplitude by ICI in 13/17 neurons. For those neurons, the time course of the recovery of the touch response was quantified by an exponential fit, yielding the intercontact interval time-constant for the half-maximal response, which we denote as ICI50. The time course of suppression

of the touch response varied strongly across the population of recorded neurons with a mean ICI50 of 87 ± 61 ms (median 63 ms; range 14 to 194 ms). Across our population of recorded C2 column layer 2/3 neurons, the mean amplitude of the touch response for long preceding ICI (>500 ms) was 8.3 ± 4.1 mV and decreased significantly to 3.1 ± 2.2 mV for short ICI (10–40 ms) (Figure 6J). The major impact of ICI on response variability could be seen by the linear relationship between the coefficient of variation of the response and the ICI50 (r = 0.94) (Figure 6K). Since the duration of the touch response was often longer than the ICI, consecutive touch responses also in many cases began from a more depolarized baseline Vm. Indeed, plotting the precontact Vm against the preceding ICI indicates a near parallel increase in precontact Vm and decrease in response amplitude at shorter ICI (Figures 6C and 6G).