The number

The number Roxadustat nmr of positive pixels and positive clusters (groups of adjacent positive pixels) within the outline was counted using ImageJ. To normalize

for variation in size of neurons, we divided the numbers of pixels and clusters by the outline perimeter. Data are presented as means ± SEM and were analyzed using ANOVAs repeated-measures and two-tailed t test (unpaired or paired) for normally distributed variables to evaluate statistical significance with p < 0.05 as level of statistical significance. See Table S2 for the average number of analyzed cells per mouse for each perisomatic marker and Table S3 for detailed statistical results. We thank K. Kan and M. Parakala for technical assistance and M. Mayford for providing

TetTag mice. We thank J. Aggleton, J. Ainsley, L. Drane, L. Feig, M. Jacob, K. GDC-0199 research buy Mackie, E. Perisse, and S. Waddell for critical reading of the manuscript. This work was supported by an NIH Director’s New Innovator Award (L.G.R.; DP2 OD006446), a Fyssen Foundation Postdoctoral Fellowship, a Bettencourt-Schueller award, and a Philippe Foundation Award (S.T.), a Sackler Dean’s Graduate Fellowship (J.S.), the Synapse Neurobiology Training Program (J.S.; T32 NS061764; PI: K. Dunlap and M. Jacob), the Tufts Center for Neuroscience Research (P30 NS047243; PI: R. Jackson), and DA011322 (PI: K. Mackie). “
“Neurons communicate with each other in dynamically modulated circuits. Functional connectivity, a measure of interactions between neurons in these circuits, can change gradually during learning (McIntosh and Gonzalez-Lima, 1998) and formation of long-term memories, or it can change rapidly, depending on behavioral context and cognitive demands. While the mechanisms underlying long-term network plasticity have been extensively documented, those underlying rapid modulation of functional connectivity remain largely unknown. At the network level, functional connectivity is affected by up-down and oscillatory states of the neural network (Gray et al., 1989). Cortical inhibition plays a key role in this process

(Cardin et al., 2009, Sohal et al., 2009 and Womelsdorf et al., 2007). Parvalbumin-positive (PV+) interneurons, which make up more than half of the inhibitory neurons in the click here cortex (Celio, 1986), are particularly important as they provide strong feedforward and feedback inhibition that can synchronize the cortical network (Cardin et al., 2009, Fuchs et al., 2007, Isaacson and Scanziani, 2011 and Sohal et al., 2009). Their precise influence on cortical networks during sensory processing, however, remains unclear. In particular, it is unknown how PV+ neurons may differentially modulate responses in different layers of the neocortex and how the anatomical organization of the cortex affects the flow of information through these networks.

They found that mEPSC amplitudes were unchanged at 6 and 18 hr po

They found that mEPSC amplitudes were unchanged at 6 and 18 hr postlesion but then increased at

24 and 48 hr, closely matching the time course of activity rate homeostasis. Because spine size is correlated with synaptic strength, and changes in a predictable manner when circuits are weakened or strengthened in response to MD in vivo (Hofer et al., 2009), Keck et al. (2013) hypothesized that in vivo scaling of synaptic strengths should have a structural correlate in altered dendritic spine size. Remarkably, they indeed found that spine size on L5 pyramidal neurons increased 24 hr after the retinal lesion and was maintained at 48 hr, thus following the same time course as the changes in DAPT price mEPSC amplitude and cortical activity in vivo. Altogether, these data and those obtained by Hengen et al. (2013) are consistent with the hypothesis that synaptic scaling could underlie homeostatic adjustments in neocortical firing rates in vivo. The studies by Hengen et al. (2013) and Keck et al. (2013) provide much anticipated evidence supporting that neuronal activity levels are homeostatically regulated in the neocortex in vivo. While both studies report an initial drop in activity levels in response to sensory deprivation, followed by a subsequent rebound, the time courses of the

two observations are dramatically different. Interestingly, the rapid sensory deprivation induced drop in overall activity levels observed by Keck et al. (2013) recovered to control levels within 24 hr, which is when Hengen et al. (2013)

obtained their first measurements selleck chemicals llc also showing baseline firing rates in excitatory neurons. Discrepancies between the two studies are evident only at 48 hr, when Hengen et al. (2013) see significant depression of firing rates in excitatory neurons, whereas Keck et al. (2013) observe baseline activity Tryptophan synthase levels. Most likely, differences are due to the widely diverse experimental conditions in the two studies—including deprivation protocols (monocular lid suture versus binocular retinal lesion), species (rat versus mouse), and ages (juvenile versus adult; Figure 1). Future experiments utilizing similar paradigms, while independently varying the individual parameters, will shed light on the mechanisms and origins of these differences. Several testable predictions arise from these studies and lead to exciting new avenues of research. While these studies support that synaptic scaling could be responsible for homeostatic regulation of firing rates in the neocortex, they do not exclude that alternative mechanisms of synaptic plasticity, such as plasticity of intrinsic excitability, anti-Hebbian mechanisms, or Hebbian modifications of excitatory or inhibitory synapses, are also at play. One prediction is that a homeostatic set point should operate bidirectionally; and consequently, enhanced firing rates due to sensory overstimulation should be homeostatically downregulated.

This leads us to believe that significant confounding due to prio

This leads us to believe that significant confounding due to prior infection with, and immune response to, non-vaccine types to be highly unlikely. Our assessment of non-specific interference using sera from HPV-naïve infants resulted in a pseudovirus neutralization assay specificity of around 99–100%. As the sera used for this study were collected within six months of the third vaccine dose and given the apparent improved immunogenicity within

this age group [31], the titers of cross-neutralizing antibodies reported here are likely to represent peak levels. Type-specific neutralizing antibodies appear to wane quite BI 6727 price quickly following vaccination to plateau several fold lower than their peak level [35] and this is likely to be true also for cross-neutralizing antibodies. We did not have repeat samples or a sufficient range in collection times to assess changes in neutralizing antibody titers over time. The detection of cross-neutralizing antibodies in vaccine sera per se does not, of course, provide sufficient evidence for antibodies being mechanistically associated with cross-protection. Furthermore,

type-specific antibody titers in genital secretions are orders Bcl-2 inhibitor of magnitude lower than those found in the periphery [12] and it is unclear whether these very low levels of cross-neutralizing antibodies found in the periphery would be sufficient to protect at the site of infection in the absence of other immune effectors [36] and [37]. However, the coincidence of the rank order of HPV types recognized by vaccinee sera in this and other studies [20] and the apparent hierarchy of protected HPV types suggested from efficacy studies [4], [16] and [17] is intriguing. Defining the mechanism(s) of cross-protection will be important to monitor vaccine effectiveness on both a population and individual level. These data may be helpful to parameterize epidemiological models to predict the impact of the current HPV vaccines on the population and to inform the development of second generation HPV vaccines. This study was given a favorable ethical opinion by the Tameside & Glossop

Local Research Ethics Committee, Manchester, UK (REC reference number 09/H1013/33). This work was supported by the UK Medical Research Council (grant number G0701217). We thank Dr. Rosemary McCann (Greater Manchester Calpain Health Protection Unit, U.K.), Dr. Ray Borrow and Elaine Stanford (Vaccine Evaluation/Seroepidemiology Unit, Manchester Royal Infirmary, U.K.) for coordinating the collection of the serum samples used in this study and Prof. Elizabeth Miller and Liz Sheasby (National Vaccine Evaluation Consortium, U.K.) for providing anonymized infant, HPV-naïve sera. We are grateful to Tom Nichols for helpful discussions on statistical analyses. We are indebted to Prof. John T. Schiller and Dr. Chris Buck (National Cancer Institute, Bethesda, U.S.A.) and Dr. H. Faust and Prof. J.

In addition, other LDL receptor family members have been implicat

In addition, other LDL receptor family members have been implicated in AD, owing to their roles in modulating the intracellular trafficking and processing www.selleckchem.com/products/Romidepsin-FK228.html of the amyloid precursor protein ( Cam and Bu, 2006). The effects of astrocyte-derived apoE in the brain are a point of ongoing study, and it remains to be described whether astrocyte-derived apoE impacts neuronal health and pathology differently

from the apoE that is synthesized within neurons. The remainder of this review will describe how induction of neuronal apoE (apoE4 > apoE3 > apoE2) in response to injury sets the stage for neuropathology and subsequent neurodegeneration. The apoE hypothesis posits that apoE genotype sets the stage for neuropathology in an isoform-dependent manner (apoE4 > apoE3 > apoE2), and “second hits” that directly induce neuronal injury or stress initiate a pathological response to injury when apoE4 is synthesized in neurons (Huang, 2010; Huang and Mucke, 2012; Mahley and Huang, 2012; Mahley et al., 2006). With respect to AD, these second hits could include aging, ischemia, trauma, inflammation, oxidative stress, or toxins like the Aβ peptide and its different assemblies. TBI causes direct damage to neurons, whereas following stroke the second

hit may be ischemia. Given the nature of many neurological diseases, where multiple FG-4592 research buy pathologies occur over a protracted period, the possibility for second hits is very high. Other genetic disorders and metabolic disturbances, such as diabetes, can also be injurious factors that contribute to apoE4’s neurotoxic effects. In response to injury, neurons induce the synthesis of apoE, presumably to participate in lipid transport and redistribution Rolziracetam for membrane repair and remodeling. However, because of varying degrees of structural instability and tendency to assume domain interaction across the apoE isoforms (which we discuss

in greater detail below), apoE can be recognized as structurally abnormal by neurons and undergo proteolytic cleavage (apoE4 > apoE3 > apoE2). The neurotoxic fragments that are generated cause mitochondrial dysfunction and cytoskeletal alterations. In the sections to follow, we describe in more detail the data supporting the apoE hypothesis (Figure 1). As mentioned previously, in the brain apoE is primarily synthesized by astrocytes under normal physiological conditions (Mahley, 1988) and the neuropathological effects of astrocyte-derived apoE are a point of ongoing study. However, apoE can also be produced by neurons under pathological conditions resulting from neuronal cell injury or stress (Huang, 2010; Huang and Mucke, 2012; Mahley et al., 2006). Xu et al. (2006) established an enhanced green fluorescent protein (EGFP)apoE-reporter mouse model in which EGFP was inserted into one allele of the apoE gene to serve as a reporter of apoE expression.

, 2007) Thus, short-lived focal increases in gamma-band power ar

, 2007). Thus, short-lived focal increases in gamma-band power are not unique to conscious states but track activation of both conscious and nonconscious local cortical circuits ( Ray and Maunsell, 2010). However, their significant enhancement on consciously perceived trials, turning into an all-or-none pattern after 200 ms, appears as a potentially more specific marker of conscious access ( Fisch et al., 2009 and Gaillard et al., 2009). The

high spatial precision and signal-to-noise ratio afforded Selleck Osimertinib by intracranial recording in epileptic patients provides essential data on this point. Gaillard et al. (2009) contrasted the fate of masked (subliminal) versus unmasked (conscious) words while recording from a total of 176 local sites using intracortical depth electrodes in ten epileptic patients. Four objective signatures of conscious perception were identified (Figure 3): (1) late (>300 ms) and distributed event-related potentials contacting sites in prefrontal cortex; (2) large and

late (>300 ms) increases in induced power (indexing local synchrony) in high-gamma frequencies (50–100 Hz), accompanied by a decrease in lower-frequency power (centered around 10 Hz); (3) increases in long-distance cortico-cortical synchrony in the beta frequency band 13–30 Hz; (4) increases in causal relations among distant cortical areas, bidirectionally but more strongly in the bottom-up direction (as assessed by Granger causality, a statistical technique that measures whether the time course of signals at one site can forecast the future evolution of signals at another Dabrafenib order distant site). Gaillard et al. (2009) noted that all four signatures coincided in the same time window (300–500 ms) and suggested that they might constitute different measures of the same state of distributed ADAMTS5 “ignition”

of a large cortical network including prefrontal cortex. Indeed, seen stimuli had a global impact on late evoked activity virtually anywhere in the cortex: 68.8% of electrode sites, although selected for clinical purposes, were modulated by the presence of conscious words (as opposed to 24.4% of sites for nonconscious words). Neuronal recordings. A pioneering research program was conducted by Logothetis and collaborators using monkeys trained to report their perception during binocular rivalry ( Leopold and Logothetis, 1996, Sheinberg and Logothetis, 1997 and Wilke et al., 2006). By recording from V1, V2, V4, MT, MST, IT, and STS neurons and presenting two rivaling images, only one of which led to high neural firing, they identified a fraction of cells whose firing rate increased when their preferred stimuli was perceived, thus participating in a conscious neuronal assembly. The proportion of such cells increased from about 20% in V1/V2 to 40% in V4, MT, or MST to as high as 90% in IT and STS.

, 2005 and Yazawa et al , 2005) Remarkably, α-synuclein is not n

, 2005 and Yazawa et al., 2005). Remarkably, α-synuclein is not normally expressed by oligodendrocytes

(Miller et al., 2005), selleck screening library and a fundamental question remains about the origin of this protein: is it taken up from neurons, or does the pathological process activate expression by glia? In fact, the pathology shows relatively little deposition of synuclein in neurons, with only occasional nuclear and cytoplasmic inclusions (Farrer et al., 2004, Jellinger and Lantos, 2010 and Nishie et al., 2004b). At the same time, oligodendrocytes do not seem to upregulate expression of α-synuclein even in MSA (Miller et al., 2005). Regardless of its source, α-synuclein accumulates to particularly high levels in MSA, Regorafenib solubility dmso suggesting a process distinct from Lewy pathology. In addition, GCI lesions are widespread in MSA but generally correlate with neuron loss in the substantia nigra, pons, cerebellum, and intermediolateral cell columns of the spinal cord, suggesting that the glial process may be primary. MSA is rarely familial (Soma et al., 2006) and mutations in α-synuclein have not been observed (Ozawa et al., 2006). However, polymorphisms in the synuclein gene may influence susceptibility to MSA (Al-Chalabi

et al., 2009 and Scholz et al., 2009). In addition, the analysis of familial MSA has recently identified mutations in COQ2, a protein required for the synthesis of coenzyme Q10 (Multiple-System Atrophy Research Collaboration, 2013). The degenerative process in MSA may thus reflect a primary lesion in mitochondria. DLB more closely resembles idiopathic PD in terms of Lewy body

pathology. Although DLB appears to be a distinct syndrome, with early cognitive impairment, fluctuating alertness, and visual hallucinations in addition to progressive parkinsonism, the distribution of Lewy pathology appears remarkably similar to that observed in PD, with a brainstem-predominant form and others involving the cortex as well (McKeith et al., 2005). Like PD and MSA, DLB also involves primarily the deposition of α-synuclein. Lewy pathology was originally considered to involve only α-synuclein, but β- and γ- can deposit in both PD and DLB (Galvin et al., 1999). Similar to α-synuclein, β- accumulates presynaptically in PD, but γ- forms axonal spheroids. β-synuclein has been suggested to ameliorate GBA3 the toxicity of α-synuclein by reducing either its aggregation or its expression (Fan et al., 2006 and Hashimoto et al., 2001). However, polymorphisms in β-synuclein predispose to DLB (Ohtake et al., 2004), and transgenic mice overexpressing the variant develop degeneration and behavioral abnormalities (Fujita et al., 2010). These animals do not develop typical Lewy pathology, but they do accumulate both α- and β-synuclein in axonal spheroids (Fujita et al., 2010). Indeed, β-synuclein appears as toxic as α-synuclein to cultured neurons (Taschenberger et al., 2013).

Furthermore, this reaction time improvement increased following a

Furthermore, this reaction time improvement increased following a sequence switch in each block (Figure 2D). We recorded the activity of 553 (230 in monkey 1, 323 in monkey 2) neurons in 3-deazaneplanocin A clinical trial the lateral prefrontal cortex (lPFC) and 453 (210 in monkey 1, 243 in monkey 2) neurons, in the dorsal striatum (dSTR) predominantly in the caudate nucleus. Neural activity was recorded simultaneously from both areas in all sessions. All reported effects were consistent in both animals. Therefore, the data were pooled. We examined activity relative to five factors; the task condition, the sequence executed in each trial,

the specific movement being executed, the color bias, and learning related action value, which was estimated using a reinforcement learning algorithm (see Experimental Procedures). Sequence and learning effects were less well defined in the random sets, but because of the consistent task structure we analyzed them as an internal control. We began by analyzing activity using an omnibus ANOVA, across conditions, and then split the data by task condition to examine more specific hypotheses. Neurons buy Saracatinib were found which were related to all variables of interest. For example, some neurons had responses which depended on the specific movement being executed, but which also depended on the task condition (Figure 3). This lPFC neuron tended

to respond strongly to the first and last movement of all sequences in both the random and the fixed condition, as has been seen in previous studies (Fujii and Graybiel, 2003).

However, it also had a robust response to the second movement in sequence one and sequence five, but only in the fixed condition. We also found neurons related to the color bias. For example, in the random sets (Figure 4A) this dSTR neuron had a very strong baseline firing rate which was additionally modulated with the color bias (Figure 4B), an effect which became statistically significant just after movement onset (Figure 4C). Sequence selection also was modeled using a reinforcement learning algorithm (see Experimental Procedures). This allowed us to track the animal’s estimate of the value of each eye movement, movement by movement and trial by trial. For example, in the fixed condition, following a switch from a block in which sequence seven had been correct to a block in which sequence two was correct the animal continued trying to execute sequence seven in the first trial and the value estimates reflected this. The first execution of the leftward movement had a high value (1.0) as this had been correct in the previous block (Figure 5A; switch + 0). After this point, the animal still believed that the sequence had not switched and therefore it executed a downward movement for the second movement. As the animal would assume this was correct, this movement would have a high value (1.0).

The crucial observation is that persistent eigenmodes of network

The crucial observation is that persistent eigenmodes of network diffusion appear homologous to characteristic atrophy patterns observed in various dementias. selleck products The first (steady-state) eigenmode, whose eigenvalue is zero, is not shown here, because it is relatively uninteresting, varying simply according to region size, in rough correspondence to atrophy

seen in normal aging. In order to ensure that these results are not due to a specific choice of volumetric algorithm or choice of anatomic atlas, we repeated the same study using volumetric data obtained by the FreeSurfer software (Fischl et al., 2002) and a different 86-region atlas (Figure 5). Measured atrophy patterns generally match the cortical atrophy seen using the automated anatomic labeling (AAL) atlas (Figure 4), but exact match is not to be expected due to both methodological as well as ROI size and shape differences. It is important to note, however, that the visual correspondence between eigenmodes and atrophy remains intact, and the former generally agree with classic

AD/bvFTD pathology, which implies that these results are not methodology-specific. We show later (Figures S5 and S6 available online) that our results are also insensitive to inter-subject variability. The second-most persistent mode (Figures 2 and 4, top rows) closely resembles typical Alzheimer’s atrophy in mesial temporal, posterior cingulate, and limbic structures, as well as lateral temporal and dorsolateral frontal cortex (Apostolova et al., 2007 and Thompson et al., 2003). This eigenmode shows strong involvement of the medial and lateral temporal lobes, which CB-839 molecular weight are involved in memory, and the dorsolateral prefrontal cortex, implicated in working memory (Curtis and D’Esposito, 2003). The main fibers connecting these regions are the superior longitudinal fasciculus (SLF), the splenium of corpus callosum, and the cingulum bundle. While agreement is good with our own volumetric findings (Figures 2 and 4, bottom

rows) and excellent with the published MycoClean Mycoplasma Removal Kit literature (see, for instance, Apostolova et al., 2007, Thompson et al., 2003 and Seeley et al., 2009), there were some areas of disagreement with our volumetric findings in the parietal, frontolateral, and frontoinsular areas. We attribute these differences to small sample size and technical limitations of tractography, co-registration, and volumetrics. The third persistent eigenmode (Figures 3 and 5, top rows) is in good agreement with our bvFTD data (Figures 3 and 5, bottom rows) and published findings (Du et al., 2007, Boxer and Miller, 2005 and Seeley et al., 2009), which indicate prominent atrophy in the orbitofrontal and anterior cingulate regions. This eigenmode is particularly strong in the lateral temporal lobe and the superior frontal, dorsolateral, and orbital cortices— areas that deal with executive function, decision making, expectation, balancing risk versus reward, and inhibition.

This process gave us τ summarizing the similarity in activity bet

This process gave us τ summarizing the similarity in activity between all pairs of trial blocks. Resampling methods were used to confirm whether a τ for two blocks was unusually low. That is, we obtained an empirical distribution of τ obtained under the null hypothesis that the pattern of activity MAPK inhibitor for two blocks was the same. The τ computed between two blocks was considered to have changed if it was lower than 99% of the τ values obtained for the empirical distribution under the null hypothesis. Using this approach, we could determine whether

a neuron changed its firing pattern from block to block. To test whether a neuron rescaled its delay activity when the delay was doubled, the same approach was taken, but the PSTH for the longer delay used time bins whose duration was also doubled. Further details on this analysis are provided in the Supplemental Experimental Procedures. To assess the effect of time on firing rate related to the rat’s position, we generated spatial firing rate maps for the delay zone as 1 × 1 cm bins, and calculated occupancy-normalized firing rates. To assess firing rates

related to head direction, we assigned each head direction observation to 1 of learn more 60, nonoverlapping 6° bins and calculated occupancy-normalized firing rates for each bin. Speed firing rate plots were based on computations of the difference in the X-Y position for successive frames, assigned to 1 of 30 speed bins that spanned 0–30 cm/s, and occupancy-normalized firing rates were calculated for each bin. ANOVAs were performed on trial-by-trial, unfiltered firing rates for each 1 s segment of the delay. We used only those bins whose firing rate could be estimated in all of the 1 s segments across trials, allowing

an ANOVA with factors time and bin to test whether time modulated neural activity. Further details on this method are provided in the Supplemental Experimental Procedures. Analysis of LFP frequency as a function of time used the multi-taper (-)-p-Bromotetramisole Oxalate functions written for MATLAB that are freely available as part of the Chronux toolbox (Mitra and Bokil, 2008; http://www.chronux.org). For the delay the trial-averaged multi-tapered spectrum was determined (mtspectrumc.m) using a window size of 1 s that started at the beginning of the delay and was slid across time using 100 ms increments. For the object and odor periods, a window size of 1.2 s was time locked to the beginning of the either period and slid with one 100 ms increment. The trial-averaged spectrum was computed separately depending on the object that was presented. For a given tetrode, in order to test whether θ (i.e., 4–12 Hz) power differed depending on the object presented during each trial period, a trial-average spectrogram was generated using the same parameters as above except that the frequency range was confined to 4–12 Hz. Further details of the ANOVAs performed on the LFPs are provided in the Supplemental Experimental Procedures.

The scan head was placed onto an inverted Nikon TE2000-U microsco

The scan head was placed onto an inverted Nikon TE2000-U microscope (Nikon) table equipped with differential micrometers (OptoSigma) for precise positioning. A custom-built laser confocal set up was used to record fluorescence simultaneously with topography. Excitation was provided by an LCS-DTL-364 laser diode (473 nm wavelength, Laser Compact). The fluorescence signal was collected using a 100× 1.3 NA oil-immersion objective, an epifluorescence filter block, and a photomultiplier with a pinhole (D-104-814, Photon Technology

International) or in nonconfocal mode using wide-field illumination and an Evolve 512 EM-CCD camera (Photometrics). Fine-tipped nanopipettes used both to probe the neuronal topography and to perform cell-attached patch-clamp selleck products recordings were pulled from borosilicate glass (OD 1 mm, ID 0.5 mm, Sutter Instruments) MDV3100 purchase using a horizontal laser-based puller P-2000 (Sutter

Instruments). The pipette resistance was in the range of ∼80–110 MΩ, corresponding to an estimated inner tip diameter of ∼90–125 nm (Figure 3E). Nanopipettes were held in voltage-clamp mode with an Axopatch 200B patch-clamp amplifier coupled to a DigiData 1322A interface (Molecular Devices). Topographic and confocal images were obtained, first, by aligning the nanopipette tip with the fluorescence microscope focal plane and, second, by recording topographic and fluorescence images while scanning the specimen in the x and y axes by the SICM Oxalosuccinic acid electronics. The laser-pulling process generates, from a single capillary, a pair of “twin” nanopipettes with virtually identical geometries. One of the pair was used as a representative of the tip geometry before pipette breaking. The other was subjected to the controlled widening procedure as described in the main text. In this set of experiments, the ultrafiltered standard extracellular

solution (20 nm syringe filter) was used both in the pipette and in the bath. The pipette resistance was monitored before and after the breaking procedure using the Seal Test function of pCLAMP 9.2 (Molecular Devices). Immediately after completion of the breaking procedure, the pipette solution was removed and the pipette tip was washed three times with ultra filtered 96% ethanol and dried. Both modified and unmodified pipettes were sputter coated with gold (15 nm coat thickness) and imaged using an FEI Quanta 3D FEG (FEI) scanning electron microscope operating in high vacuum mode at 30 kV. Dimensions and cone angle of pipette tips were measured in ImageJ (U.S. National Institutes of Health). After topographic and confocal images were obtained, the coordinates of a defined ROI on the neuron surface (synaptic bouton or dendrite) were used for precise positioning of the SICM pipette for cell-attached patch-clamp recording. The nanopipette was then lowered by the z axis piezo control until it made contact with the cell surface.