Thus, despite receiving substantial excitatory input from recepto

Thus, despite receiving substantial excitatory input from receptor neurons, the MCs have a very small response (defined here as the change in the MC firing rate with respect to spontaneous baseline activity). The reduction in the response is due to the strong inhibition provided by the GC. These inhibitory inputs almost completely cancel the excitation provided Fludarabine clinical trial by the receptor neurons (see “The State Dependence of the MC Code” in Experimental Procedures for a more quantitative analysis). The

balance between excitation and inhibition has implications for olfactory code carried by the MCs. For the MCs that receive inhibitory inputs from the GC, the odorant responses are substantially reduced. If all MCs receive these inhibitory inputs, only weak (i.e.,

undetectable) activity that is necessary to drive GC above the firing threshold remains (Figure 2B). Because the inputs to all of the MCs are substantially balanced by inhibition and none of the MCs displays strong odorant responses in this case, we call this complete combinatorial compensation. On the other hand, as shown in Figure 2C, the responses of a subset of MCs may accurately reproduce the inputs that they receive from the receptor neurons. This is because these cells do not receive the compensating inhibition from the GC, which is therefore incomplete. Inhibitory inputs from GCs selectively reduce the responses of some MCs, while leaving other MCs responsive. The Selleck Rucaparib sustained combinatorial representation carried by MCs becomes sparse. Therefore, our model can yield sparse sustained MC responses observed experimentally (Rinberg et al., 2006). Sparsening of the responses of MCs reduces redundancy in the representation of odorants (Figure 3A). The role of GCs in this case is to remove overlaps between combinatorial receptor inputs. The removal of overlaps makes MC activation patterns Thiamine-diphosphate kinase more independent for different odorants. Redundancy reduction may occur in a task-dependent manner. This means that the particular overlap that is removed depends on the activation of the centrifugal cortico-bulbar projections (Figure 3A versus

Figure 3B). By activating/deactivating the particular subsets of GCs, these projections may change the MC code to better discriminate the set of odorants relevant to specific behavior. Figure 2A illustrates the regime when GCs are never or rarely active. The MC code in this case is dense and reflects glomerular inputs. Activation of GCs, as shown in Figure 2B, leads to sparse odorant representations. Because the transition between full and sparse codes occurs upon transition between anesthetized and awake states, we suggest that Figures 2A and 2B illustrate these two regimes of the bulbar network. The prediction of this model is therefore that GCs are less active in anesthetized animals than in awake and behaving animals. We now consider the case of several GCs.

, 2011) Of the proteins that bound selectively to ecto-LPHN3-Fc,

, 2011). Of the proteins that bound selectively to ecto-LPHN3-Fc, FLRT2 and FLRT3 were among the most abundant

and were of particular interest due to similarities in domain organization to previously identified postsynaptic organizing molecules such as the LRRTMs (de Wit et al., 2011), which were not detected in our purification Roxadustat mw (Figure 1B). We also identified proteins in the Teneurin family (also named ODZs), which have recently been reported as ligands for LPHN1 (Silva et al., 2011) (see Figure S1A available online). Because FLRT3 was the most abundant FLRT protein identified in the ecto-LPHN3-Fc pull-down, we carried out complementary experiments with ecto-FLRT3-Fc to confirm this interaction MAPK Inhibitor Library (Figures 1B and S1A). Affinity chromatography and mass spectrometry using ecto-FLRT3-Fc resulted in the identification of a large number of LPHN1 and LPHN3 peptides, with relatively fewer LPHN2 peptides, but not the abundant presynaptic organizing protein NRXN1 (Figure 1C). UNC5B (Figure S1B), a previously reported FLRT3 interactor, was also identified, but at much

lower abundance (Karaulanov et al., 2009, Söllner and Wright, 2009 and Yamagishi et al., 2011). When total spectra counts from proteins identified in both purifications were compared, LPHN3 and FLRT3 stood out clearly as the proteins most frequently detected in both purifications (with each as bait in one condition and prey in the other) (Figure 1D). To support our mass spectrometry results, we verified the association of FLRT3 with LPHN3 by western blot in similar ecto-Fc pull-down assays on rat brain extract and transfected heterologous cell lysate (Figures 1E–1I). Together, these findings suggest that FLRTs likely represent endogenous ligands for latrophilins. To test whether FLRT3 and LPHN3 can all bind to one another in a cellular context, we expressed FLRT3-myc

in HEK293 cells and applied ecto-LPHN3-Fc or control Fc protein. We observed strong binding of ecto-LPHN3-Fc to cells expressing FLRT3-myc, but no binding of Fc (Figure 1J). Ecto-LPHN3-Fc did not bind to cells expressing myc-LRRTM2 (Figure S1D), showing that the LPHN3-FLRT3 interaction is specific. Ecto-LPHN3-Fc also bound strongly to the other FLRT isoforms, FLRT1 and FLRT2 (Figure S1D), and ecto-LPHN1-Fc bound to all FLRT isoforms as well (Figure S1C). Complementarily, ecto-FLRT3-Fc, but not control Fc, bound strongly to cells expressing LPHN3-GFP (Figure 1K). Ecto-FLRT3-Fc also bound the previously identified interactors UNC5A, UNC5B, and UNC5C, but did not bind to NRXN1β(+ or −S4)-expressing cells (data not shown). We also confirmed that ecto-LPHN3-Fc, but not ecto-FLRT3-Fc, could bind to cells expressing teneurin 3, confirming that LPHNs and teneurins can indeed interact (Figure S1E). Thus, we find that LPHNs and FLRTs strongly interact, with promiscuity between isoforms.

433, p = 0 083; Figure S1) In contrast, such a correlation was n

433, p = 0.083; Figure S1). In contrast, such a correlation was not observed in control subjects. Together, Obeticholic Acid purchase these results demonstrate that patients with aMCI had increased activation in the left DG/CA3 subregion of the hippocampus and performed significantly worse on trials with lures, compared to the age-matched control group. These results are consistent with earlier reported findings (Yassa et al., 2010), providing

a sample of aMCI patients appropriate for testing the functional significance of hippocampal hyperactivity. To assess whether low dose levetiracetam treatment reduces the observed increased DG/CA3 activation, we compared functional data during the fMRI memory task in aMCI patients upon completion of the placebo and drug treatment

phases. Within the neuroanatomical region of task related activity (as shown click here in Figure 2C), low-dose levetiracetam significantly reduced BOLD activation relative to the placebo treatment (t = 2.537, p = 0.022; Figure 2F). This effect was not influenced by order of treatment. Thus there was no evidence for any carryover 6 weeks after drug treatment was terminated (4 week washout and placebo for 2 weeks before the second assessment). Further, BOLD activation in the DG/CA3 in aMCI patients after levetiracetam treatment did not differ from activity in that region observed in the normal age-matched control subjects. To confirm this finding, a separate analysis was conducted in which voxel selection was based on a one-way ANOVA of trial type only in control subjects. This analysis resulted in an area of task related activity similarly localized in the left DG/CA3 subregion of the hippocampus (Figure 2D). The effect of drug treatment was confirmed by comparing fMRI activation in that area of task related activity in aMCI patients on placebo and levetiracetam.

In that analysis, patients with aMCI taking low dose levetiracetam again showed significantly reduced BOLD activation relative to their activity under placebo treatment (t = 2.192, p = 0.044; Figure 2G). These effects were obtained with a drug dose well below that used and clinically for the treatment of epilepsy; drug levels in patients were determined to be 4.4 mcg/ml ± 0.53 (mean and SEM), as compared to typical ranges for efficacy as an antiepileptic with doses of 1,000–3,000 mg/day achieving levels of 10–40 mcg/ml (Lyseng-Williamson, 2011). Treatment with low dose levetiracetam also improved behavioral performance on the lure trials. Analysis of the responses on those trials showed an overall significant effect of response type (F(1,16) = 5.992, p = 0.026), and, importantly, a significant interaction effect of treatment (drug versus placebo) × response (F(1,16) = 5.

Statistical analysis revealed no effect of group, F (1, 21) = 0 7

Statistical analysis revealed no effect of group, F (1, 21) = 0.7, p = 0.412, but an effect of devaluation, F (1, 21) = 4.71, p = 0.042, and a group × devaluation interaction, F (1, 21) = MLN0128 research buy 4.40, p = 0.048; whereas the Sham, F (1, 21) = 5.10, p = 0.035 and Ipsi groups, F (1, 21) = 3.84, showed reliable devaluation effects, the Contra group did not, F (1, 21) = 0.211, p = 0.651. In the outcome-selective reinstatement test (Figure 4H), Group Sham and Group Ipsi both showed selective reinstatement but Group Contra did not.

There was no effect of group, F (1, 21) = 0.38, p = 0.545, a main effect of responding in the pre versus post periods, F (1, 21) = 12.61, p = 0.002; however, the postoutcome reinstatement was specific to the lever associated with that outcome only in Group Sham, F (1, 21) = 6.81,

p = 0.016, and Group Ipsi, F (1, 21) = 6.1, p = 0.022, but was divided equally between levers in Group Contra, F (1, 21) = 0.17, p = 0.898. The impairments observed in Group Contra here echo those previously observed as a result of bilateral Pf lesions. To confirm the effect of the Pf find more lesions on CIN function in the pDMS, we examined p-Ser240-244-S6rp intensity in ChAT-immunoreactive neurons in the intact pDMS in rats drawn from the Sham, Ipsi, and Contra groups perfused immediately after the reinstatement test. To assess specificity, we also compared p-S6rp intensity in ChAT-immunoreactive neurons in the dorsolateral striatum (DLS) in these groups. The results of these analyses are presented in Figures 5A, 5B, and 5C. As is clear from these figures, different levels of p-S6rp L-NAME HCl intensity in the pDMS were observed among groups: p-S6rp signal was significantly reduced in CINs from Group Contra (the disconnection group) compared to CINs from both Group Ipsi and Group Sham (the controls), based on the quantification presented in Figures 5B and 5C (F (1, 9) = 17.54, p < 0.001). These differences were

specific to the pDMS and, as observed previously (cf. Figure 2G), were not observed in the DLS; F (1, 9) = 0.32, p = 0.587. Using brain sections from the same experiment, we further examined whether the Pf lesion principally affected CINs in the pDMS, or whether the medium spiny neurons (MSNs) in this region were affected as well, based on the proportion of Pf glutamatergic inputs to MSNs and the complex regulation of MSNs by the Pf (Ellender et al., 2013). We took advantage of the phospho-Thr202-Tyr204-extracellular regulated kinase 1/2 (phospho-ERK1/2) detection in the striatum, a method shown to reliably reflect neuronal activation in MSN populations (Bertran-Gonzalez et al., 2008; Shiflett and Balleine, 2011a, 2011b).

, 2004 and Riley, 1981) Thus, synapse elimination seems to be un

, 2004 and Riley, 1981). Thus, synapse elimination seems to be underway just as animals are being born. Because of the large volume being reconstructed, it was possible in some cases to trace the axons back far enough to assess whether the same axon Palbociclib in vitro was innervating

more than one of the three adjacent neuromuscular junctions. Of the 26 terminal axon branches innervating these three junctions, seven were traceable back to branch points where they bifurcated to give rise to innervation to two of the three junctions (Figure 5). In six of the seven cases, the axons innervated comparably sized percentages of each of the junctions (6% versus 10%; 16% versus 10%; 8% versus Selleck Afatinib 17%; 4% versus 10%; 17% versus 14%; 21% versus 16%). In one case, however, we saw that one of the axon branches did not establish a synaptic contact with the neuromuscular junction site but rather terminated in a bulb

just proximal to one of the junctions. The ultrastructural appearance of this axonal bulb suggested, as described above, that it was a retracting axon, i.e., there were nearby local shed axosomes, and it had a smooth shape (Bishop et al., 2004), rather than a growth cone (i.e., it showed no filopodia or lamellopodia). This result suggested that an axon branch was already in the process of retracting in the first postnatal day. Another branch of the same axon innervated 15% of the neuromuscular junction area on an adjacent neuromuscular junction

(axon 7 in Figure 5). Thus, this early stage of branch loss is occurring asynchronously among the branches of one axon. This result lends further support to the conclusion that the initial axon pruning decisions are being made at the level of terminal branches and not more proximally in the axon arbor. Moreover, the fact that most axons are being maintained at a neuromuscular junction while one is being removed supports the idea that beginning at birth, during the earliest stages of synapse elimination, different axons are being sequentially removed from junctions rather than synchronously. The serial reconstructions also provided information about the way multiple until axons coinnervated neuromuscular junctions at birth. Many of these features were different from both adult singly innervated neuromuscular junctions and later-stage multiply innervated junctions. The synaptic contacts of the axons were highly intermixed, showing no evidence of the interaxonal segregation found at later stages of the elimination process (Gan and Lichtman, 1998) (Figure 4A). The branches of the different axons were not only intermixed but also were closely juxtaposed to each other, with their membranes abutting without intervening Schwann cell processes (Figure S1A, boxed region).

For example, to distinguish direct from indirect synaptic connect

For example, to distinguish direct from indirect synaptic connections, the uncertain parameter of “time after injection” was often used as a determinant EPZ-6438 order for this critical distinction (Jovanovic et al., 2010,

Rathelot and Strick, 2006 and Ugolini, 2010). A recently introduced modification to this technology now allows for an unambiguous assignment of synaptic connectivity in the central nervous system (CNS) (Callaway, 2008 and Wickersham et al., 2007). In this strategy, genomic deletion of the gene encoding a glycoprotein (Gly) essential for transsynaptic spread renders the rabies virus spreading incompetent. Introduction of Gly expression by genetic or viral tools to selectively complement Gly-deficient rabies in primarily infected neurons reestablishes the ability for transsynaptic spread to label neurons presynaptic to primary infection but prohibits subsequent Thiazovivin mw spread due to absence of Gly in presynaptic neurons (Callaway, 2008 and Wickersham et al., 2007). This monosynaptically restricted transsynaptic rabies virus system was used in two recent studies to map the three-dimensional distribution

of spinal interneurons with direct synaptic connections to motor neurons in mice (Stepien et al., 2010 and Tripodi et al., 2011). Using retrograde motor axonal coinfection strategy from specific muscles, transsynaptic spread is initiated from functionally defined motor neuron Methisazone pools (Figures 5A and 5B). Analysis of the overall distribution patterns of spinal premotor interneuron connectivity to an individual motor neuron pool demonstrates a high degree of reproducibility across animals (Stepien et al., 2010). In contrast, analysis of premotor interneurons

connecting to motor neuron pools with distinct function in motor behavior reveals striking differences in overall distribution (Stepien et al., 2010 and Tripodi et al., 2011). These observations uncover the existence of anatomical or structural engrams at the premotor circuit level that correlate with motor function. The results raise a number of interesting and currently unresolved questions. Premotor neurons encompass a diverse array of neuronal subpopulations, including distinct neurotransmitter phenotypes, synaptic input driving their activation, and additional synaptic partners contacted. It will be interesting to determine the relationship between connectivity-based anatomical maps and functional maps assessing activity patterns in relation to locomotor output. At present, it is unclear which of the many premotor interneurons are required for or involved in the core components of interneuron circuits that give rise to rhythm generation and perpetuation. In addition, motor neuron pools may tap into connections from distinct possible premotor interneuron populations differentially.

One key idea implicit in both algorithmic frameworks is the idea

One key idea implicit in both algorithmic frameworks is the idea of abstraction layers—each level of the hierarchy need only be concerned with the “language” of its input area and its local job. For example, in the serial chain framework, while workers in the middle of a car assembly line might put in the car engine, they do not need to know the job description of early line

workers (e.g., how to build a chassis). In this analogy, the middle line workers are abstracted away from the job description of the early line workers. Most complex, human-engineered systems have PD-0332991 order evolved to take advantage of abstraction layers, including the factory assembly line to produce cars and the reporting organization of large companies to produce coordinated action. Thus, the possibility that each cortical area can abstract away the details below its input area may be critical for

leveraging a stack of visual areas (the ventral stream) to produce an untangled object identity representation (IT). A key advantage of such abstraction is that the “job description” of each worker is locally specified and maintained. The trade-off is that, in its strongest instantiation, no one oversees the online operation of the entire processing chain and there are many workers at each level operating in parallel without explicit coordination (e.g., distant parts of V1). Thus, the proper selleck products upfront job description at each local cortical subpopulation must be highly robust to that lack of across-area and within-area supervision. In principle, such robustness could arise from either an ultraprecise, stable set of instructions given to each worker upfront (i.e., precise genetic control of all local cortical synaptic weights within the subpopulation), or from a less precise “meta” job description—initial instructions that are augmented by learning that continually

refines the daily job description of each worker. Such learning mechanisms could involve feedback (e.g., Hinton et al., 1995; MTMR9 see above) and could act to refine the transfer function of each local subpopulation. We argue above that the global function of the ventral stream might be best thought of as a collection of local input-output subpopulations (where each subpopulation is a “worker”) that are arranged laterally (to tile the visual field in each cortical area) and cascaded vertically (i.e., like an assembly line) with little or no need for coordination of those subpopulations at the time scale of online vision. We and others advocate the additional possibility that each ventral stream subpopulation has an identical meta job description (see also Douglas and Martin, 1991, Fukushima, 1980, Kouh and Poggio, 2008 and Heeger et al., 1996). We say “meta” because we speculate about the implicit goal of each cortical subpopulation, rather than its detailed transfer function (see below).

, 2004 and Towne et al , 2010) Recently, a modified retrograde a

, 2004 and Towne et al., 2010). Recently, a modified retrograde approach has been developed to map the entire synaptic network converging onto a single cell, labeled with in vivo microelectroporation (Marshel et al., 2010), a technical

advance that could well dovetail with optogenetic control. As described above, the limitations imposed by packaging capacity in viral systems Paclitaxel in vivo can be overcome using single-component optogenetic tools (for example, by using recombinase-dependent opsin-expressing viruses and/or by leveraging relevant anatomy for projection targeting). Beyond the benefits of speed, flexibility, spatiotemporal targeting precision, and high gene copy-number, virus injection into recombinase driver lines also can uncouple promoter specificity from expression strength, since opsin expression is related to the copy number of the virus with its strong nonspecific promoter, and resulting transcription can exceed endogenous transcription from tissue-specific promoters. Protein Tyrosine Kinase inhibitor However, another major class of strategy, generation of mouse transgenic lines directly expressing opsin genes under local promoter-enhancer regions (i.e., not in a recombinase-dependent fashion), provides a distinctly useful means of achieving cell-type-specific opsin expression. While transgenic mouse

lines require effort, time, and cost associated with production and maintenance, the convenience and reliability of homogeneous opsin-expressing animals STK38 provides major experimental leverage. The Thy1::ChR2-EYFP mouse lines (Arenkiel et al., 2007 and Wang et al., 2007) express ChR2 under control of the Thy1 promoter. While as discussed above promoters

do not suffice to completely define cell types and the complement of labeled cells must be considered in each case, Thy1-driven expression is largely restricted to projection neurons, enabling several studies in which optogenetics was applied to study cortical connectivity (Wang et al., 2007), transmission from the olfactory bulb to cortex (Arenkiel et al., 2007), aspects of ganglion cell function in visual impairment (Thyagarajan et al., 2010), cortical information processing (Sohal et al., 2009), and parkinsonian circuitry (Gradinaru et al., 2009). For example, in the latter study it was found that therapeutic deep brain stimulation (DBS) in the subthalamic nucleus (STN) arising from a point source (e.g., electrode or fiber) is by far most effective when the direct target is afferent axons within the structure (these axons then efficiently modulate both downstream and upstream neurons—and indeed potently reduce local STN spiking); much weaker effects were seen with direct modulation of local cell bodies in the STN by a point source of control, suggesting electrical DBS might be best designed to target axonal tracts rather than gray matter. A defined local cell type was targeted in a pioneering study by Kiehn and colleagues (Hägglund et al.

We semiquantitatively assessed ThioS staining using a blinded rat

We semiquantitatively assessed ThioS staining using a blinded rater who gave a score from 1 (no staining) to 5 (maximum staining) in all control and anti-tau antibody-treated mice. By semiquantitative assessment, HJ8.5 treatment significantly reduced ThioS staining compared to PBS and HJ3.4 (Figures 6A and 6B). We also stained mice treated with PBS, HJ8.5, and HJ9.3 (n = 6 from each group) with PHF1 monoclonal antibody, which recognizes tau phospho-residues Ser396 and Ser404 (Otvos et al., 1994). AT8 and PHF1 staining significantly correlated (r = 0.630, p = 0.005) (Figure S7A), showing that two anti-phospho tau antibodies to different tau epitopes give similar

results. Many neurodegenerative diseases, including tauopathies, exhibit microglial activation in areas of the brain surrounding protein

aggregation and cell injury. We assessed microglial activation in the treatment groups using anti-CD68 antibody R428 cell line (Macauley et al., 2011) (Figures S6C–S6G). HJ8.5 and HJ9.3 treatment reduced microglial activation in piriform cortex, entorhinal cortex, and amygdala compared to controls (Figures S6C–S6G). HJ9.4 had a weaker effect in the piriform cortex compared to HJ8.5 and HJ9.3 (Figures S6E–S6G), consistent Docetaxel nmr with the AT8 staining results (Figure 4A). Microglial activation strongly correlated with AT8 staining in all samples (r = 0.511, p = 0.0038) (Figure S7B). To determine the level of soluble and insoluble tau in the cortex, we performed sequential biochemical extraction with RAB (aqueous buffer), radio immunoprecipitation assay (RIPA) (detergent buffer), and 70% formic acid (FA) to solubilize the final pellet. We quantified total tau by ELISA with anti-tau antibody HJ8.7, which detects both human and mouse tau with the same

KD (0.34 pM). We excluded the possibility that the treatment antibodies would interfere with the ELISA by spiking positive control samples with these antibodies prior to analysis and observing no interference (data not shown). We analyzed all mice that were assessed by pathological analysis in Figure 5. Total tau levels in the RAB (Figure 6A)- or RIPA (Figure 6B)-soluble fractions were similar among all groups. We analyzed the detergent-insoluble/70% Cediranib (AZD2171) FA-soluble fractions by neutralizing the samples prior to ELISA and western blot. We analyzed every animal studied and found that HJ8.5 and HJ9.3 decreased detergent-insoluble tau by >50% versus controls (Figure 6C). Representative samples (n = 4 from each group) illustrate by western blot decreased levels of insoluble tau in mice treated with HJ8.5 and HJ9.3 (Figure S7C). Insoluble tau levels were no different in HJ9.4-treated groups versus PBS or HJ3.4. We also assessed human and mouse tau specifically in the detergent-insoluble/70% FA-soluble fractions in n = 6 mice per group in which the mean AT8 staining reflected the mean values of results in Figure 4.

, 2010) and point to a critical role for inhibitory GABAergic neu

, 2010) and point to a critical role for inhibitory GABAergic neurons in gating active touch sensory responses in supragranular pyramidal cells (Gabernet et al., 2005, Sun et al., 2006, Haider et al., 2010 and Cruikshank et al., 2010). Future studies should define more precisely the role of different subtypes of inhibitory GABAergic neurons during active touch, which can now be approached through two photon targeted recordings of cell-type-specific GFP-expressing mouse lines (Margrie et al., 2003, Liu et al., 2009 and Gentet et al., 2010)

or by selectively manipulating cortical neuron subpopulations during functional operation through combinations of optogenetics, viral gene transduction, and mouse genetics (Boyden et al., 2005, Cardin et al.,

2009 and Sohal et al., 2009). Through such further experimentation in combination with computational modeling, it will be of great interest to investigate GSK1349572 the circuit determinants of the hyperpolarized touch-evoked reversal potentials and whether the PSP reversal potential is fixed for a given neuron or whether it can be modulated by context, behavior, and learning. Here, in this study, we provide detailed measurements of the synaptically driven membrane potential dynamics of identified neurons within a specific MLN0128 well-defined cortical column in actively sensing mice. Such data form an essential step toward a causal and mechanistic explanation for the functional operation of neocortical microcircuits during behavior at the level of individual neurons and their synaptic inputs. The experimental procedures are described in detail in the Supplemental Information. All experimental procedures were approved by the Swiss Federal Veterinary Office. C57BL6J or GAD67-GFP mice were implanted with a metal head-fixation post and trained for head-restraint. All whiskers of the mouse except C2 were trimmed

before the recording session. The left C2 barrel column was functionally located using intrinsic optical imaging (Grinvald et al., 1986) through the intact bone (Ferezou et al., 2006). A small craniotomy (<0.5 mm in diameter) was then opened Sodium butyrate to allow for the insertion of the patch pipette within the C2 barrel column. The recording chamber was filled with Kwik-Cast (WPI) to protect the exposed brain and the animal recovered in its cage for 2-4 hr before the recording session began. Electrophysiological recordings, targeted to the C2 barrel column identified by intrinsic optical imaging, were carried out following previously described methods (Crochet and Petersen, 2006, Poulet and Petersen, 2008 and Gentet et al., 2010). The whole-cell recording solution contained (in mM): 135 potassium gluconate, 4 KCl, 10 HEPES, 10 sodium phosphocreatine, 4 MgATP, 0.3 Na3GTP (adjusted to pH 7.