These findings,

These findings, RAD001 taken together with our previous bulk tracing results (Wimmer et al., 2010), indicate that such experience-dependent rewiring of the thalamocortical projection may occur in

as little as 3 days. Rapid receptive field changes in any TC-innervated layer, as recently observed for L5 (Jacob et al., 2012), may partially derive from rapid rewiring of TC anatomy. Given that interbouton distances along axons were unperturbed by trimming, our results indicate a striking reduction in the number of thalamocortical synapses. This reduction was highly unexpected because the sensory responses of single units in L4 are largely regarded as stable, whereas other layers seem robustly plastic (Feldman and Brecht, 2005, Fox, 2002 and Karmarkar and Dan, 2006). We too observed that L4 response magnitudes are

relatively stable. Our results demonstrate that single-unit recordings from a neuronal population do not necessarily allow the inference of anatomical changes among its inputs. One possible explanation is that feedforward inhibition in the thalamocortical circuit maintains L4 responsiveness in the face of TC pruning. Trimming would simultaneously decrease both feedforward excitation and inhibition, possibly leaving L4 response magnitudes unchanged. In this scenario, other functional aspects of cortical activity, beyond the magnitude of sensory-evoked responses, might be plastic. Sensory information may be robustly encoded by near-synchronous discharges of neurons rather than by uncoordinated KPT330 increases in their firing rates (reviewed in Bruno, 2011). For example, the degree of millisecond-timescale synchrony among TC neurons and consequent L4 discharges varies depending on features of whisker stimuli (Bruno and Sakmann, 2006, Temereanca et al., 2008 and Wang et al., 2010). Experience-induced reduction in TC axonal arborization in and of itself would reduce the common input shared by cortical neurons, which

in the simplest case would decrease correlated discharges among L4 neurons during sensory stimulation. Our data show, however, that reduced TC innervation does not guarantee reduced L4 synchrony, indicating that additional elements of the thalamocortical circuit are plastic. aminophylline The loss of afferent input might additionally trigger homeostatic rescaling of the strength of synapses—afferent and/or intracortical—onto an excitatory L4 neuron to maintain its normal firing rate. Consistent with this possibility, we observed that trimming enhances the strengths of common inputs shared by L4 neurons. Synaptic rescaling of intracortical connections within layer 4 is thought to switch off during development but has not yet been studied for thalamocortical connections (Turrigiano, 2011). Reduced TC innervation may directly or indirectly lead to potentiation of unpruned TC synapses.

8 ± 1 0 cells per mouse; KO: 17 4 ± 2 1 cells per mouse; Figure 2

8 ± 1.0 cells per mouse; KO: 17.4 ± 2.1 cells per mouse; Figure 2A) and analyzed units (CT: n = 48; KO: n = 92) with significant activity on the track (place field peak > 1 Hz). Fine quantification PD-0332991 mw revealed no differences in these responses across multiple measures (Figure 2; see also Figure S1). Specifically, single units in KO exhibited normal place field sizes (F(1,138) = 0.01, NS; Figure 2B), normal firing rates within place fields (F(1,138) = 0.56, NS; Figure 2C), no difference in the normal tendency of units to fire more in one direction than another (F(1,138) = 0.19, NS; Figure 2D),

and no difference in sparsity (F(1,138) = 0.85, NS; Figure 2E), which is a measure of the localization of place fields (Jung et al., 1994). In addition, no difference was observed in spatial information index (F(1,138) = 0.02, NS; Figure 2F), which measures how informative about position a spike from a place cell is (Markus et al., 1994), and spatial coherence (F(1,138) = 0.92, NS; Figure 2G), which measures the local smoothness of a firing rate pattern of spikes (Muller and Kubie, 1989). Next, to determine whether excitability might be evident in the precise timing of single spikes, we further examined run-time unit activity

on a finer timescale. Since hippocampal single units exhibit complex LY294002 nmr spikes, made up of a burst of several spikes occurring 2–10 ms apart (Quirk and Wilson, 1999), we first measured the number of spikes during bursts. Both KO and CT units exhibited similar numbers of spikes per burst (F(1,142) = 0.01, NS; Figure 2H) and a similar percentage of burst spikes (F(1,142) = 0.40, NS; Figure 2I). Interestingly, however, we found that

bursts in KO tended to be faster, as measured by burst interspike interval (CT: 5.70 ± 0.70 ms; KO: 4.99 ± 0.78 ms; F(1,142) = 29.16, p < 10−6; Figure 2J), and extracellular spike amplitude attenuation, which is associated with complex spikes (Harris et al., 2001 and Quirk and Wilson, 1999), was also increased in KO (CT: 2.84% ± 0.39%; KO: 5.93% ± 0.38%; F(1,142) = 31.36, p < 10−6; Figure 2J). Taken together, these results indicated that the spatial representation at the level of single science cells in KO appears to be preserved during exploratory behavior, in spite of the bias toward enhanced synaptic strength, with little change in spike timing during bursts. Since the place responses of single units in calcineurin KO were largely normal during run, we next examined whether unit activity during immobile periods, specifically SWRs, was also unaltered. In both KO and CT mice, single units exhibited spikes during SWR events (Figure 3A). Place cells in KO, however, fired more than double the number of spikes during each SWR event as compared to those in CT (CT: 1.11 ± 0.14 spikes per SWR; KO: 2.56 ± 0.54 spikes per SWR; F(1,81) = 4.84, p < 0.05; Figure 3B).

The first transatlantic fiber-optic cable connecting the US and U

The first transatlantic fiber-optic cable connecting the US and UK and allowing 40,000 simultaneous telephone calls was lauded as a communications milestone. The first virus infected the internet, still a largely academic communication medium. The antidepressant Prozac was first introduced to the US market and quickly became one of the top-selling drugs in history. In science, Sir

James W. Black, Gertrude B. selleck screening library Elion, and George H. Hitchings were awarded the Nobel Prize in Physiology and Medicine for their work on “important principles for drug treatment,” developing new drugs targeted for specific biochemical pathways. Fiscal budgets were tight, the NIH was facing cuts to its operating budget, and the scientific community was worried. And amid all of this, the first issue of Neuron was launched in March of 1988 and contained papers on axon branching, channel biophysics, hippocampal LTP, and molecular analyses of gene expression. A lot has changed in the past 25 years for science and the world, but in many ways, the issues that preoccupied us then, as scientists and world citizens, continue to preoccupy us today. Since its inception, Neuron has Selleck Hydroxychloroquine been positioned as a journal that reaches out broadly to the neuroscience community. In the first issue, the

journal’s founders put forth a vision grounded on the pillars of exciting and innovative science, interdisciplinary thinking, the value of basic mechanistic research, and the catalyzing opportunities afforded by technology. In an Editorial in the first issue, the founding editors wrote, “By bringing together papers using methods ranging from biophysics to advanced structural analysis to molecular genetics, the journal can encourage, educate, and sustain a readership of broad technical literacy that shares an interest in common biological questions.” This core vision holds as true today as it did in 1988. This issue is a celebration of the journal and the developments in the field over the past 25 years. We have brought together a series of essays that build on this vision, reflect on the history of the field, and project forward

to the future. One of the most enjoyable and satisfying aspects of compiling an issue like this (and hopefully this applies to reading it as well) is the chance to step back and reflect. Too often in our world, we are busy Rolziracetam looking forward and rarely does one have the chance to admire the full view. In considering topics for this special collection, it was difficult to winnow down the list and capture in full scope the tremendous progress and excitement that we’ve seen in this field over the last few decades. Each Perspective tackles a different subject area in the field, and yet, it is interesting to see some common themes emerge: The importance of interdisciplinary science. The brain is a complex puzzle and no one system or methodology will be sufficient to crack it.

When ICNs are obtained using hierarchical graph clustering of fun

When ICNs are obtained using hierarchical graph clustering of functional correlation networks, they are mathematically congruent with eigenmodes of network

Laplacian (Kondor and Lafferty, 2002). This congruence may explain why dementias appear to fall into distinct ICNs—a strictly mechanical consequence of diffusive network dynamics. Even with possibly random starting configurations, network dynamics are sufficient to produce regional specificity; however, this does not imply that the conventional focal origin or selective vulnerability hypotheses are incorrect. It could be that the starting configuration is dictated by selective vulnerability due to various stressors (Saxena and Caroni, 2011, Palop et al., 2006, Braak et al., 2000 and Seeley et al., 2009), but the subsequent patterns are determined by MS-275 macroscopic network dynamics. Our model merely accommodates a conception of these diseases

that is fully consistent with known findings but does not require focal origin or selective vulnerability. Our model is based on current evidence of prion-like protein misfolding, which propagates within neurons as well as transsynaptically, where retrograde axonal transport deficits cut off the growth-factor supply to projection find more neurons, begetting axonal degeneration, synapse loss, and postsynaptic dendrite retraction. There is mounting neuropathological evidence that numerous disease proteins, including tau, alpha-synuclein, beta-amyloid, and TDP-43, have the capacity to misfold and march through neural circuits via transsynaptic spread ( Palop and Mucke, 2010 and Frost et al., 2009b). If a common concentration-dependent diffusive prion-like process can reproduce subsequent atrophy patterns, this raises a somewhat unorthodox

possibility that diverse degenerative etiologies have common macroscopic consequences. Indeed, our model does not differentiate between individual proteopathic carriers, bunching them together into a generalized old “disease factor.” This is justified on two grounds. First, there is a considerable diversity of published opinion on the etiology of neurodegeneration (Saxena and Caroni, 2011) and the effect of individual misfolding proteins (Whitwell et al., 2009, Palop and Mucke, 2010 and Frost et al., 2009b). Second, and more important, the specific biochemical properties of the prion-like agent may be inconsequential for the macroscopic and chronic manifestation of disease, as evidenced from recent joint histopathological/morphometric studies. The idea that proteopathic carriers with varied etiology can have a shared progression mechanism via “permissible templating” was first raised by Hardy (2005). The spatial distribution of beta amyloid pathology in AD is poorly correlated with whole brain atrophy patterns ( Rabinovici et al., 2010), while tau is well-correlated.

, 2003)

The similarity between spontaneous and evoked pa

, 2003).

The similarity between spontaneous and evoked patterns is not restricted only to global activity patterns but has also been found in spike-timing relations among neurons. At the microcircuit level, the precise temporal sequence of spiking evoked by external stimuli is more similar to spontaneously occurring patterns than predicted by chance. This has been demonstrated both in vitro (MacLean et al., 2005) and in vivo (Luczak et al., 2009). These data suggest that the adaptation of ongoing activity to the statistical nature PF-06463922 cost of experienced stimuli can also involve sculpting the corresponding microcircuit architecture (Luczak and Maclean, 2012). Other data from freely moving animals suggest that such changes in sequential spiking are related to behaviorally relevant learning and memory processes. AZD9291 supplier Population recordings in hippocampus or neocortex have revealed that spiking sequences observed during behavior were subsequently replayed in similar temporal order during following resting periods (Euston et al., 2007, Ji and Wilson, 2007 and Skaggs

and McNaughton, 1996). Despite the likely importance of understanding the mechanisms by which stimulus-evoked sequences are “imprinted” in spontaneous activity, advances have been limited by the technological difficulty of recording neuronal population activity and manipulating neural processes in behaving animals. The hallmark of memory formation in the brain activity of freely over moving animals is the emergence of stimulus-induced

(or behavior-induced) sequential activity patterns that are later spontaneously replayed (Euston et al., 2007, Ji and Wilson, 2007 and Skaggs and McNaughton, 1996). Although many previous studies have emphasized replay during slow-wave sleep, there is abundant evidence that it can occur during periods of wakeful quiescence, even relatively brief ones, when the hippocampus exhibits large irregular activity containing sharp wave ripple (SPWR) events and the cortex is in a relatively synchronized state, exhibiting up-down state transitions. Moreover, the actual reactivation events occur during the up states, which can be considered as brief episodes of cortical desynchronization. Finally, there is also evidence that long-term potentiation (LTP) is suppressed during slow-wave sleep in general (Leonard et al., 1987) but is transiently re-enabled during SPWR events that are associated with neocortical up-state transitions (Buzsáki, 1984). To investigate if a similar phenomenon could be also studied in simpler (anesthetized) preparations and to study how the formation of sequential patterns depends on the brain state, we used population recordings in urethane-anesthetized rats.

The cDNA was amplified with forward, tcttgcaaggacagttgcac, and re

The cDNA was amplified with forward, tcttgcaaggacagttgcac, and reverse, tcaggcacttgtctcacagc, primers bracketing nucleotides 248–629 in the open reading frame of chicken prestin (GenBank accession number EF028087.1; Schaechinger and Oliver 2007) using Invitrogen Platinum Taq DNA Polymerase (Life Technologies). As a positive control, 200 ng/μl of pEGFP-N1 plasmid containing chicken prestin (Schaechinger and Oliver, 2007; Tan et al., 2011) click here was also amplified with the same primer set. PCR products were electrophoresed on 1% agarose gel. A polyclonal antibody against the N-terminal peptide sequence of rat prestin, CKYLVERPIFSHPVLQE (Bethyl

Laboratories), which is closely homologous to the equivalent sequence in chicken prestin, was used to label the basilar papilla. The antibody was affinity purified and shown to immunolabel rat OHCs (Mahendrasingam et al., 2010). Immunoblots were performed on tissue from fifteen E19 chicken papillae and twelve P11 mice cochleas; mice were decapitated and cochleas dissected out using procedures approved by the Institutional Animal Care and Use Committee of the University of Wisconsin-Madison. Proteins were extracted with Tissue Extraction Reagent I (Invitrogen Life Technologies) plus protease inhibitor C59 wnt solubility dmso cocktail (Sigma-Aldrich), denatured and electrophoretically separated

on a 7.5% SDS-PAGE and blotted onto a 0.45 μm nitrocellulose membrane. Blotted membranes were incubated with the prestin antibody (1:100 dilution) at 4°C overnight then incubated in secondary goat anti-rabbit horseradish Calpain peroxidase-conjugated antibody (1:1000, Invitrogen) for 90 min at room temperature and stained with Novex chemiluminescent reagent. Chicken papillae were fixed in 4% paraformaldehyde in phosphate buffered saline (PBS) for 30 min, washed then permeabilized with 0.5% Triton-X for 30 min. Fixed papillae were immersed in 10% goat serum (Invitrogen) for 1 hr at room temperature and incubated overnight at 4°C with

the prestin antibody (dilution 1:50) and the mouse monoclonal HCS-1 antibody (dilution 1:400) which labels otoferlin in chicken hair cells (Goodyear et al., 2010). After rinsing in PBS, specimens were incubated with Alexa Fluor 488 goat anti-rabbit IgG antibody (1:200; Invitrogen) and goat anti-mouse Alexa Fluor 568 for 90 min and Alexa Fluor 647 phalloidin (1:200; Invitrogen Life Sciences) for 60 min at room temperature. Preparations were mounted in Fluoromount-G medium (SouthernBiotech) with coverslips and viewed under a 60× oil-immersion objective (NA = 1.4) on a Nikon A1 laser scanning confocal microscope. Work was supported by grant RO1 DC01362 from the National Institutes on Deafness and other Communication Disorders to R.F. We thank Dan Yee for constructing electrical equipment, Ana Garic for assistance with molecular biology, Lance Rodenkirch for advice on confocal imaging, Dominik Oliver for the pEGFP-N1 plasmid containing chicken prestin, and Jeff Corwin for the HCS-1 antibody.

Further, these regions respond during mental imagery of big and s

Further, these regions respond during mental imagery of big and small objects, which is a characteristic property of other nearby category-selective regions. Finally,

we find that these regions reflect information about the category of the object rather than how big the object was conceived. Broadly, these results show that real-world size this website is a large-scale dimension that differentiates distributed object representations in occipitotemporal cortex. We propose a potential account of this organization, in which the size of objects in the world naturally give rise to systematic biases in visual experience which are extracted in early visual areas and ultimately dictate where high-level object representations will be in anterior occipitotemporal cortex. In Experiment 1a, observers were presented with images Epacadostat of isolated big objects (e.g., car, piano) and isolated small objects (e.g., strawberry, safety pin), presented at the same retinal size on the screen (Figure 1A; for all stimuli see Figure S1 available online; see Experimental Procedures). The experiment consisted of one run of 8.8 min

of scanning, during which 200 distinct big objects and 200 distinct small objects were presented in a standard blocked design (see Experimental Procedures). To compare the neural response of big and small objects, we conducted a size-preference map analysis and a whole-brain contrast analysis. In the first analysis, we visualized the spatial distribution

of small and big object preferences across occipitotemporal cortex. Size-preference maps were computed reflecting whether the voxels had a preference for big objects (blue) or small objects (orange) within an object-responsive mask (see Experimental Procedures), and these are shown on an inflated cortical surface in Figure 1. We observed a striking large-scale organization along the ventral surface, evident at the group level and at the single-subject level, with big and small object preferences arranged in a medial to lateral organization across both hemispheres. Further, not this organization was mirrored along the lateral surface of the cortex, with small to big object preferences arranged from inferior to superior (Figures 1B and 1C). Importantly, these data should not be interpreted as evidence that big and small objects are represented in separate swaths of cortex. Both big and small objects activate most of this object-responsive cortex to varying degrees, illustrated in Figures 2, consistent with accounts of distributed activation profiles of these objects (e.g., Haxby et al., 2001). However, voxels with a big-object preference are consistently found along medial ventral temporal cortex, while voxels with a small-object preference were consistently found along lateral temporal cortex (Figures 2C and 2D).

, 2008 and Zenisek et al , 2000]) In WT astrocytes (data not sho

, 2008 and Zenisek et al., 2000]). In WT astrocytes (data not shown) and in Tnf−/− astrocytes incubated with TNFα, ( Figure 4E) the two pools underwent exocytosis in a clear biphasic temporal sequence: during the first phase (0–400 ms) most of the fusing vesicles belonged to the “resident” pool (80.6%, n = 7 cells), whereas during the second phase (500 ms–2 s), to the “newcomers” pool (82.5%). This temporal segregation reflects the different readiness to fusion of the two pools, in particular the fact that most “resident” vesicles, contrary to “newcomers,”

have already undergone the docking steps and are ready for fusion (i.e., are functionally docked [ Ohara-Imaizumi et al., 2007 and Toonen et al., 2006]). However, in Tnf−/− Bortezomib astrocytes,

the situation was very different. Events attributable to “residents” decreased in percentage (20% instead of 40%; n = 3680 vesicle fusions analyzed, n = 7 cells). Moreover, importantly, events due to “residents” IDH inhibitor and “newcomers” occurred randomly, without the expected temporal segregation. This indicates that even the residual “resident” pool seen in Figure 4A is defective in Tnf−/− astrocytes, because it is not ready/competent to fuse. Most likely, these vesicles dock only transiently and, like all the others, most in the absence of TNFα are hampered in reaching the stage of functional docking

allowing them to undergo rapid fusion ( Toonen et al., 2006). We conclude that constitutive TNFα is necessary for the correct reception of glutamatergic vesicles to release sites, a precondition for efficient exocytosis upon stimulation. In parallel TIRF experiments, we studied local submembrane Ca2+ events, previously shown to be temporally locked to exocytic events (Marchaland et al., 2008). Indeed, in WT astrocytes, 2MeSADP stimulation induced a burst of submembrane Ca2+ events whose temporal pattern mirrored the one of VGLUT1-pHluorin fusion events, with two peaks of Ca2+ events, each one slightly preceding the corresponding peak of vesicular fusions (Figure 4B, inset). Importantly, and in full agreement with the observations in situ, this pattern was totally preserved in Tnf−/− astrocytes ( Figure 4C, inset), further confirming that TNFα does not act on the coupling between GPCR and [Ca2+]i elevation, and indicating that this step of gliotransmission can be perfectly normal while the downstream signaling is dramatically defective.

These complications should be borne in mind in considering whethe

These complications should be borne in mind in considering whether different repeat sizes within the C9ORF72 gene may provide divergent symptoms/diseases or different severity

of phenotypes. A gain-of-RNA-toxicity mechanism for a repeat expansion disease is best characterized in myotonic dystrophy 1 (DM1), which is caused by up to 2,500 of CTG repeats in the 3′UTR of the myotonic dystrophy protein kinase (DMPK) gene (Lee and Cooper, 2009). Two proteins, CUG-BP1 and muscleblind, were identified selleck chemicals llc to bind to the CUG repeat-containing RNA (Miller et al., 2000 and Timchenko et al., 1996). Of these two proteins, only muscleblind shows repeat-length-dependent association and is selectively sequestered into pathogenic RNA foci (Mankodi et al., 2001). Nevertheless, misregulation of both muscleblind and CUG-BP1 play roles in DM1 pathogenesis. Indeed, CUG repeats lead to activation of protein kinase C (PKC), which in turn phosphorylates CUG-BP1, whose phosphorylated form has increased activity from increased protein stability, thereby activating multiple splicing changes toward fetal isoforms (Kuyumcu-Martinez et al., 2007 and Roberts et al., 1997). The function of the C9ORF72 gene and its predicted protein product are unknown. Recent bioinfomatical analysis implies a potential involvement of the C9ORF72

protein in membrane trafficking and autophagy ( Levine et al., 2013 and Zhang et al., 2012), but this remains to be determined. A 50% reduction of mRNA levels corresponding to both short and long mRNA isoforms of C9ORF72 ( DeJesus-Hernandez et al., 2011 and Gijselinck et al., 2012) has been reported and GSI-IX nmr much is consistent with partial or complete silencing of the expanded allele ( Figure 4A), although it should be noted that the reduction of the corresponding C9ORF72 proteins has not been demonstrated. Antisense oligonucleotide-mediated reduction of C9ORF72

in zebrafish produces reduced axon lengths of motor neurons and locomotion deficit ( Ciura et al., 2013), consistent with the notion that partial loss of the C9ORF72 gene could contribute to disease pathogenesis. Intranuclear RNA foci containing the C9ORF72 hexanucleotide repeat have been reported (DeJesus-Hernandez et al., 2011), which may trap one or more RNA-binding proteins, thereby inhibiting their functions, especially in RNA processing (Figure 4B). While two RNA-binding proteins, hnRNP-A3 (Mori et al., 2013a) and Pur-α (Xu et al., 2013), have been reported to bind GGGGCC repeats in vitro and both were reported to be components of p62-positive TDP-43-negative inclusions in C9ORF72 patients, their role in pathogenesis is unproven. Neither has been demonstrated to localize at RNA foci formed by the hexanucleotide repeat and the predicted loss of RNA processing function that would follow from sequestration of hnRNP-A3 and Pur-α has not been demonstrated in cells and tissues expressing the hexanucleotide repeat-containing RNA.

Another important finding is that VGLUT3 expression suffices for

Another important finding is that VGLUT3 expression suffices for the induction of vesicular glutamate uptake and release in nonglutamatergic neurons. While Imatinib supplier recent data from VGLUT3-deficient neurons demonstrated the necessity of VGLUT3 function for glutamatergic neurotransmission in auditory hair cells and pain pathways (Obholzer et al.,

2008, Ruel et al., 2008, Seal et al., 2008 and Seal et al., 2009), our demonstration that this effect is a direct result of VGLUT3′s ability to function as a classical vesicular transporter is critical in interpreting morphological data showing that VGLUT3 localizes to terminals from serotonergic, cholinergic, and GABAergic neurons. Based on our findings, these synapses IPI 145 are very likely coreleasing glutamate along

with their classical neurotransmitters, which implies a fast excitatory signaling component at these classically modulatory synapses. Previous studies have shown that VGLUT levels are endogenously and bidirectionally regulated during development (Boulland et al., 2004 and Nakamura et al., 2005), in disease states (Eastwood and Harrison, 2005, Kashani et al., 2007 and Smith et al., 2001), with pharmacological manipulation (De Gois et al., 2005 and Wilson et al., 2005), and according to circadian rhythms (Yelamanchili et al., 2006). Our data suggest these alterations would be accompanied by changes in neuronal firing patterns and perhaps circuit behavior. For example, Ribonucleotide reductase differences between VGLUT1 and VGLUT2/3 could be important during development, where the early, transient expression of VGLUT2 and VGLUT3 in neurons that later express VGLUT1 could increase the chance of glutamate release at synapses that may contain fewer synaptic vesicles

than mature synapses. It is possible that neurons or networks of neurons actively use specific VGLUT isoform expression to regulate the efficiency of glutamate release. The mechanism by which endophilin levels regulate release efficiency is still unknown. Because endophilin is a protein known primarily for its role in endocytosis, it is possible that it acts by altering either the size of the RRP or its rate of replenishment. However, overexpression and knockdown of endophilin did not affect the RRP. Instead they increased and decreased the EPSC charge, suggesting that endophilin directly alters the fusion efficiency of synaptic vesicles. Because this effect does not require the SH3 domain, it is not likely to involve increased recruitment of dynamin or synaptojanin. The effect is, however, dependent on membrane binding and dimerization. Although it is likely that many of endophilin’s actions are dependent on interactions with synaptojanin and dynamin, recent evidence suggests endophilin’s main endocytic function requires only the BAR domain and occurs at the plasma membrane prior to vesicle scission (Bai et al., 2010).