Then, IVs were fitted with a cubic, and the zero crossing (Nernst

Then, IVs were fitted with a cubic, and the zero crossing (Nernst potential) was determined analytically. Residues of the alignment in Figure 1B were colored with Jalview 2 (Waterhouse et al., 2009) in modified Zappo color scheme (hydrophobic I, L, V, A, and M = pink; aromatic F, W, and Y = orange; positively charged K, R, and H = red; negatively charged D and E = blue; hydrophilic S, T, N, and Q = green;

P and G = magenta; C = yellow). Values are reported as mean ± SEM. We would like to thank H. Okada and W. Chu for help with the cloning and the members of the Isacoff lab for discussion. This work was supported by postdoctoral fellowships for prospective and advanced researchers from the Swiss National learn more Science Foundation (SNSF; PBELP3-127855 and PA00P3_134163) (T.K.B.) Navitoclax and by a grant from the National Institutes of Health (R01 NS35549) (E.Y.I.). “
“Cortical circuits display fine functional and structural organization (Feldmeyer et al., 2002, Lefort et al., 2009 and Petreanu et al., 2009) that is carefully established and tuned by sensory experience (Bender et al., 2003, Buonomano and Merzenich, 1998, Feldman and Brecht, 2005 and Stern et al., 2001). Modification of synapses includes Hebbian plasticity

mechanisms where correlated (or uncorrelated) activity leads to structural as well as functional alternations, such as changes in spine morphology (Alvarez and Sabatini, 2007), or synaptic insertion or removal of AMPA receptors (Kessels and Malinow, 2009, Malenka and Bear, 2004, Newpher and Ehlers, 2008 and Nicoll et al., 2006). In parallel to such Hebbian

mechanisms, neurons are also equipped with homeostatic-scaling machinery that may serve to avoid instability problems of network activity (Turrigiano and Nelson, 2004). Such scaling can globally regulate synaptic strength by altering the number of AMPA receptors in individual synapses (Turrigiano et al., 1998). Although a number of molecular and cellular mechanisms underlying these plasticity mechanisms have been identified, how synapses on a dendritic branch cooperate with each other to drive such plasticity is not well understood. Accumulating in vitro and theoretical evidence suggests that there exists biochemical compartmentalization on dendrites that leads to clustered synaptic plasticity (Branco Tyrosine-protein kinase BLK and Häusser, 2010, Govindarajan et al., 2006, Häusser and Mel, 2003, Iannella and Tanaka, 2006 and Larkum and Nevian, 2008). For example NMDA receptor-dependent Ca2+ influx caused by a dendritic spike (Golding et al., 2002, Schiller et al., 2000 and Wei et al., 2001), spread of Ras activity during long-term potentiation (LTP) (Harvey et al., 2008), and exocytosis of AMPA receptors into dendritic membrane during LTP (Lin et al., 2009, Makino and Malinow, 2009, Patterson et al., 2010 and Petrini et al., 2009) all occur locally on short stretches of a dendrite and could contribute to synaptic potentiation at nearby synapses.

(2013) makes a strong case for excess glutamate to be a pathogeni

(2013) makes a strong case for excess glutamate to be a pathogenic click here driver, use of antipsychotic drugs may hasten the transition to psychosis and, therefore, should be avoided during the prodromal stage of the illness. Prevention strategies are now commonplace in most fields of medicine. This includes neurology, where dietary and other lifestyle recommendations have become mainstream for prevention or reduction of

illness progression for brain disorders such as Alzheimer’s disease. Early identification of individuals who are at high risk to develop schizophrenia holds the promise of advancing preventive treatment to psychiatry, and in particular to prevention of schizophrenia. Research so far indicates that pathophysiological changes in the prodrome and first-episode patients may be distinct from that observed in chronic patients (Kaur and Cadenhead, 2010) and that early intervention with traditional antipsychotic drugs may signaling pathway not only be ineffective, but may actually worsen the outcome. This makes longitudinal and mechanistically oriented translational approaches such as that applied by Schobel et al. (2013) especially critical for the design of prevention strategies to prevent psychosis in individuals

at high clinical risk for schizophrenia. “
“For years, it was routine that each new generation of microtubule researchers would learn to make tubulin preps from bovine or porcine brain (Miller and Wilson, 2010). This would involve regular trips to a slaughterhouse, waiting for the brains to become available so that they could

immediately be put on ice and rushed back to the lab, minced, and then put into a blender with cold buffer. From there, the strategy was based on the simple principle that microtubules would disassemble into soluble tubulin found in the cold and reassemble into microtubules when the prep was warmed. Through cycles of warming and cooling, the prep would become progressively more enriched for tubulin. For most microtubule labs, those days are gone, because molecular approaches can now accomplish what used to require this tedious procedure. Reminiscing about those earlier days brings to mind a fundamental issue about tubulin and microtubules. The initial huge pellet produced by spinning down the brain homogenate was normally washed down the drain without further consideration—but this pellet actually contained a significant amount of tubulin that was not soluble in the cold. Although this was long known, the cold-stable tubulin fraction was given little attention by most investigators—one notable exception being Scott Brady. Nearly three decades ago, Brady et al. (1984) compared the biochemical properties of tubulin in the cold-stable fraction with the properties of the temperature-cycled tubulin. Much of the tubulin in the cold-stable fraction was shown to be extremely basic in charge, as assessed by two-dimensional electrophoresis.

4% of the total variance in force The first factor consisted of

4% of the total variance in force. The first factor consisted of body mass, muscle circumferences, and skinfolds accounting for 47.8% of the variance in force. In contrast, the third factor included height and limb lengths and accounted for only 7% of the total variance in force. The regression analysis for males in the present study is consistent with the findings of Scanlan et al.20 because a circumference measure (ELB) had a greater impact on the equation for predicting elbow flexion strength than a length measure (L3). In contrast, the inclusion of L3 to a prediction equation with BW had a greater impact for females than it did for males, in terms of accounting for

additional variance in elbow flexion strength. The large contribution of limb length to the strength prediction equation for females may be explained by the relationship between the length of a muscle and the number of sarcomeres in series.21 and 22 The number of Tanespimycin solubility dmso sarcomeres in parallel (physiological cross-sectional area) is proportional to the amount of tension that is produced whereas the number of sarcomeres in series (muscle fiber length) is proportional to the velocity at which tension is.23 and 24 While

the dependent measure in this study was mean torque, not velocity of shortening, it has been suggested that selleck products the number of sarcomeres in series, and therefore the length of a muscle, has a relationship with the amount of force being produced.22 and 25 over This relationship was demonstrated for sprint performance and leg characteristics in female sprinters. Abe and colleagues26 found that increased fascicle length was highly correlated with increased shortening velocity and concurrently, sprint performance. These physiological characteristics combined with females’ decreased proportion of lean tissue mass may explain the large contribution of limb length compared to weight and circumference measurements.

The contribution of muscle activation in addition to muscle size to the prediction of strength was assessed by incorporating RMS sEMG amplitude to equations consisting of BW and a second anthropometric variable. The addition of sEMG RMS resulted in a significant (p < 0.05) increase in the variance-accounted-for by each equation, except when the second variable was L3 for females. The minimal contribution may have been due to the immense contribution of L3 alone (partial R2 = 39.1%). Excluding this particular case, on average, sEMG RMS accounted for an additional 10.1% of the variance in strength. Surprisingly, the addition of a third anthropometric variable instead of sEMG RMS resulted in superior prediction equations for both males and females. The majority of the literature on force and sEMG is focused on the linear versus non-linear nature of the relationship, to create a calibrating equation throughout the range of muscle forces (0–100% maximal voluntary contraction).

, 2002, Verdi et al , 1999 and Zhong et al , 2000) and is also in

, 2002, Verdi et al., 1999 and Zhong et al., 2000) and is also involved in promoting neurogenesis and differentiation ( Klein et al., 2004, Wakamatsu et al., 1999 and Zilian et al., 2001). Numb is a membrane-associated adaptor protein containing C59 wnt multiple protein-binding motifs, and numb associates with a number of other proteins. Via its specific binding to AP-2 and other endocytic proteins (such as Eps15 and EHD4), numb plays roles in regulated endocytosis of receptors, including those of the Notch, integrin, L1, and Trk families. Many of the biological effects of numb on neurodevelopment are therefore probably related to regulating endocytosis of specific

receptors (Santolini et al., 2000). Interaction of numb with integrin promotes endocytosis and directional cell migration (Nishimura and Kaibuchi, 2007). Par3-dependent phosphorylation of numb by aPKC regulates polarized localization of numb and its association with clathrin coated structures (Smith et al., 2007). Negative regulation of numb by aPKC appears to play an important

role in integrin endocytosis and integrin-based cell migration (Nishimura and Kaibuchi, 2007). Numb has also been reported to play 3-Methyladenine nmr a critical role in cerebellar granule cell polarization during migration via a distinct mechanism (Zhou et al., 2011). Numb interacts with activated TrkB and promotes TrkB endocytosis and polarization. BDNF-induced phosphorylation of numb by aPKC increases binding to TrkB and promotes chemotactic responses to BDNF. Thus, in the

context of TrkB endocytosis, phosphorylation by aPKC increases affinity of numb for its cargo. In addition to its role in mediating receptor internalization, numb has a novel function in endosomal recycling of receptors. For example, numb has been implicated in regulating the postendocytic trafficking of Notch1. In mammalian cell lines, overexpression of numb promotes trafficking and degradation of Notch 1, whereas depletion of numb facilitates recycling of Notch1. Numb mutants defective Digestive enzyme in binding to endocytic proteins such as α-adaptin, fail to promote Notch 1 degradation, suggesting numb suppresses Notch activity by regulating postendocytic sorting pathways that lead to Notch degradation (McGill et al., 2009 and McGill and McGlade, 2003). Neurons frequently show cell-type-specific responses to the same environment. One mechanism of cell-type specificity involves differential expression of adhesion and guidance receptors (Kolodkin and Tessier-Lavigne, 2011 and Stein and Tessier-Lavigne, 2001). Interestingly, neurons can have different responses even when expressing the same receptors. Subsets of cortical neurons, identified by differential expression of the transcription factors Satb2 or Ctip2 respond differentially to Sema3A cues even though both populations similarly express the Neuropilin1, L1, and plexinA4 receptors (Carcea et al., 2010).

To address this we quantified the bulk movement of the entire pho

To address this we quantified the bulk movement of the entire photoactivated pool by using intensity-center shift analysis (Figure 2A). Briefly, the intensity center in a given frame is a quantitative center of the distribution of binned fluorescence intensities along a line scan within the photoactivated zone.

A bulk vectorial movement of fluorescent molecules within the axon would lead to a corresponding shift in the intensity center as well. Anterograde shifts were consistently seen in the photoactivated pools of both synapsin and CamKIIa though there were differences in the overall kinetics IWR-1 clinical trial as shown in Figure 2B. Specifically, for CamKIIa, there was a slight lag in the initiation of intensity-center shift as well as a periodic variation that was more pronounced than that for synapsin. Note that there is an expected flattening of the wave after the initial rise as the

fluorescent molecules leave the analyzed area (photoactivated zone) over time. Despite these differences in the overall nature of the intensity-center Nutlin-3 molecular weight shifts between synapsin and CamKIIa, there were similarities in their initial intensity shifts, suggesting possible commonalities between mechanisms transporting these cytosolic proteins—a notion supported by radiolabeling studies as well (Garner and Lasek, 1982). To investigate this in more detail we focused on the initial intensity shift, imaging the activated zone with higher time compression. Indeed there was some similarity in the overall transport behaviors of both proteins (Figure 2B, lower panel). Linear regression slopes were 0.008 and 0.01 respectively, equivalent to predicted average rates of 0.008–0.01 μm/s for the entire population. These data are comparable to known slow rates of synapsin and CamKIIa from in vivo radiolabeling experiments, ≈0.01–0.03 μm/s (Baitinger and Willard, 1987, Lund and McQuarrie, 2001 and Petrucci et al., 1991), providing confidence in the validity of this assay in evaluating axonal transport of these cytosolic cargoes. Note

that similar analysis of untagged soluble PAGFP diffusion does not show any bias in its mobility (Figure 2B, lower right panel). The anterogradely Phosphatidylethanolamine N-methyltransferase biased transit of synapsin and CamKIIa, distinct from the bidirectional unbiased movement of untagged soluble PAGFP in axons, suggests that the movement is not a simple diffusive process. We further tested this notion by analyzing synapsin and CamKIIa transport in the presence of micromolar amounts of the alkylating agent N-ethyl maleimide (NEM), a known inhibitor of kinesins, dyneins, and myosins (Pfister et al., 1989). Most recently, NEM was used to inhibit molecular motors to dissect their role in mobilizing vesicles at synapses (Shakiryanova et al.

, 2012, Hasselmo and Wyble, 1997, Lisman and Grace, 2005 and Yass

, 2012, Hasselmo and Wyble, 1997, Lisman and Grace, 2005 and Yassa and Stark, 2011). The VE-821 ic50 hippocampus may play a similar role in perception, tracking the strength of relational match/mismatch. These findings suggest that the hippocampus does not generally produce a state-based signal in long-term memory, but may produce state- or strength-based signals depending on the nature of the materials and demands of the task. In the current perception

study, we found a linearly graded signal from the hippocampus, which may be a result of complex, feature-ambiguous materials and/or a graded comparison process. The critical point is that it is necessary to assess state- and strength-based memory and perception to elucidate the role of the hippocampus

in these cognitive domains. Further studies examining the conditions in which the hippocampus produces state-based or strength-based output will be important. The current neuroimaging and patient findings converge to indicate that the hippocampus is involved in, and is necessary for, perceptual judgments of scenes, and this role is specific to perceptual judgments based on continuously graded strength signals. Scene perception based on discrete states of Akt inhibitor identifying specific differences does not seem to depend on the hippocampus. The findings highlight the surprising reach of the hippocampus, affording precision in both memory and perception. Both studies were approved by the University of California, Davis Institutional Review Board. Informed consent was obtained from all individuals prior to their participation. Mean age of the patients was 49.2 years (SD = 14.1) and mean education was 14.8 years (SD = 2.7). Mean age

of the controls was 47.7 years (SD = 15.6) and mean education was 15.2 years (SD = 2.0). Patients and controls were not significantly Activator different in age or education (t’s < 1). Each patient had 1–3 controls that were closely matched to the patient’s age and education. Patients. Patient characteristics and neuropsychological scores are shown in Table 1. Patient 1 had selective hippocampal damage following a traumatic brain injury due to a car accident. Clinical scans appeared normal with the exception of volume reductions in the hippocampus. Table 2 provides estimates of gray matter volume for MTL structures for this patient and age-matched controls. The left and right hippocampus were significantly reduced in volume for the patient compared to controls; no other MTL structure showed significant volume reduction ( Figure 1). Patient 2 had limbic encephalitis, and MRI scans suggested damage to the hippocampus bilaterally, with no damage apparent in the surrounding parahippocampal gyrus (Figure 1).

Dimer formation was enhanced by oxidizing conditions (100 μM CuPh

Dimer formation was enhanced by oxidizing conditions (100 μM CuPhen) and eliminated by treatment with reducing agent (100 mM DTT). To test if the crosslinking of subunits in GluA2-A665C was specific to functional receptors, i.e., those that could be controlled by iGluR ligands, we tested for dimerization in various conditions. A substantial dimer fraction was observed in the presence of 500 μM glutamate (30% ± 4%; n = 11 blots). This dimerization was specific to the introduction of cysteine at position 665 because the nearby R661C mutant, which exhibited minimal inhibition in electrophysiological assays, showed indistinguishable

dimer formation from the background in the same conditions (14% ± 2%; n = 5; p = 0.99 versus GluA2 7 × Cys

(−); Dunnett’s post hoc test; Figures 3D and S3D). Dimerization of the A665C mutant www.selleckchem.com/products/MLN8237.html Sunitinib research buy was reduced to control levels by crosslinking in the presence of 10 mM glutamate (14% ± 3%; n = 6 blots; p = 0.026 versus 500 μM glutamate; both with 100 μM cyclothiazide [CTZ]; Figure 3D). Inclusion of DNQX and CTZ produced a level of dimerization in between that of control (R661C) and A665C with 500 μM glutamate, but the difference from either was not significant (n = 5 blots, p = 0.41 versus A665C; p = 0.82 versus R661C; Dunnett’s post hoc test). Our structural, biochemical, and electrophysiological findings suggest that the LBD assembly can adopt a distinct CA conformation that occurs readily in full-length receptors. The CA conformation might be unstable in full-length receptors, but the crosslinked LBD tetramer structure is stabilized by an intersubunit disulfide bond. The absence of the ATD and TMD perhaps also allows the LBDs to adopt this configuration unhindered. What then are the expected consequences of the CA conformation in full-length channels? To investigate whether OA-to-CA transitions

in an intact receptor would require rearrangements of the ATD tetramer conformation, we measured the distance between the Cα atoms of T394 (lobe 1 of the LBD, proximal to the ATD). In an OA-to-CA transition, the pairwise intersubunit distances would likely either decrease or stay about the same (Table S1). Thus, consistent with a minimal role for ATD transitions in gating, OA-to-CA transitions are predicted to not disrupt the conformation of the ATD layer observed in the full-length Endonuclease receptor structure. One measure of the extent to which the four LBDs provide impetus to gate the channel is the distances between LBD segments proximal to the TMD, i.e., the Cα atoms of P632, for each pair of subunits (Lau and Roux, 2011 and Sobolevsky et al., 2009). We examined these distances in the crosslinked LBD tetramer structure, the full-length GluA2 structure, and several modeled conformations of the LBDs (Table S1). The analysis indicates that the CA conformation results in greater P632-P632 distances relative to the OA conformation.

, 2005), although it has also been suggested that this “anticorre

, 2005), although it has also been suggested that this “anticorrelation” may reflect a statistical

artifact ( Murphy et al., 2009; Anderson et al., 2011). Given the proposal that the neural systems mediating Cilengitide purchase attention and memory are anatomically segregated, and perhaps even in opposition, it is unclear what neural systems are involved when visual attention is recruited during episodic retrieval. Does the recruitment of visual attention by episodic retrieval engage the same brain regions implicated in top-down visual attention in the perceptual domain (dorsal attention network), brain regions associated with episodic retrieval (default network), or both? In the experiment described here, we directly investigated the recruitment of visual attention during episodic retrieval. Specifically, we dissociated attention

to specific perceptual detail and successful retrieval of specific perceptual detail. We accomplished this goal selleck screening library using a paradigm we recently developed that shows that gist-based false recognition, which occurs when one mistakenly recognizes an item that is similar to an item that was previously encountered ( Reyna and Brainerd, 1995; Koutstaal and Schacter, 1997), occurs primarily because of a failure to retrieve detailed information that is still stored in memory ( Guerin et al., 2012). Critically, our data established that attention to the specific perceptual details relevant to the task is not sufficient to overcome this failure. Rather, reinstatement of the studied item, a potent cue that enables participants to retrieve diagnostic details from memory, Carnitine dehydrogenase is required to substantially reduce gist-based false recognition. Thus, attention to specific perceptual details can occur in the absence of successful retrieval of task-relevant perceptual

details. In addition to shedding light on the mechanisms leading to memory distortion, this experimental paradigm also enables us to isolate and directly investigate the recruitment of visual attention during episodic retrieval. The experimental paradigm is illustrated in Figure 1. Participants study a series of pictures. Then, they undergo a memory test while brain activity is indirectly measured with fMRI. On each trial of the recognition test, participants are presented with three pictures. Their task is to select one of the pictures as a previously studied item or reject all three items as novel. Note that the task is not a forced-choice recognition task: on some trials, no target is presented and the correct response is to reject all three items as new. In contrast to standard yes/no recognition, in the present task participants are switching their attention between test items over the course of the trial.

The thalamocortical brain slice preparation allows TC input to ba

The thalamocortical brain slice preparation allows TC input to barrel cortex to be selectively activated by extracellular stimulation in VPM and resulting synaptic responses to be monitored with extracellular or patch-clamp recordings (Agmon and Connors, 1991, Crair and Malenka, 1995 and Isaac et al., 1997). Extracellular field potential recordings were made to measure TC fEPSPs evoked by electrical

stimulation in VPM. TC inputs are glutamatergic, with the fEPSP mediated by AMPARs (Agmon and O’Dowd, 1992, Crair and Malenka, 1995, Kidd and Isaac, 1999 and Lu et al., 2001). Consistent with selleck chemicals this and previous work (Agmon and Connors, 1992 and Crair and Malenka, 1995), the fEPSP was reversibly blocked by 10 μM NBQX, an AMPAR antagonist, or a Ca2+-free extracellular solution (Figure S5). These manipulations did not block the small early downward deflection confirming that this small deflection is a presynaptic fiber volley. The strength of the TC input to layer 4 (contralateral to the intact whisker-pad) selleck products in slices prepared from sham or IO rats was compared by measuring the fEPSP: fiber volley (FV) ratio at different stimulus

intensities (Figure 5). This input/output (I/O) relationship was significantly steeper in slices from IO rats compared to sham, demonstrating an increase in TC input strength in the spared input side following IO nerve resection. There was a 47% increase in TC synaptic strength in the IO rats compared to sham. To examine whether intracortical (IC) synapses in L4 barrel cortex are strengthened following IO nerve resection, in a separate set of experiments we measured TC fEPSPs and IC fEPSPs in layer 4 (Figure S6). We confirmed the increase in the input/output relationship for TC fEPSPs but found no increase in the input/output relationship for IC fEPSPs in slices Doxorubicin from IO rats.

Thus, intracortical synaptic strength in layer 4 is not increased in spared barrel cortex in IO rats, indicating strengthening of TC synapses. The mechanism(s) underlying the increase in the TC fEPSP in the spared barrel cortex were studied with patch-clamp recordings. GABAergic feedforward inhibition in L4 barrel cortex is strongly engaged by TC afferent activity and serves to regulate coincidence detection, truncate the EPSP, and limit spike output in L4 (Chittajallu and Isaac, 2010, Cruikshank et al., 2007, Daw et al., 2007a, Gabernet et al., 2005 and Porter et al., 2001). A change in the engagement of feedforward inhibition could contribute to the change of the TC fEPSP observed in the IO rats. Whole-cell voltage-clamp recordings from L4 stellate cells were performed to measure the feedforward inhibition and feedforward excitation onto the same stellate cells using established techniques (Chittajallu and Isaac, 2010 and Daw et al., 2007a).

Ca2+ can inhibit or activate Ca2+ channels,

mediated by s

Ca2+ can inhibit or activate Ca2+ channels,

mediated by several EF-hand Ca2+-binding proteins. Specifically, calmodulin appears to both facilitate and inhibit voltage-dependent activation of Cav2.1 P/Q-type Ca2+ channels via binding to discrete sites in the cytoplasmic Ca2+ channel tail sequences (DeMaria et al., 2001 and Lee et al., 2003). In addition, another EF-hand Ca2+-binding protein called check details “calcium-binding protein 1” (CaBP1) increases inactivation of P/Q-type Ca2+ channels (Lee et al., 2002), whereas a third EF-hand Ca2+-binding protein called visinin-like protein 2 (VILIP-2) slows the rate of Ca2+ channel inactivation and enhances facilitation (Lautermilch et al., 2005). Moreover, Ca2+ channels are powerfully inhibited by G protein mediated mechanisms activated by presynaptic receptors, and such inhibition can also contribute to short-term synaptic plasticity. RG7204 order For example, GABAB-autoreceptors mediate short-term synaptic depression of inhibitory synapses during stimulus trains in insular cortex, illustrating this mode of short-term

synaptic plasticity (Kobayashi et al., 2012). However, most G protein mediated presynaptic inhibition of release by suppression of Ca2+ channel activation probably does not operate via autoreceptors, but via receptors for neuromodulators such as neuropeptides, endocannabinoids, acetylcholine, and catecholamines. The most prominent example of this process is depolarization-induced suppression of inhibition, a form of short-term plasticity where postsynaptically released endocannabinoids suppress presynaptic release of GABA by inhibiting presynaptic Ca2+ channels (Wilson and Nicoll, 2001). This widespread mechanism also operates outside of short-term plasticity to modulate entire neuronal ensembles, as seen for example in the

suppression of excitatory synaptic transmission at Schaffer collateral synapses in the CA1 region of the hippocampus by presynaptic Heterotrimeric G protein muscarinic receptors (Vogt and Regehr, 2001). In addition to short-term synaptic plasticity due to the interplay of residual Ca2+ and vesicle depletion and to the modulation of presynaptic Ca2+ channels, a third class of mechanisms mediates short-term plasticity via direct changes in the release machinery. Mutations in several proteins associated with the release machinery alter short-term plasticity in a manner independent of the first two sets of mechanisms, for example mutations in synapsins (Rosahl et al., 1995), Munc13 (Augustin et al., 1999), and RIMs (Schoch et al., 2002). The mechanisms by which these mutations cause such changes are largely unclear, except for one protein: Munc13. As we discussed earlier, Munc13 is an active zone protein that is essential for synaptic vesicle priming, probably because it catalyzes SNARE-complex formation via its MUN domain, and that is directly regulated by RIM proteins.