Understanding how Munc18 proteins govern exocytosis is crucial because mutations of this protein cause severe secretion deficits in neuronal and immune cells. Munc18-2 has indispensable roles in the degranulation of mast cell, partly by binding and chaperoning a subset of syntaxin isoforms. However, the key syntaxin that, crucially, participates in the degranulation – whose levels and intracellular localization are regulated by Munc18-2 – remains unknown. Here, we demonstrate that double knockdown of Munc18-1 and Munc-2 in mast cells results in greatly reduced degranulation accompanied with strikingly compromised expression levels and localization of syntaxin-3. This phenotype is fully rescued by wild-type Munc18 proteins but not by the K46E, E59K and K46E/E59K mutants of Munc-18 domain 1, each of which exhibits completely abolished binding to ‘closed’ syntaxin-3. Furthermore, knockdown of syntaxin-3 strongly impairs degranulation. Collectively, our data argue that residues Lys46 and Glu59 of Munc18 proteins are indispensable for mediating the interaction between Munc18 and closed syntaxin-3, which is essential for degranulation by chaperoning syntaxin-3. Our results also indicate that the functional contribution of these residues differs between immune cell degranulation and neuronal secretion.

Munc18-2 (also known as STXBP2) plays indispensable functions in immune cells. In humans, mutations of Munc18-2 have recently been associated with a fatal immune disorder, familial hemophagocytic lymphohistiocytosis type 5 (FHL5) (Cetica et al., 2010; Côte et al., 2009; Meeths et al., 2010; zur Stadt et al., 2009). In the immune tissue of these patients, there is a markedly reduced degranulation of lytic granules from cytotoxic T-lymphocytes and natural killer cells (Côte et al., 2009). In mice, Munc18-2 heterozygous knockdown results in significant impairment of mast cell degranulation (Kim et al., 2012). However, despite its importance, understanding the functions and its structural determinants of Munc18-2, which are necessary for immune cell exocytosis, has just begun recently (Bin et al., 2013).

In contrast, the precise roles of Munc18-1 (also known as STXBP1), a neuronal isoform of Munc18, have been studied extensively. The essential function of Munc18-1 in neuronal exocytosis (Verhage et al., 2000) is known to be mediated by its complex interactions with syntaxin-1 and/or the neuronal SNARE complex (Han et al., 2010). At least two important functions of Munc18-1 have been proposed: (1) priming of membrane fusion through direct interaction with the SNARE complex (Han et al., 2013; Han et al., 2014; Hu et al., 2010; Ma et al., 2013; Parisotto et al., 2014; Rodkey et al., 2008; Südhof and Rothman, 2009; Tareste et al., 2008; Xu et al., 2010); (2) molecular chaperone function of syntaxin-1 regulating appropriate intracellular localization and expression of syntaxin-1 (Arunachalam et al., 2008; Han et al., 2009; Malintan et al., 2009; McEwen and Kaplan, 2008; Medine et al., 2007; Rowe et al., 2001; Rowe et al., 1999; Shen et al., 2007). A similar chaperoning function (regulation of expression level as well as intracellular trafficking of its cognate syntaxin partner) of Munc18-2 has also been discovered; some FHL5 patients, who lack functional Munc18-2 protein, present with strongly reduced syntaxin-11 expression levels in their immune cells (Côte et al., 2009; zur Stadt et al., 2009). However, the underlying mechanisms governing such chaperoning function of Munc18-2 and its significance in the degranulation process remain unclear.

Therefore, two key questions remain unanswered about the Munc18-2 chaperoning function in mast cells: (1) What are the key cognate syntaxin isoform(s) that exhibit indispensable roles in mast cell degranulation and whose level and intracellular localization are regulated by Munc18-2? (2) Does the chaperoning activity of Munc18-2 directly depend on the binding of the Munc18-2 domain-1 cleft and those syntaxin(s) in the ‘closed’ conformation? We demonstrate here for the first time that loss of function in Munc18 impairs the expression level and intracellular localization of syntaxin-3. Furthermore, we show that such chaperoning activity of Munc18 is dependent on the direct binding between closed syntaxin-3 and the amino acid (aa) residues Lys46 and Glu59 in domain 1 of Munc18. Importantly, the functional contribution of these two residues in governing chaperoning activity towards syntaxin-3 is different compared with that of Munc18-1 and syntaxin-1 of neuronal exocytosis. Finally, strong and stable knockdown of syntaxin-3 results in vast deficits in mast cell degranulation. Collectively, our data invariably argue that the chaperoning activity of Munc18 towards syntaxin-3 is indispensable for the degranulation of mast cell.

Munc18 proteins regulate expression levels of syntaxin-3 and syntaxin-11 in mast cells

We have previously found that a stable Munc18-1–Munc18-2 double-knockdown (Munc18-1/2 DKD) in RBL-2H3 mast cells causes severe defects in degranulation, accompanied with the specific downregulation of syntaxin-11 of the various syntaxin isoforms (syntaxin-2, -3, -4, -6, -7, -8 and -11) examined by us (Bin et al., 2013). However, a stable knockdown of syntaxin-11 mast cells generated by us did not exhibit significant impairment of mast cell degranulation. Thus, it is unlikely that the degranulation defects observed in our Munc18-1/2 DKD mast cells are due to the reduction of syntaxin-11 levels itself. Therefore, we hypothesized that – if one of the primary functions of Munc18 proteins (hereafter referred to as Munc18) is to mediate the chaperoning action toward the syntaxins – there must be another key cognate syntaxin isoform, whose level is regulated by Munc18. Munc18-1 and Munc18-2 are known to bind to syntaxin-2 and syntaxin-3 in addition to syntaxin-11 and neuronal syntaxin-1 (Riento et al., 1998; Tamori et al., 1998). Recent work suggested that syntaxin-3 potentially substitutes for the function of syntaxin-11 in syntaxin-11-deficient mast cells (Hackmann et al., 2013). Thus, we examined a possible change of syntaxin-3 levels in our Munc18-1/2 DKD mast cells, by using a recently generated rabbit monoclonal anti-syntaxin-3 antibody (EPR8543, Abcam). Immunoblot analysis revealed a vast reduction in the levels of syntaxin-3 within Munc18-1/2 DKD mast cells (Fig. 1A). Our quantification revealed that more than 60% of syntaxin-3 was decreased in the DKD cells when compared with control cells (P<0.05, Fig. 1B). However, when we tested again for syntaxin-3 levels with the previously used rabbit polyclonal anti-syntaxin-3 antibody (Synaptic Systems), we failed to observe any obvious change in the expression of syntaxin-3 (data not shown). We think that this may be because the latter antibody is crossreacting more to other proteins due to its polyclonal nature, which hampered us to detect the reduction of syntaxin-3 level in our previous study (Bin et al., 2013). We also confirmed our previous findings of the mild decrease in the levels of syntaxin-11 in Munc18-2 single knockdown (KD) mast cells and of a more drastic decrease in Munc18-1/2 DKD mast cells (Fig. 1A).

Fig. 1.

Drastic reductions of syntaxin-3 and syntaxin-11 in Munc18-1/2 DKD mast cells and restoration by wild-type Munc18-1 or Munc18-2. (A) Twenty micrograms of stable Munc18-1, Munc18-2 single KD or Munc18-1/2 DKD RBL-2H3 cell homogenate were analyzed by SDS-PAGE and immunoblotting using the antibodies indicated on the right. *, non-specific band observed following the use of anti-Munc18-1 antibody. (B) Quantification of syntaxin-3 expression levels in Munc18-1, Munc18-2 single KD or Munc18-1/2 DKD RBL-2H3 cells. Blots on the film were quantified for densitometry and normalized to the respective control using ImageJ. Error bars indicate s.e.m. (n=6). The statistical significance of the differences in syntaxin-3 level between control and Munc18-1/2 DKD is indicated. *, P<0.05 (Student's t-test). (C) Stable EmGFP, wild-type Munc18-1- or Munc18-2–EmGFP-rescued Munc18-1/2 DKD RBL-2H3 cells were generated by using a lentivirus expression system. Twenty micrograms of cell homogenate were analyzed by SDS-PAGE and immunoblotting using antibodies indicated on the right. *, non-specific band observed with anti-Munc18-2 antibody, which co-migrated with Munc18-2–EmGFP.

Fig. 1.

Drastic reductions of syntaxin-3 and syntaxin-11 in Munc18-1/2 DKD mast cells and restoration by wild-type Munc18-1 or Munc18-2. (A) Twenty micrograms of stable Munc18-1, Munc18-2 single KD or Munc18-1/2 DKD RBL-2H3 cell homogenate were analyzed by SDS-PAGE and immunoblotting using the antibodies indicated on the right. *, non-specific band observed following the use of anti-Munc18-1 antibody. (B) Quantification of syntaxin-3 expression levels in Munc18-1, Munc18-2 single KD or Munc18-1/2 DKD RBL-2H3 cells. Blots on the film were quantified for densitometry and normalized to the respective control using ImageJ. Error bars indicate s.e.m. (n=6). The statistical significance of the differences in syntaxin-3 level between control and Munc18-1/2 DKD is indicated. *, P<0.05 (Student's t-test). (C) Stable EmGFP, wild-type Munc18-1- or Munc18-2–EmGFP-rescued Munc18-1/2 DKD RBL-2H3 cells were generated by using a lentivirus expression system. Twenty micrograms of cell homogenate were analyzed by SDS-PAGE and immunoblotting using antibodies indicated on the right. *, non-specific band observed with anti-Munc18-2 antibody, which co-migrated with Munc18-2–EmGFP.

To examine whether the observed reduction in syntaxin-3 expression levels was specifically due to the knockdown of Munc18 proteins and not the result of any off-target small hairpin RNA (shRNA) effects, we re-expressed shRNA-resistant wild-type Munc18-1 or Munc18-2 fused to emerald GFP (EmGFP) again in Munc18-1/2 DKD mast cells by using a lentivirus mediated expression system (see Materials and Methods). We succeeded in finding full restoration of reduced syntaxin-3 levels in both wild-type (WT) Munc18-1–EmGFP- and Munc18-2–EmGFP-rescued mast cells, indicating the specificity of the syntaxin-3 downregulation phenotype in our Munc18-1/2 DKD mast cells (Fig. 1C). We also rescued syntaxin-11 expression levels in our wild-type Munc18-rescued cells once more (Fig. 1C). Our results provide strong evidence that, in addition to syntaxin-11, Munc18 plays a crucial role in the regulation of syntaxin-3 expression levels in mast cells.

Munc18 controls syntaxin-3 intracellular localization in mast cells

Severe mislocalization of syntaxin-1 is observed following the single knockdown of Munc18-1 or the double knockdown of Munc18-1 and Munc18-2 in the neuroendocrine tumor PC12 cell line (Arunachalam et al., 2008; Han et al., 2011; Han et al., 2009). Similarly, in unc-18-null C. elegans, the anterograde trafficking and synaptic localization of UNC-64, an ortholog of syntaxin-1, is significantly perturbed (McEwen and Kaplan, 2008). In contrast, we have previously found that the endosomal localization of syntaxin-11 is unchanged in our Munc18-1/2 DKD mast cells (Bin et al., 2013). Because we hypothesize that Munc18 exhibits chaperoning activity toward syntaxin-3, we next asked whether the subcellular localization of syntaxin-3 could also be perturbed, in addition to its reduced expression, in the Munc18-1/2 DKD mast cells. Although syntaxin-2 expression levels were unchanged in our Munc18-1/2 DKD mast cells (Bin et al., 2013, see Fig. 1 therein), we tested for a possible alteration in syntaxin-2 localization because Munc18 proteins are known to physically interact with both syntaxin-2 and syntaxin-3. For a direct comparison between the two syntaxin isoforms in a systematic manner, we overexpressed hemagglutinin (HA)-tagged exogenous syntaxin-2 and syntaxin-3 in our control cells and in Munc18-1/2 DKD mast cells, and examined their intracellular localization by using anti-HA monoclonal antibody. HA tags were attached in tandem to the C-terminal end of these proteins (Fig. 2A), and we stably expressed them in control and Munc18-1/2 DKD mast cells (see Materials and Methods). Immunoblotting with anti–HA antibody revealed similar levels of overexpressed syntaxin-2–HA and syntaxin-3–HA in control and Munc18-1/2 DKD cells (Fig. 2B).

Fig. 2.

Systematic analysis of the intracellular localization of syntaxin-2 and syntaxin-3 in control and Munc18-1/2 DKD mast cells. (A) A schematic representation of recombinant rat syntaxin-2–HA (aa residues 1–290) and syntaxin-3–HA (aa residues 1–289) proteins. Lys 289 (K289) is located in the C-terminus of rat syntaxin-3 downstream of the TMR domain. Sequence VDSRDL constitutes a linker between syntaxins and the tandem HA tags. (B) Stable control and Munc18-1/2 DKD RBL-2H3 mast cells overexpressing rat syntaxin-2–HA or syntaxin-3–HA were generated by using a lentivirus expression system. Twenty micrograms of cell homogenate were analyzed by SDS-PAGE and immunoblotting using mouse monoclonal anti-HA antibody. (C,D) Syntaxin-2–HA or syntaxin-3–HA overexpressing control cells and Munc18-1/2 DKD RBL-2H3 mast cells were permeabilized and stained with mouse monoclonal anti-HA antibody followed by Alexa-Fluor-488-conjugated goat anti-mouse secondary antibody and DAPI. Green, syntaxin-2–HA (C) and syntaxin-3–HA (D); blue, DAPI. Control (top panels), Munc18-1/2 DKD (bottom panels). Left panels show wide views of the cells, whereas right panels show magnified view focusing on individual cells. Scale bars: 10 µm.

Fig. 2.

Systematic analysis of the intracellular localization of syntaxin-2 and syntaxin-3 in control and Munc18-1/2 DKD mast cells. (A) A schematic representation of recombinant rat syntaxin-2–HA (aa residues 1–290) and syntaxin-3–HA (aa residues 1–289) proteins. Lys 289 (K289) is located in the C-terminus of rat syntaxin-3 downstream of the TMR domain. Sequence VDSRDL constitutes a linker between syntaxins and the tandem HA tags. (B) Stable control and Munc18-1/2 DKD RBL-2H3 mast cells overexpressing rat syntaxin-2–HA or syntaxin-3–HA were generated by using a lentivirus expression system. Twenty micrograms of cell homogenate were analyzed by SDS-PAGE and immunoblotting using mouse monoclonal anti-HA antibody. (C,D) Syntaxin-2–HA or syntaxin-3–HA overexpressing control cells and Munc18-1/2 DKD RBL-2H3 mast cells were permeabilized and stained with mouse monoclonal anti-HA antibody followed by Alexa-Fluor-488-conjugated goat anti-mouse secondary antibody and DAPI. Green, syntaxin-2–HA (C) and syntaxin-3–HA (D); blue, DAPI. Control (top panels), Munc18-1/2 DKD (bottom panels). Left panels show wide views of the cells, whereas right panels show magnified view focusing on individual cells. Scale bars: 10 µm.

Upon immunostaining, we found that the syntaxin-2–HA had specifically localized to the plasma membrane (Fig. 2C, top panels), whereas syntaxin-3–HA was shown as punctate staining, suggesting its localization to secretory lysosomal granules (Fig. 2D, top panels). In addition, we found strong expression of syntaxin-3–HA at the tips of the mast cells. Upon Munc18-1/2 double knockdown, we found that such plasmalemmal localization of syntaxin-2–HA was unchanged (Fig. 2C, bottom panels). Strikingly, however, we observed severe perturbation of secretory lysosomal localization of syntaxin-3–HA in Munc18-1/2 DKD mast cells (Fig. 2D, bottom panels). Specifically, we found vast aggregation of syntaxin-3–HA in perinuclear regions of Munc18-1/2 DKD mast cells and its expression at the tips of the mast cells was largely reduced. This observation implies that Munc18 helps syntaxin-3–HA to translocate towards the tips of the mast cells and, therefore, in their absence, syntaxin-3–HA accumulates at the perinuclear region. Thus, our results indicate that Munc18 selectively regulates the intracellular localization of syntaxin-3 but not the plasmalemmal localization of syntaxin-2, even though it physically binds to both isoforms.

We then sought to reveal whether similar subcellular localization profiles can be seen for endogenous syntaxin-3 by staining the control and the Munc18-1/2 DKD mast cells with rabbit monoclonal syntaxin-3 antibody that we used for the immunoblotting. In accordance with the syntaxin-3–HA overexpression data (Fig. 2D), we found that endogenous syntaxin-3 was shown as discrete punctate pattern with particularly strong expression at the tips of elongated mast cells (Fig. 3A, top panels). Subsequently, for the Munc18-1/2 DKD mast cells, we were able to recapitulate the aggregation of syntaxin-3 in the perinuclear region and its absence at the tips of the cells (Fig. 3A, bottom panels). Also, we found that overall intensity for syntaxin-3 staining in the Munc18-1/2 DKD cells was much less compared with control, consistent with a lower level of syntaxin-3 expression in these cells, as analyzed by immunoblotting (Fig. 1A). Quantification analysis of confocal images revealed, compared with control, an overall syntaxin-3 intensity of ∼35% in Munc18-1/2 DKD (Fig. 3B, left panel). This reduction was similar to the reduction of syntaxin-3 expression levels in our immunoblot analysis (Fig. 1B). In addition, we found that most of syntaxin-3 was expressed at the tip of the control cells (70% at the tips versus 30% in the soma). However, we found syntaxin-3 strikingly – with nearly 80% – aggregated in the soma, whereas only 20% was found at the tips of Munc18-1/2 DKD cells (Fig. 3B, right panel). This result, therefore, further confirms that the localization of syntaxin-3 is strikingly perturbed by the Munc18-1/2 double knockdown in the mast cells.

Fig. 3.

Confocal microscopy of endogenous syntaxin-3 reveals striking mislocalization in Munc18-1/2 DKD mast cells that was rescued upon re-introduction of either Munc18-1–EmGFP or Munc18-2–EmGFP. (A) Control and Munc18-1/2 DKD RBL-2H3 mast cells were permeabilized and stained with rabbit monoclonal anti-syntaxin-3 antibody followed by Alexa-Fluor-488-conjugated goat anti-rabbit secondary antibody and DAPI. Green, endogenous syntaxin-3; blue, DAPI. Control (top panels), Munc18-1/2 DKD (bottom panels). Left panels show wide views of the cells, whereas right panels show magnified view focusing on an individual cell. Scale bars: 10 µm. (B) Quantification of syntaxin-3 expression in control and Munc18-1/2 DKD RBL-2H3 mast cells was performed using ImageJ. Left graph shows the fluorescence intensity; right graph indicates distribution of the fluorescence signal (tip versus soma). Analyzed were 22 cells for control and 32 cells for DKD cells, and their mean values are presented. Error bars indicate s.e.m. (C) Confocal images of Munc18-1/2 DKD RBL-2H3 mast cells infected with lentivirus that express either EmGFP alone (top panels), wild-type Munc18-1–EmGFP (middle panels) or wild-type Munc18-2–EmGFP (bottom panels). Cells were permeabilized and co-stained with mouse monoclonal anti-GFP antibody and rabbit monoclonal anti-syntaxin-3 antibody followed by Alexa-Fluor-488-conjugated goat anti-mouse secondary antibody, Rhodamine-x-conjugated goat anti-rabbit secondary antibody and DAPI. Green, EmGFP, Munc18-1–EmGFP and Munc18-2–EmGFP (left panels); red, syntaxin-3 (middle panels) and merged images (right panels). Notice the reduced staining intensity of syntaxin-3 in EmGFP rescued cells compared to wild-type Munc18 rescued cells. Scale bars: 10 µm.

Fig. 3.

Confocal microscopy of endogenous syntaxin-3 reveals striking mislocalization in Munc18-1/2 DKD mast cells that was rescued upon re-introduction of either Munc18-1–EmGFP or Munc18-2–EmGFP. (A) Control and Munc18-1/2 DKD RBL-2H3 mast cells were permeabilized and stained with rabbit monoclonal anti-syntaxin-3 antibody followed by Alexa-Fluor-488-conjugated goat anti-rabbit secondary antibody and DAPI. Green, endogenous syntaxin-3; blue, DAPI. Control (top panels), Munc18-1/2 DKD (bottom panels). Left panels show wide views of the cells, whereas right panels show magnified view focusing on an individual cell. Scale bars: 10 µm. (B) Quantification of syntaxin-3 expression in control and Munc18-1/2 DKD RBL-2H3 mast cells was performed using ImageJ. Left graph shows the fluorescence intensity; right graph indicates distribution of the fluorescence signal (tip versus soma). Analyzed were 22 cells for control and 32 cells for DKD cells, and their mean values are presented. Error bars indicate s.e.m. (C) Confocal images of Munc18-1/2 DKD RBL-2H3 mast cells infected with lentivirus that express either EmGFP alone (top panels), wild-type Munc18-1–EmGFP (middle panels) or wild-type Munc18-2–EmGFP (bottom panels). Cells were permeabilized and co-stained with mouse monoclonal anti-GFP antibody and rabbit monoclonal anti-syntaxin-3 antibody followed by Alexa-Fluor-488-conjugated goat anti-mouse secondary antibody, Rhodamine-x-conjugated goat anti-rabbit secondary antibody and DAPI. Green, EmGFP, Munc18-1–EmGFP and Munc18-2–EmGFP (left panels); red, syntaxin-3 (middle panels) and merged images (right panels). Notice the reduced staining intensity of syntaxin-3 in EmGFP rescued cells compared to wild-type Munc18 rescued cells. Scale bars: 10 µm.

Since re-introduction of either wild-type Munc18-1 or Munc18-2 into Munc18-1/2 DKD mast cells restored the reduced expression of syntaxin-3 (Fig. 1C), we next asked whether its perturbed localization can also be rescued. For this purpose, Munc18-1/2 DKD RBL-2H3 mast cells expressing EmGFP, wild-type Munc18-1–EmGFP or wild-type Munc18-2–EmGFP were co-stained for GFP and syntaxin-3 in corresponding rescued cells by using mouse anti-GFP monoclonal and rabbit anti-syntaxin-3 monoclonal antibodies; syntaxin-3 localization was then visualized using confocal microscopy (Fig. 3C). First we noticed that, in comparison to EmGFP-rescued cells, both WT Munc18-1–EmGFP- and WT Munc18-2–EmGFP-rescued cells exhibited brighter syntaxin-3 intensity (an ∼4-fold increase in intensity compared to that of EmGFP, see Fig. 7 for quantification), consistent with their restored expression observed in the immunoblots. (Fig. 1C). In addition, we found that the strong syntaxin-3 expression in the tips re-appeared in both Munc18-1–EmGFP- and Munc18-2 WT–EmGFP-rescued cells, but not in EmGFP-rescued cells. Importantly, we found that the perinuclear aggregation of syntaxin-3 seen in EmGFP-rescued cells was largely alleviated in both Munc18-1–EmGFP- and Munc18-2 WT–EmGFP-rescued cells (Fig. 3C). Quantification analysis revealed that almost 80% of syntaxin-3 was found to be aggregated in the soma of EmGFP rescued cells, yet its strong expression pattern at the tips was restored in both Munc18-1–EmGFP- and Munc18-2–EmGFP-rescued cells with nearly 60% of syntaxin-3 in the tips versus ∼40% in the soma (Fig. 7). Collectively, our results strongly advocate that Munc18 functions as a chaperone of syntaxin-3 by regulating both its expression levels and intracellular trafficking. In addition, we illustrate here that Munc18-mediated syntaxin trafficking is not only present in neuronal cells (Arunachalam et al., 2008; Han et al., 2011; Han et al., 2009; McEwen and Kaplan, 2008) but also in mast cells.

Aggregated syntaxin-3 – as seen in Munc18-1/2 DKD mast cells – does not colocalize with Golgi or ER markers

We found that in the Munc18-1/2 DKD mast cells syntaxin-3 aggregated near the nucleus. Since we hypothesize that Munc18 acts as a molecular chaperone by mediating the trafficking of syntaxin-3, we asked whether such aggregation occurs in the Golgi complex or compartments of the endoplasmic reticulum (ER) of Munc18-1/2 DKD cells. To test for syntaxin-3 localization in the Golgi complex, we co-stained for syntaxin-3 with two different Golgi markers – GM130 for cis-Golgi and TGN38 for the trans-Golgi network – in the Munc18-1/2 DKD cells and analyzed their localizations using confocal microscopy (supplementary material Fig. S1A). We found that aggregated syntaxin-3 does not colocalize with either of the Golgi markers. Then we asked whether aggregation occurs in the ER compartment by staining cells with anti-calnexin antibody (supplementary material Fig. S1B). Since this antibody was made against rabbit calnexin, we were unable to use the anti-syntaxin-3 antibody for co-staining. However, we noticed that calnexin expression was different compared with the pattern of aggregated syntaxin-3, being more spread throughout the cells unlike the aggregated syntaxin-3 pattern in the Munc18-1/2 DKD cells. Therefore, we conclude that, although we see perinuclear syntaxin-3 aggregation, syntaxin-3 does not occur in the ER or the Golgi compartments of the Munc18-1/2 DKD mast cells.

Lys46 and Glu59 residues in the domain 1 of Munc18 play crucial roles in mediating binary interaction with monomeric syntaxin-2 and syntaxin-3

Binding of Munc18-1 to monomeric closed syntaxin-1 is mediated by interactions of many residues, as revealed by X-ray crystallography (Misura et al., 2000). Among those residues, we have previously shown that Lys46 and Glu59 in the domain 1 of Munc18-1 play a crucial role in governing the binary interaction (interaction between Munc18 and monomeric syntaxin). We found that either of the two point mutations (i.e. K46E or E59K) significantly disrupted the binding to closed syntaxin-1, whereas the K46E/E59K double mutation resulted in complete abolishment of interaction, as seen using yeast two-hybrid assays and isothermal titration calorimetry (ITC) (Han et al., 2011; Han et al., 2009). Lys46 and Glu59 are conserved not only between Munc18-1 and Munc18-2 but also in UNC-18 of C. elegans and Rop in Drosophila melanogaster (Fig. 4A). Also, a homology model of Munc18-2 in complex with monomeric syntaxin-3 based on the crystal structure of Munc18-1–syntaxin-1A suggested that these two residues are in close proximity to Asp230 and Arg114 of closed syntaxin-3, possibly mediating electrostatic interactions (Fig. 4B). Therefore, we next investigated whether mutation of these residues in Munc18-1 and Munc18-2 would affect the binary interactions (if so, how much) with syntaxin-2 and syntaxin-3 in yeast two-hybrid assays (Fig. 4C–F).

Fig. 4.

See next page for legend.

K46E and E59K mutations in domain 1 of Munc18-1 or Munc18-2 result in the abolishment of binding to closed syntaxin-3. (A) Sequence alignment of domain 1 residues of rat Munc18-2 and Munc18-1, C. elegans UNC-18, and Drosophila Rop indicate conservation of Lys46 and Glu59 among different Munc18 isoforms and homologs (highlighted in red). (B) Predicted structure of Munc18-2 in complex with syntaxin-3, by using homology modeling based on Munc18-1/syntaxin-1A crystal structure (Misura et al., 2000). Each domain of Munc18-2 is represented in a different color. Domains 1, 2, and 3 are shown in green, purple and pink, respectively. Syntaxin-3 is shown in cyan. An enlarged representative structure is shown on the right, illustrating the location of Lys46 and Glu59 of Munc18-2, and their potential electrostatic interaction partners Asp230 and Arg114, respectively, in syntaxin-3. Direct binding between Munc18-1 or Munc18-2 wild-type and the binary interaction mutants (K46E, E59K, K46E/E59K) with syntaxin-3 (C,D), and syntax-2 (E,F) was analyzed by yeast two-hybrid assays. Munc18 proteins were expressed as bait, whereas the cytoplasmic domains of wild-type syntaxins were expressed as prey. In each assay, β-galactosidase activities of the transformed yeast clones were quantified and normalized so that the activity of the yeast clones transformed with the wild-type Munc18 was set to 100%. (G) The syntaxin-3 LE open mutant was expressed as prey and binding to wild-type Munc18-1 or Munc18-2 was tested. Resulting β-galactosidase activity of the syntaxin-3 LE open mutant was quantified and normalized to the binding of wild-type syntaxin-3 to Munc18. pLexN indicates empty bait vector. Error bars indicate s.e.m. (n=6–15).

Fig. 4.

See next page for legend.

K46E and E59K mutations in domain 1 of Munc18-1 or Munc18-2 result in the abolishment of binding to closed syntaxin-3. (A) Sequence alignment of domain 1 residues of rat Munc18-2 and Munc18-1, C. elegans UNC-18, and Drosophila Rop indicate conservation of Lys46 and Glu59 among different Munc18 isoforms and homologs (highlighted in red). (B) Predicted structure of Munc18-2 in complex with syntaxin-3, by using homology modeling based on Munc18-1/syntaxin-1A crystal structure (Misura et al., 2000). Each domain of Munc18-2 is represented in a different color. Domains 1, 2, and 3 are shown in green, purple and pink, respectively. Syntaxin-3 is shown in cyan. An enlarged representative structure is shown on the right, illustrating the location of Lys46 and Glu59 of Munc18-2, and their potential electrostatic interaction partners Asp230 and Arg114, respectively, in syntaxin-3. Direct binding between Munc18-1 or Munc18-2 wild-type and the binary interaction mutants (K46E, E59K, K46E/E59K) with syntaxin-3 (C,D), and syntax-2 (E,F) was analyzed by yeast two-hybrid assays. Munc18 proteins were expressed as bait, whereas the cytoplasmic domains of wild-type syntaxins were expressed as prey. In each assay, β-galactosidase activities of the transformed yeast clones were quantified and normalized so that the activity of the yeast clones transformed with the wild-type Munc18 was set to 100%. (G) The syntaxin-3 LE open mutant was expressed as prey and binding to wild-type Munc18-1 or Munc18-2 was tested. Resulting β-galactosidase activity of the syntaxin-3 LE open mutant was quantified and normalized to the binding of wild-type syntaxin-3 to Munc18. pLexN indicates empty bait vector. Error bars indicate s.e.m. (n=6–15).

We expressed wild-type and the binary interaction mutants (K46E, E59K and K46E/E59K) of rat Munc18 as bait and the cytoplasmic domains of rat syntaxin-2 and syntaxin-3 as prey (see Materials and Methods). When we tested the binding of the mutants to monomeric syntaxin-2 and syntaxin-3, we found significant disrupting effects of mutations K46E and E59K on the binary interactions (Fig. 4C–F). However, we found intriguing differences between syntaxin-2 and syntaxin-3. Strikingly, even the single mutation of K46E or E59K in Munc18-1 or Munc18-2 completely abolished the binding to syntaxin-3, and double mutation of K46E/E59K did not exhibit any additive effects (Fig. 4C,D). In contrast, the single binary interaction mutants of Munc18-1 and, to a lesser extent, those of Munc18-2 retained some ability to bind syntaxin-2. However, as observed in syntaxin-3, the double mutants of Munc18-1 or Munc18-2 similarly abrogated the binding to syntaxin-2 (Fig. 4E,F).

Based on our homology model illustrating Munc18-2 bound to closed syntaxin-3 (Fig. 4B), we speculated that the wild-type syntaxin-3 adopts a closed conformation upon binding to Munc18. To verify that, indeed, wild-type syntaxin-3 adopts a closed conformation and that Munc18-2 primarily binds to this closed conformation, we examined whether binding between Munc18-2 and syntaxin-3 is significantly reduced when syntaxin-3 adopts an open conformation. In a previous NMR study of syntaxin-1A (Dulubova et al., 1999), mutations (L165A/E166A) in a linker region between its N-terminal Habc domain and its SNARE domain were found to cause this protein to open. These residues are highly conserved from nematode to human; similar mutations (L165A/E166A) in the syntaxin-1 ortholog UNC-64 also cause this protein to adopt an open conformation (Richmond et al., 2001). Therefore, we generated a syntaxin-3 mutant that has an open conformation (L165E/E166A, hereafter referred to as LE open mutant) and then tested its interaction with both isoforms of Munc18 in yeast two-hybrid assays. We found that, compared to wild-type syntaxin-3, the syntaxin-3 open mutant merely exhibited ∼10% of binding to either Munc18-1 or Munc18-2 isoforms (Fig. 4G). This indicates that the interaction of wild-type syntaxin-3 with Munc18 is mainly governed by syntaxin-3 in a closed conformation. Importantly, the fact that K46E and E59K alone completely abolished the interaction with wild-type syntaxin-3 indicates that these two chaperoning mutants play an essential role in mediating binary interaction with closed syntaxin-3 (Fig. 4C,D). In addition, although the binding mode of Munc18-1 and Munc18-2 towards closed syntaxin-2 seems similar to the binding mode of Munc18-1 and closed syntaxin-1 in neuronal exocytosis, as previously reported (Han et al., 2011; Han et al., 2009), their interactions with closed syntaxin-3 are substantially different, in that even a single mutation is enough to disrupt such interaction.

Lys46 and Glu59 of Munc18 are essential for chaperoning of syntaxin-3 and mast cell degranulation

To test the functional outcome of these binary interaction mutations with regard to the regulation of syntaxin-3 expression levels, localization and mast cell degranulation, we stably expressed the EmGFP-tagged mutants of Munc18-2 (K46E and E59K single mutants, and K46E/E59K double mutant) in Munc18-1/2 DKD mast cells together with wild-type and examined their ability to restore compromised syntaxin-3 expression level, subcellular localization as well as defective degranulation. Immunoblot analysis illustrated that all mutants were expressed at levels similar to that of wild-type Munc18-2, implying that there were no major folding issues caused by these point mutations (Fig. 5A). However, we found that none of the binary interaction mutants of Munc18-2 was able to restore the reduced syntaxin-3 expression level of the Munc18-1/2 DKD mast cells (Fig. 5A). We then examined whether the mislocalization of syntaxin-3 seen in the Munc18-1/2 DKD mast cells can be rescued by the mutants (Fig. 5B). Consistent with the reduced expression of syntaxin-3, we found that overall staining intensities were substantially less in the each of the mutant-rescued cell in comparison to the wild-type-rescued cells. In fact, quantification of signal intensities showed similarly less intense signals in Munc18-2 EmGFP-tagged mutant-rescued cells and EmGFP-rescued cells (Fig. 7A). Strikingly, we observed expression of syntaxin-3 in the binary-interaction-mutant-rescued cells in an aggregated manner and failed to see syntaxin-3 expression at the tips of these cells (Fig. 5B, Fig. 7B for quantification). Finally, we tested the ability of these mutants to rescue the defective degranulation phenotype of Munc18-1/2 DKD mast cells by measuring release of β-hexosaminidase (Fig. 5C). This protein resides inside of secretory granules and/or lysosomes of mast cells and we triggered its release using immunoglobulin E (IgE) and 2,4-dinitrophenyl hapten conjugated to human serum albumin (DNP-HSA) or ionomycin as two different stimulation methods (Bin et al., 2013). IgE-DNP antibody is recognized by the high-affinity IgE receptor (FcεRI), a cell surface receptor of mast cells. Upon cross-linking its antigen DNP-HSA, signaling pathways eventually lead to the increase in intracellular Ca2+ levels, resulting in degranulation. Ionomycin is a potent Ca2+ ionophore that directly induces Ca2+ influx, thus triggering the degranulation. Strikingly, we found that none of the Munc18-2 binary mutants were able to rescue β-hexosaminidase secretion in response to either stimulation (Fig. 5C), indicating that the degranulation defects lie within the fusion machineries of mast cells. In accordance with the results of our yeast two-hybrid assays (Fig. 4D), we did not see any additive effects of the double mutant (K46E/E59K) over the single mutants (K46E or E59K), suggesting that even a single mutation is detrimental, i.e. abolishes the functional outcome of the binary interaction towards syntaxin-3 in the mast cells.

Fig. 5.

Binary interaction mutants of Munc18-2 do not restore syntaxin-3 levels, localization or degranulation defects of Munc18-1/2 DKD mast cells. (A) Munc18-1/2 DKD RBL-2H3 cells were infected with lentiviruses that express EmGFP, wild-type Munc18-2–EmGFP or the binary interaction mutants of Munc18-2–EmGFP (K46E, E59K, K46E/E59K). Infected cells were selected with blasticidin (20 µg/ml). Homogenates of surviving cells (20 µg) were analyzed by SDS-PAGE and immunoblotting using the antibodies indicated on the right. *, a non-specific band observed with anti-Munc18-2 antibody, which co-migrated with Munc18-2–EmGFP. (B) Munc18-1/2 DKD RBL-2H3 expressing the Munc18-2–EmGFP variants were permeabilized, and co-stained with mouse monoclonal anti-GFP and rabbit monoclonal anti-syntaxin-3 antibodies followed by Alexa-Fluor-488-conjugated goat anti-mouse secondary antibody, Rhodamine-x-conjugated goat anti-rabbit secondary antibody and DAPI. Green, Munc18-2–EmGFP of wild-type and mutants (left panels); red, syntaxin-3 (middle panels) and merged images (right panels). Notice the absence of tip staining and the increased perinuclear aggregation of syntaxin-3 in the mutant-rescued cells compared to the wild-type Munc18-rescued cells. Scale bars: 10 µm. (C) β-hexosaminidase release from the rescued cells was stimulated by 1-hour incubation with 0.01 µg/ml DNP-IgE and 50 ng/ml DNP-HSA, or 0.5 µM or 2.5 µM ionomycin. Error bars indicate s.e.m. (n=6).

Fig. 5.

Binary interaction mutants of Munc18-2 do not restore syntaxin-3 levels, localization or degranulation defects of Munc18-1/2 DKD mast cells. (A) Munc18-1/2 DKD RBL-2H3 cells were infected with lentiviruses that express EmGFP, wild-type Munc18-2–EmGFP or the binary interaction mutants of Munc18-2–EmGFP (K46E, E59K, K46E/E59K). Infected cells were selected with blasticidin (20 µg/ml). Homogenates of surviving cells (20 µg) were analyzed by SDS-PAGE and immunoblotting using the antibodies indicated on the right. *, a non-specific band observed with anti-Munc18-2 antibody, which co-migrated with Munc18-2–EmGFP. (B) Munc18-1/2 DKD RBL-2H3 expressing the Munc18-2–EmGFP variants were permeabilized, and co-stained with mouse monoclonal anti-GFP and rabbit monoclonal anti-syntaxin-3 antibodies followed by Alexa-Fluor-488-conjugated goat anti-mouse secondary antibody, Rhodamine-x-conjugated goat anti-rabbit secondary antibody and DAPI. Green, Munc18-2–EmGFP of wild-type and mutants (left panels); red, syntaxin-3 (middle panels) and merged images (right panels). Notice the absence of tip staining and the increased perinuclear aggregation of syntaxin-3 in the mutant-rescued cells compared to the wild-type Munc18-rescued cells. Scale bars: 10 µm. (C) β-hexosaminidase release from the rescued cells was stimulated by 1-hour incubation with 0.01 µg/ml DNP-IgE and 50 ng/ml DNP-HSA, or 0.5 µM or 2.5 µM ionomycin. Error bars indicate s.e.m. (n=6).

We then performed additional experiments to investigate the effect of the same mutations in Munc18-1 on the chaperoning of syntaxin-3 and the mast cell degranulation (Fig. 6). We have previously found that the ability of Munc18-1 to rescue the phenotype of Munc18-1/2 DKD PC12 cells is correlated to their ability to bind syntaxin-1. We observed that the K46E and E59K single mutants retained some binding to syntaxin-1 and, thus were able to restore β-hexosaminidase secretion to 70% compared to the wild-type rescue, whereas the double mutant lost its ability to bind to syntaxin-1 and, therefore, rescued secretion at the level of merely 10% compared to wild-type (Han et al., 2011). In addition, synaptic transmission of hippocampal neurons that lack Munc18-1 was relatively well rescued by the E59K single mutant, which exhibited EPSC size that was comparable to that of wild-type (Meijer et al., 2012). However, in the rescue of the Munc18-1/2 DKD mast cells, we found that – as for Munc18-2 – even the single mutations of K46E or E59K of Munc18-1 were unable to restore reduced syntaxin-3 expression levels, its mislocalization or the impaired degranulation (Fig. 6 and Fig. 7C,D for quantification). Thus, in the case of the mast cells, the ability of Munc18-1 and Munc18-2 to rescue the loss of syntaxin-3 chaperoning and defective degranulation in Munc18-1/2 DKD cells was strikingly different from the results of Munc18-1 and syntaxin-1 interaction in neuronal secretion. Also, this functional phenotype of the binary interaction mutants was well reflected on their ability to directly bind to syntaxin-3, which we saw in our yeast two-hybrid assays (Fig. 4C).

Fig. 6.

Binary interaction mutants of Munc18-1 do not restore syntaxin-3 levels, localization or degranulation defects of Munc18-1/2 DKD mast cells. (A) Lentivirus expressing EmGFP, wild-type and binary interaction mutants (K46E, E59K, K46E/E59K) of Munc18-1–EmGFP were applied to Munc18-1/2 DKD RBL-2H3 cells. Homogenates of isolated cells (20 µg) were analyzed by SDS-PAGE and immunoblotting using antibodies indicated on the right. (B) Munc18-1/2 DKD RBL-2H3 cells expressing Munc18-1–EmGFP variants were permeabilized and co-stained with mouse monoclonal anti-GFP antibody and rabbit monoclonal anti-syntaxin-3 antibody followed by Alexa-Fluor-488-conjugated goat anti-mouse secondary antibody, Rhodamine-x-conjugated goat anti-rabbit secondary antibody and DAPI. Green, Munc18-1–EmGFP of wildtype and binary interaction mutants (left panels); red, syntaxin-3 (middle panels) and merged images (right panels). Notice the absence of tip staining and the decreased intensity of syntaxin-3 in the mutant-rescued cells compared to the wild-type Munc18-1-rescued cells. Scale bars: 10 µm. (C) β-hexosaminidase release from the rescued cells was stimulated by 1-hour incubation with 0.01 µg/ml DNP-IgE and 50 ng/ml DNP-HSA,or 0.5 µM or 2.5 µM ionomycin. Error bars indicate s.e.m. (n=6).

Fig. 6.

Binary interaction mutants of Munc18-1 do not restore syntaxin-3 levels, localization or degranulation defects of Munc18-1/2 DKD mast cells. (A) Lentivirus expressing EmGFP, wild-type and binary interaction mutants (K46E, E59K, K46E/E59K) of Munc18-1–EmGFP were applied to Munc18-1/2 DKD RBL-2H3 cells. Homogenates of isolated cells (20 µg) were analyzed by SDS-PAGE and immunoblotting using antibodies indicated on the right. (B) Munc18-1/2 DKD RBL-2H3 cells expressing Munc18-1–EmGFP variants were permeabilized and co-stained with mouse monoclonal anti-GFP antibody and rabbit monoclonal anti-syntaxin-3 antibody followed by Alexa-Fluor-488-conjugated goat anti-mouse secondary antibody, Rhodamine-x-conjugated goat anti-rabbit secondary antibody and DAPI. Green, Munc18-1–EmGFP of wildtype and binary interaction mutants (left panels); red, syntaxin-3 (middle panels) and merged images (right panels). Notice the absence of tip staining and the decreased intensity of syntaxin-3 in the mutant-rescued cells compared to the wild-type Munc18-1-rescued cells. Scale bars: 10 µm. (C) β-hexosaminidase release from the rescued cells was stimulated by 1-hour incubation with 0.01 µg/ml DNP-IgE and 50 ng/ml DNP-HSA,or 0.5 µM or 2.5 µM ionomycin. Error bars indicate s.e.m. (n=6).

Fig. 7.

Quantification of syntaxin-3 expression, and expression at the tips versus soma of wild-type Munc18-1- or Munc18-2-rescued and binary mutant-rescued Munc18-1/2 DKD mast cells. (A,B) Overall syntaxin-3 expression levels (A) and syntaxin-3 expression at the tips versus soma (B) of EmGFP, Munc18-2 WT, Munc18-2 K46E-, E59K- and K46E/E59K-rescued Munc18-1/2 DKD RBL-2H3 cells. Confocal microscopy images were quantified by ImageJ. (C,D) The same quantification was performed for Munc18-1 WT-, Munc18-1 K46E-, E59K- and K46E/E59K-rescued Munc18-1/2 DKD RBL-2H3 cells illustrating overall syntaxin-3 expression levels (C) and syntaxin-3 expression at the tips versus soma (D). See Materials and Methods for the quantification method. Note, in both Munc18 isoforms, WT rescued both overall intensity as well as tip localization of syntaxin-3, whereas all binary mutants failed to rescue this phenotype. Error bars indicate s.e.m. (n=23–40).

Fig. 7.

Quantification of syntaxin-3 expression, and expression at the tips versus soma of wild-type Munc18-1- or Munc18-2-rescued and binary mutant-rescued Munc18-1/2 DKD mast cells. (A,B) Overall syntaxin-3 expression levels (A) and syntaxin-3 expression at the tips versus soma (B) of EmGFP, Munc18-2 WT, Munc18-2 K46E-, E59K- and K46E/E59K-rescued Munc18-1/2 DKD RBL-2H3 cells. Confocal microscopy images were quantified by ImageJ. (C,D) The same quantification was performed for Munc18-1 WT-, Munc18-1 K46E-, E59K- and K46E/E59K-rescued Munc18-1/2 DKD RBL-2H3 cells illustrating overall syntaxin-3 expression levels (C) and syntaxin-3 expression at the tips versus soma (D). See Materials and Methods for the quantification method. Note, in both Munc18 isoforms, WT rescued both overall intensity as well as tip localization of syntaxin-3, whereas all binary mutants failed to rescue this phenotype. Error bars indicate s.e.m. (n=23–40).

Knockdown of syntaxin-3 results in severe impairment of degranulation in mast cells

If the chaperoning of syntaxin-3 by Munc18-1 and Munc18-2 is functionally important in mast cell degranulation, we would anticipate that this syntaxin isoform plays an important independent role in exocytosis. To test this hypothesis, we generated two stable RBL-2H3 mast cell lines whose endogenous syntaxin-3 was knocked down (syntaxin-3–5 KD and syntaxin-3–8 KD, see Materials and Methods). Immunoblotting with rabbit monoclonal anti-syntaxin-3 antibody revealed some reduction in the level of syntaxin-3 in both cell-lines, but syntaxin-3–8 KD resulted in a stronger knockdown than syntaxin-3–5 KD (Fig. 8A,C). Also, we did not observe any obvious changes in the expression of syntaxin-2 and syntaxin-11 in cells subjected to the stronger syntaxin-3 knockdown, indicating that these two isoforms were not upregulated (Fig. 8A). When performing degranulation assays using IgE- and ionomycin-dependent stimulations, we found significant degranulation defects in both syntaxin-3 KD mast cells (Fig. 8B,D). However, the degree of degranulation defect was correlated to the level of syntaxin-3 reduction seen in two different syntaxin-3 KD cell lines. We saw modest, yet statistically significant, reduction in degranulation in syntaxin-3–5 KD cells. However, we found a reduction greater than 4-fold (∼46% versus ∼11%) in IgE-induced degranulation, whereas ionomycin-induced degranulation showed a 2–3-fold reduction (∼35% versus ∼12% for 0.5 µM and ∼45% versus ∼10% for 2.5 µM) in syntaxin-3–8 KD mast cells compared to the control. This result suggests that syntaxin-3 plays an important independent role in mast cell degranulation, and the severe secretion defects seen in the Munc18-1/2 DKD mast cells might be attributed to the lack of syntaxin-3 chaperoning.

Fig. 8.

Stable knockdown of syntaxin-3 in mast cells results in severe impairment of β-hexosaminidase release. (A,C) Stable syntaxin-3 knockdown RBL-2H3 cells were generated by two different lentivirus-mediated shRNAs targeting different sequences of rat syntaxin-3. Cell homogenates (20 µg) were analyzed by SDS-PAGE and immunoblotting using antibodies indicated on the right. The signals were detected using an enhanced chemiluminescence detection system. Numbers on the left indicate molecular mass markers. (B,D) β-hexosaminidase release from the syntaxin-3–8 KD (B) and syntaxin-3–5 KD (D) KD RBL-2H3 cells was stimulated by applying 0.01 µg/ml DNP-IgE and 100 ng/ml DNP-HSA, or 0.5 µM or 2.5 µM ionomycin for 1 hour. Error bars indicate s.e.m. (n=19, B; and n=8, D).

Fig. 8.

Stable knockdown of syntaxin-3 in mast cells results in severe impairment of β-hexosaminidase release. (A,C) Stable syntaxin-3 knockdown RBL-2H3 cells were generated by two different lentivirus-mediated shRNAs targeting different sequences of rat syntaxin-3. Cell homogenates (20 µg) were analyzed by SDS-PAGE and immunoblotting using antibodies indicated on the right. The signals were detected using an enhanced chemiluminescence detection system. Numbers on the left indicate molecular mass markers. (B,D) β-hexosaminidase release from the syntaxin-3–8 KD (B) and syntaxin-3–5 KD (D) KD RBL-2H3 cells was stimulated by applying 0.01 µg/ml DNP-IgE and 100 ng/ml DNP-HSA, or 0.5 µM or 2.5 µM ionomycin for 1 hour. Error bars indicate s.e.m. (n=19, B; and n=8, D).

In the present study, we examined the role of binary interaction between domain 1 of Munc18 and its cognate syntaxin isoforms in the mast cell exocytosis. This interaction mode has been regarded as of importance in neuronal secretion, as the mutations in Munc18-1 that disrupt such interaction have demonstrated lack of chaperoning activity towards syntaxin-1 thus resulting in compromised neuronal exocytosis (Han et al., 2011; Han et al., 2009). However, no attempts have been made to study the binary interaction mode in the mast cell exocytosis, despite the fact that similar Munc18-regulated secretion processes take place. We demonstrated, for the first time, the crucial role of the binary interaction between Munc18 and syntaxin-3 in mast cell degranulation.

Importantly, we have shown that Munc18 performs an essential role in chaperoning (i.e. regulating the expression levels as well as intracellular trafficking) of its cognate partner syntaxin-3. We saw that loss of Munc18 function in mast cells impairs the expression levels of syntaxin-3 (Fig. 1A,B). These decreased syntaxin-3 levels were completely restored upon re-expression of Munc18-1 or Munc18-2, illustrating the specificity of such phenotype (Fig. 1C). Furthermore, our finding of mislocalized syntaxin-3 in the perinuclear region of Munc18-1/2 DKD mast cells, which was rescued by re-introduction of wild-type Munc18 (Figs 2, 3), provides additional strong evidence that Munc18 plays a crucial role in chaperoning syntaxin-3. In an attempt to investigate an independent role of syntaxin-3 in mast cell degranulation, we generated stable syntaxin-3 KD mast cells, which exhibited strong impairments in both IgE- and ionomycin-dependent degranulation (Fig. 8). This is different compared with our previous attempt to see the role of syntaxin-11 in the mast cell exocytosis, where strong knockdown of endogenous syntaxin-11 did not result in any defect of degranulation (Bin et al., 2013). Our new findings are in accordance with the previous studies suggesting the importance of Munc18-2 and syntaxin-3 in mast cell degranulation (Brochetta et al., 2014; Tadokoro et al., 2007). Thus, we now state that, although Munc18 proteins regulate both syntaxin-3 and syntaxin-11 isoforms in mast cells, the former interactions provide more functional contributions to the mast cell exocytosis.

Recent analysis of Munc18-2 and syntaxin-3 in RBL-2H3 cells illustrates the significant function of these two proteins in the mast cell degranulation (Brochetta et al., 2014). In that study, the authors separately knocked down Munc18-2 and syntaxin-3 RBL-2H3 by using small interfering RNA (siRNA) and found substantial reduction of degranulation within both these KD cell lines. However, the downregulation of syntaxin-3 expression levels or perturbation of its intracellular localization was not found in Munc18-2 KD cells, leading to the conclusion that Munc18-2 and syntaxin-3 participate in distinct steps of mast cell degranulation that function independently of each other. Our results, however, clearly indicate that, in Munc18-1/2 DKD RBL-2H3 cells, levels of syntaxin-3 are profoundly reduced – accompanied with its severe mislocalization – in addition to a defective degranulation phenotype. Moreover, the fact that this phenotype was completely rescued following the re-introduction of wild-type Munc18 but not of its binary interaction mutants provide strong evidence that Munc18 participates in the process of mast cell exocytosis by regulating syntaxin-3 expression levels and trafficking. One possible difference between the study by Brochetta et al. and our current study could be that Munc18-2 single KD cells were used in the former study, whereas we used Munc18-1/2 DKD mast cells. The fact that wild-type Munc18-1 can rescue the phenotype of Munc18-1/2 DKD cells indicates that Munc18-1 and Munc18-2 are functionally redundant (Figs 1C, 3, 5, 6). Therefore, in Munc18-2 single KD cells, the remaining Munc18-1 is likely to overtake the role of missing Munc18-2 and is, thus, still providing adequate chaperoning activity toward syntaxin-3.

Although we found that the syntaxin-3 knockdown caused strong impairments in the mast cell degranulation (Fig. 8), we cannot fully attribute the almost complete abolishment of degranulation found Munc18-1/2 DKD mast cells to the reduction in syntaxin-3 levels alone. Moreover, we have previously observed that even Munc18-2 single KD mast cells exhibit severe degranulation defects (Bin et al., 2013). But, in our present study, we show that the level of syntaxin-3 was unchanged in Munc18-2 single KD mast cells (Fig. 1A). Thus, we speculate that there are other possibilities through which Munc18 participates in the mast cell degranulation in addition to the chaperoning of syntaxin-3. For example, Munc18 might regulate other syntaxin isoforms including syntaxin-11 and/or might directly interact with the SNARE complex, thereby controlling the priming step of the exocytosis. Further analyses to test the veracity of these possibilities will provide a clearer view on how Munc18 is involved in mast cell exocytosis.

In contrast to neuronal exocytosis, where fusions occur predominantly between the plasma membrane and the secretory vesicle, mast cells have been shown to undergo compound fusions (multivesicular fusions) allowing substantial release of contents in response to minimal stimulation (Blank and Rivera, 2004). To see the possible role of Munc18 in the compound granule fusions, we performed total internal reflection fluorescence microscopy (TIRFM) with control and Munc18-1/2 DKD RBL-2H3 cells expressing neuropeptide Y (NPY) tagged to the pH-sensitive GFP pHluorin (NPY–pHluorin) (Zhu et al., 2007). In this setting, we can only see the fluorescent signals from the fusion of granules and, by analyzing the intensity as well as radius of fluorescent signals, we can distinguish primary fusion (single granules fusing with the plasma membrane) and compound fusion. After simultaneous stimulation with ionomycin and PMA (phorbol ester analogue), we found rapidly brightening fluorescent signals as a result of fusions of granules in the control cells. We saw that most fusion events were primary fusions, less than 10% of total fusion events were compound fusions. When we performed the same experiments on Munc18-1/2 DKD cells, we found much fewer granule fusions than in the control cells (data not shown). Therefore, we conclude that primary fusion is the main type of fusion event in the RBL-2H3 cells, which was greatly inhibited in the Munc18-1/2 DKD cells.

In neuronal exocytosis, a single binary interaction mutation is not sufficient to disrupt binding between Munc18-1 and monomeric syntaxin-1 (Han et al., 2011; Malintan et al., 2009). However, our present study illustrates that a single binary interaction mutant (K46E or E59K) of either Munc18-1 or Munc18-2 is as detrimental as the double mutant (K46E/E59K) in both binding closed syntaxin-3 (Fig. 4) and rescuing the phenotype of the Munc18-1/2 DKD mast cells (Figs 5, 6). These results are strikingly different from the binary interaction between Munc18-1 and syntaxin-1, and their functional implications. In neuronal exocytosis, compared to wild-type, single binary interaction mutants exhibit only marginally reduced ability to rescue the defective secretion phenotype of Munc18-1/2 DKD PC12 cells and the synaptic transmission of hippocampal neurons that lack Munc18-1 (Han et al., 2011; Han et al., 2009; Meijer et al., 2012). Therefore, our results clearly indicate an important difference in the functional contribution residues Lys46 and Glu59 have on their own in mediating the binary interaction in immune cell compared with neuronal exocytosis. We speculate that, although the importance of the binary interaction between Munc18 and cognate syntaxin partners is well-conserved in both neuronal and immune cell exocytosis, the binding affinity and, thus, its functional outcome are different between the two types of cell. Since even a single mutant was able to fully abrogate the binary interaction between Munc18 and syntaxin-3, we hypothesize that the binary interaction between Munc18 and cognate syntaxin in immune cells is weaker than that in neuronal cells. Therefore, if even one single residue is altered, the whole binary interaction would be abolished because binding by the other residue is not strong enough to sustain the interaction on its own. This theory fits well with our previous finding that the interaction between the N-peptide of syntaxin and the hydrophobic pocket of Munc18 is also functionally important in mast cell degranulation (Bin et al., 2013). Because the binary interaction itself is weak, binding between the Munc18 hydrophobic pocket and the syntaxin N-peptide may work in concert to maintain interaction of Munc18 with its syntaxin partners in mast cells. Thus, unlike neuronal exocytosis, where the latter interaction mode is regarded to be of limited function (Malintan et al., 2009), both modes of interactions are equally valuable molecular actions between Munc18 and syntaxin that are essential for mast cell exocytosis.

A number of point mutations in Munc18-2 were recently identified to be implicated with FHL5, a lethal immune disease during which affected patients' innate immune players, such as natural killer cells and cytotoxic T-lymphocytes are compromised in their degranulation capability (Cetica et al., 2010; Côte et al., 2009; Meeths et al., 2010; zur Stadt et al., 2009). Based on our results illustrating the importance of the chaperoning function of Munc18 toward syntaxin-3 in degranulation, we speculate that those Munc18-2 mutants that are associated with FHL5 are incompetent in the regulation of its chaperoning protein, such as syntaxin-3, to govern correct expression levels as well as intracellular trafficking. Thus, collectively these Munc18-2 mutants might contribute to the lack of proper exocytosis in immune cells of FHL5 patients. As such, our study sheds light on the fact that not only Munc18-2 itself but also the syntaxin protein it chaperones are potentially important therapeutic targets in treating patients with such a devastating immune disorder.

General materials

Parental pLKO-puro plasmid for lentivirus-mediated knockdown was purchased from Sigma-Aldrich (Oakville, ON, Canada). pLVX-IRES-puro plasmid for lentivirus-mediated expression was purchased from Clontech Laboratories (Mountain View, CA). psPAX2 was purchased from Addgene (Cambridge, MA). We obtained rabbit monoclonal anti-syntaxin-3 antibody (clone EPR8543) from Abcam (Cambridge, UK), mouse monoclonal antibodies against GM130 (clone 35), Munc18-1 (clone 31), TGN38 (clone 2) from BD Biosciences (Mississauga, ON, Canada); GAPDH (clone 6C5) from Millipore (Billerica, MA); HA.11 (clone 16B21) from Covance (Princeton, NJ); GFP (clone 9F9.F9) from Novus Biologicals (Lilleton, CO), rabbit polyclonal antibodies against syntaxin-3 from Synaptic Systems (Göttingen, Germany) and syntaxin-11 from Proteintech (Chicago, IL). Rabbit polyclonal anti-Munc18-2 antibody was a kind gift from Dr Vesa Olkkonen.

Construction of Munc18-1 and Munc18-2 expression plasmids

Detailed methods for generating rat Munc18-1 and Munc18-2 expression plasmids have been previously described (Bin et al., 2013). We made binary interaction point mutations (K46E, E59K and K46E/E59K) containing silent nucleotide mutations (SNM) using site-directed mutagenesis. Munc18-1 and Munc18-2 inserts (wild-type and mutants, all without stop codons), which had been digested from pCMV5 vector with EcoRI/XbaI, were subcloned into pLVX–EmGFP–IRES-blast. Sequences of all created plasmids were verified by DNA sequencing. Expression plasmids were co-transfected with psPAX2 and pCMV-VSVG into HEK-293FT cells to generate recombinant lentivirus.

Construction of syxtain-2–HA and syxtain-3–HA expression plasmids

pLVX–HA–IRES-blast was generated by subcloning the hemagglutinin (HA) sequence in tandem into pLVX-IRES-blast using BamHI/XbaI sites. Rat syntaxin-2 and syntaxin-3 without stop codon inserts were digested from pCMV5 vector with EcoRI/XbaI for subcloning into the same sites of pLVX–HA–IRES-blast. Sequences of all created plasmids were verified by DNA sequencing. Expression plasmids were co-transfected with psPAX2 and pCMV-VSVG into HEK-293FT cells in order to generate recombinant lentivirus.

Isolation of stable Munc18-rescued, and syntaxin-2–HA- and syntaxin-3–HA-overexpressing RBL-2H3 cells

Munc18-1/2 DKD RBL-2H3 mast cells (Bin et al., 2013) were maintained in DMEM containing 10% calf serum (HyClone Laboratories, Logan, UT), penicillin (100 U/ml)/streptomycin (0.1 mg/ml) (Sigma-Aldrich) and 250 ng/ml amphotericin B (Sigma-Aldrich). Lentivirus expressing EmGFP (control), EmGFP-tagged Munc18-1 or Munc18-2 wild-type and binary interaction mutants (K46E, E59K, K46E/E59K) were applied to the Munc18-1/2 DKD RBL-2H3 cells in the presence of polybrene (8 µg/ml) (Sigma-Aldrich) for 2 days. For the isolation of cells that were successfully infected with lentivirus, we selected the cells with blasticidin (20 µg/ml) (Invivogen, San Diego, CA). For overexpression of syntaxin-2–HA and syntaxin-3–HA, lentivirus expressing HA-tagged syntaxin-2 or syntaxin-3 were applied to control and Munc18-1/2 DKD RBL-2H3 cells in the presence of polybrene (8 µg/ml) (Sigma-Aldrich) for 2 days; blasticidin (20 µg/ml) (Invivogen) was used for selection.

β-Hexosaminidase-release assays from RBL-2H3 cells

A detailed protocol has been described previously (Bin et al., 2013). Briefly, RBL-2H3 cells were plated in 24-well plates; 2–3 days after plating, cells that were to be stimulated through an IgE-dependent pathway were sensitized with 0.01 µg/ml DNP-IgE Spe7 antibody (Sigma-Aldrich) overnight. The next day, cells were washed twice with physiological saline solution (PSS) containing 145 mM NaCl, 5.6 mM KCl, 2.2 mM CaCl2, 0.5 mM MgCl2, 5.6 mM glucose, 15 mM HEPES, and 0.1% BSA at pH 7.4. Degranulation was stimulated for 1 hour by 50 or 100 ng/ml DNP-HSA (Sigma-Aldrich), 0.5 µM ionomycin (Sigma-Aldrich) or 2.5 µM ionomycin. After transferring the supernatant, cells were solubilized in 0.5% Triton X-100 to obtain the pellet portion. To evaluate release of β-hexosaminidase, 1 mM p-nitrophenyl N-acetyl-β-D-glucosaminide (Sigma-Aldrich) was added and incubated for 1 hour. Reaction was quenched by adding 0.05 µM sodium carbonate buffer and absorbance at 405 nm was measured by using a spectrophotometer; (OD405 of supernatant/OD405 of total)×100% was calculated to be the release of β-hexosaminidase in percent.

Yeast two-hybrid assays

Rat full-length wild-type Munc18-1 or Munc18-2 and one mutant (K46E, E59K or K46E/E59K) with SNM (Bin et al., 2013) were subcloned into the EcoRI/PstI sites of a bait vector, pLexN. The cytoplasmic part of rat syntaxin-2 and syntaxin-3 was subcloned into the EcoRI/BglII sites of prey vector pVP16-3 (Okamoto and Südhof, 1997). The syntaxin-3 open mutant (L165E/E166A) (Dulubova et al., 1999) was generated by site-directed mutagenesis on pVP16-3-rat syntaxin-3. Yeast strain L40 (Vojtek et al., 1993) was transfected with bait and prey vectors by using the lithium-acetate method (Schiestl and Gietz, 1989). Transformants were plated on selection plates lacking uracil, tryptophan and leucine. After 2 days of incubation at 30°C, colonies were inoculated into supplemented minimal medium lacking uracil, tryptophan and leucine, and placed in a shaking incubator at 30°C for 2 days. β-galactosidase assays were performed as described previously (Han et al., 2011).

Cell preparation for confocal immunofluorescence microscopy

Sterilized circular glass coverslips (0.25 mm in width, 1.8 cm in diameter) were placed in 2.2-cm wells within 12-well cell-culture plates. The coverslips were then coated for 1 hour with poly-D-lysine (0.1 mg/ml). RBL-2H3 cells were allowed to adhere to the coverslips for 3 days. For cells overexpressing HA-tagged syntaxin-2 or syntaxin-3, the cells were washed with phosphate-buffered saline (PBS) and fixed for 15 min with PBS containing 4% paraformaldehyde (PFA). Cells were rinsed three times (10 minutes each time) with 1 ml of PBS per well. The fixed cells were permeabilized with PBS containing 0.2% Triton X-100 and 0.3% BSA for 5 minutes, followed by three washes with PBS. Nonspecific sites were blocked for 1 hour at room temperature in PBS containing 0.3% BSA. The mouse monoclonal HA antibody was applied as the primary antibody (1:1000) for 1 hour. After three washes in blocking buffer, Alexa-Fluor-488–conjugated goat anti-mouse secondary antibody (diluted 1:1000; Invitrogen, Carlsbad, CA) was applied for 1 hour. Samples were washed again three times in blocking buffer and 300 nM DAPI was applied for 30 minutes. The slides were mounted in Fluoromount-G reagent (Southern Biotechnology, Birmingham, AL).

For all other stainings, cells were washed once with 0.1% PBS+Tween-20 and fixed for 20 minutes with PBS containing 2% paraformaldehyde. Then, the cells were rinsed twice (10 minutes each time) with 1 ml of 0.1% PBS+Tween-20 followed by permeabilization with 0.1% Triton X-100 for 5 min. Nonspecific sites were blocked for 1 hour at room temperature in PBS containing 10% normal goat serum (Gibco, Burlington, ON, Canada). Then, the primary antibodies were applied at 1:50 for rabbit monoclonal syntaxin-3 antibody, 1:250 for mouse monoclonal GFP antibody, 1:1000 for mouse monoclonal TGN38 and GM130, and 1:1000 for rabbit polyclonal calnexin antibody overnight at room temperature. After three washes in blocking buffer, either Alexa-Fluor-488 or Rhodamine-x-conjugated goat anti-mouse or anti-rabbit secondary antibodies were applied as 1:1000 for 1 hour at room temperature. Samples were washed again three times in blocking buffer and 300 nM DAPI was applied for 30 minutes. Samples were mounted using Fluoromount-G reagent. Immunofluorescence staining of cells overexpressing HA-tagged syntaxin-2 or syntaxin-3 was recorded with a laser confocal scanning microscope (LSM510; Carl Zeiss, Jena, Germany), all other staining was recorded using an LSM710 confocal microscope (Carl Zeiss, Jena, Germany) with an oil immersion objective lens (63×).

Quantification of confocal immunofluorescence microscopy images

The quantification of syntaxin-3 expressions was performed using ImageJ. First, all raw files of confocal images were exported to TIFF then each image was opened on ImageJ. For syntaxin-3 overall intensity, an entire RBL-2H3 cell was manually selected using Freehand Selection, then mean intensity was calculated using the ‘Analyze→Measure’ option. For soma versus tip analysis of syntaxin-3 expression, either soma or tip of the RBL-2H3 cell was selected using Freehand Selection. Then, the intensity was calculated using the ‘Analyze→Measure’ option. The total intensity of syntaxin-3 expression was calculated as above. Then, the value of intensity at the soma or tip was normalized to the total intensity of syntaxin-3 and given in percent. At least 20 cells were analyzed for each knockdown or rescue experiment and the mean values were presented.

Construction of syntaxin-3 knockdown plasmids and isolation of stable syntaxin-3 knockdown RBL-2H3 cells

To knock down the rat syntaxin-3 gene, we generated two different constructs targeting (1) the 21-nucleotide sequence 5′-GCTCGAAAGAAATTGATAATT-3′ (named syntaxin-3–8 KD) and, (2) the 21-nucleotide sequence 5′-GTTTGTGGAGGTGATGACAAA-3′ (named syntaxin-3–5 KD) in rat syntaxin-3. The hexameric sequence CTCGAG was used as a linker sequence in both constructs. These oligonucleotides, containing sense and antisense of the target sequences, were annealed, phosphorylated and ligated into AgeI/EcoRI sites of pLKO-puro (Sigma-Aldrich) for syntaxin-3–8 KD and pLKO-neo (Sigma-Aldrich) for syntaxin-3–5 KD, generating the syntaxin-3 knockdown plasmids. These knockdown plasmids were co-transfected with psPAX2 and pMD.G into HEK-293FT cells to generate recombinant lentivirus. Then the lentivirus, expressing shRNA against rat syntaxin-3 and an empty vector control, was applied to RBL-2H3 wild-type cells (ATCC, Manassas, VA) in the presence of polybrene (8 µg/ml) (Sigma-Aldrich) for 2 days, and infected cells were isolated by using puromycin (10 µg/ml) (Bioshop, Burlington, ON, Canada).

Homology modeling of Munc18-2 and syntaxin-3

The homology model of rat Munc18-2 (GenBank accession no. U20283.1) and rat syntaxin-3 (GenBank accession no. L20820.1) was constructed based on a template crystal structure of neuronal Sec1–syntaxin-1A complex (PDB 3C98) (Misura et al., 2000). The aa sequences of Munc18-2 and syntaxin-3 were inputted into Zdock Server software and the predicted structure of the protein complex was visualized using Jmol software.

Visualization of granule fusion using TIRFM

The control and Munc18-1/2 DKD RBL-2H3 cells were transiently expressed with of NPY–pHluorin by electroporating 15 µg of pcDNA3.1-NPY–pHluorin (Zhu et al., 2007). After 2 days, cells were plated on the 25 mm circular glass coverslip coated with poly-D-lysine (0.1 mg/ml) and were allowed to attach for 2 days. Total internal reflection fluorescence microscopy (TIRFM) setup, image acquisition and data analysis were performed as described here (Xie et al., 2013). Stimulation of exocytosis was triggered by adding 2.5 µM of ionomycin and 0.1 µM of PMA.

We thank Dr Vesa Olkkonen (Minerva Foundation Institute for Medical Research, Helsinki, Finland) for reagents used in this study.

Author contributions

N.-R.B. mainly contributed to designing, performing and analyzing the experiments, and to writing of the manuscript. C.H.J. generated syntaxin 3-8 knockdown RBL-2H3 cells. B.K. contributed to the confocal microscopy experiments. P.C. contributed to writing of the manuscript. E.T. and H.-S.S. provided conceptual help, and analyzed confocal microscopy experiments. D.Z. and H.Y.G. provided conceptual help and analyzed TIRFM experiments. S.S. contributed to designing and analyzing the experiments, and writing of the manuscript.

Funding

This research was supported by the Natural Sciences and Engineering Research Council of Canada [grant number 298461-09], the Heart and Stroke Foundation [grant number T6700], and the Canadian Institute of Health Research [grant number MOP-130573]. The Alexander Graham Bell Canada Graduate Scholarships-Doctoral (CGS D) from Natural Sciences and Engineering Research Council of Canada (NSERC) was issued to N.-R.B.

Arunachalam
,
L.
,
Han
,
L.
,
Tassew
,
N. G.
,
He
,
Y.
,
Wang
,
L.
,
Xie
,
L.
,
Fujita
,
Y.
,
Kwan
,
E.
,
Davletov
,
B.
,
Monnier
,
P. P.
, et al. 
(
2008
).
Munc18-1 is critical for plasma membrane localization of syntaxin1 but not of SNAP-25 in PC12 cells
.
Mol. Biol. Cell
19
,
722
-
734
.
Bin
,
N. R.
,
Jung
,
C. H.
,
Piggott
,
C.
and
Sugita
,
S.
(
2013
).
Crucial role of the hydrophobic pocket region of Munc18 protein in mast cell degranulation
.
Proc. Natl. Acad. Sci. USA
110
,
4610
-
4615
.
Blank
,
U.
and
Rivera
,
J.
(
2004
).
The ins and outs of IgE-dependent mast-cell exocytosis
.
Trends Immunol.
25
,
266
-
273
.
Brochetta
,
C.
,
Suzuki
,
R.
,
Vita
,
F.
,
Soranzo
,
M. R.
,
Claver
,
J.
,
Madjene
,
L. C.
,
Attout
,
T.
,
Vitte
,
J.
,
Varin-Blank
,
N.
,
Zabucchi
,
G.
, et al. 
(
2014
).
Munc18-2 and syntaxin 3 control distinct essential steps in mast cell degranulation
.
J. Immunol.
192
,
41
-
51
.
Cetica
,
V.
,
Santoro
,
A.
,
Gilmour
,
K. C.
,
Sieni
,
E.
,
Beutel
,
K.
,
Pende
,
D.
,
Marcenaro
,
S.
,
Koch
,
F.
,
Grieve
,
S.
,
Wheeler
,
R.
, et al. 
(
2010
).
STXBP2 mutations in children with familial haemophagocytic lymphohistiocytosis type 5
.
J. Med. Genet.
47
,
595
-
600
.
Côte
,
M.
,
Ménager
,
M. M.
,
Burgess
,
A.
,
Mahlaoui
,
N.
,
Picard
,
C.
,
Schaffner
,
C.
,
Al-Manjomi
,
F.
,
Al-Harbi
,
M.
,
Alangari
,
A.
,
Le Deist
,
F.
, et al. 
(
2009
).
Munc18-2 deficiency causes familial hemophagocytic lymphohistiocytosis type 5 and impairs cytotoxic granule exocytosis in patient NK cells
.
J. Clin. Invest.
119
,
3765
-
3773
.
Dulubova
,
I.
,
Sugita
,
S.
,
Hill
,
S.
,
Hosaka
,
M.
,
Fernandez
,
I.
,
Südhof
,
T. C.
and
Rizo
,
J.
(
1999
).
A conformational switch in syntaxin during exocytosis: role of munc18
.
EMBO J.
18
,
4372
-
4382
.
Hackmann
,
Y.
,
Graham
,
S. C.
,
Ehl
,
S.
,
Höning
,
S.
,
Lehmberg
,
K.
,
Aricò
,
M.
,
Owen
,
D. J.
and
Griffiths
,
G. M.
(
2013
).
Syntaxin binding mechanism and disease-causing mutations in Munc18-2
.
Proc. Natl. Acad. Sci. USA
110
,
E4482
-
E4491
.
Han
,
L.
,
Jiang
,
T.
,
Han
,
G. A.
,
Malintan
,
N. T.
,
Xie
,
L.
,
Wang
,
L.
,
Tse
,
F. W.
,
Gaisano
,
H. Y.
,
Collins
,
B. M.
,
Meunier
,
F. A.
, et al. 
(
2009
).
Rescue of Munc18-1 and -2 double knockdown reveals the essential functions of interaction between Munc18 and closed syntaxin in PC12 cells
.
Mol. Biol. Cell
20
,
4962
-
4975
.
Han
,
G. A.
,
Malintan
,
N. T.
,
Collins
,
B. M.
,
Meunier
,
F. A.
and
Sugita
,
S.
(
2010
).
Munc18-1 as a key regulator of neurosecretion
.
J. Neurochem.
115
,
1
-
10
.
Han
,
G. A.
,
Malintan
,
N. T.
,
Saw
,
N. M.
,
Li
,
L.
,
Han
,
L.
,
Meunier
,
F. A.
,
Collins
,
B. M.
and
Sugita
,
S.
(
2011
).
Munc18-1 domain-1 controls vesicle docking and secretion by interacting with syntaxin-1 and chaperoning it to the plasma membrane
.
Mol. Biol. Cell
22
,
4134
-
4149
.
Han
,
G. A.
,
Bin
,
N. R.
,
Kang
,
S. Y.
,
Han
,
L.
and
Sugita
,
S.
(
2013
).
Domain 3a of Munc18-1 plays a crucial role at the priming stage of exocytosis
.
J. Cell Sci.
126
,
2361
-
2371
.
Han
,
G. A.
,
Park
,
S.
,
Bin
,
N. R.
,
Jung
,
C. H.
,
Kim
,
B.
,
Chandrasegaram
,
P.
,
Matsuda
,
M.
,
Riadi
,
I.
,
Han
,
L.
and
Sugita
,
S.
(
2014
).
A pivotal role for Pro335 in balancing the dual functions of Munc18-1 domain-3a in regulated exocytosis
.
J. Biol. Chem. 289, 33617-33628.
Hu
,
S. H.
,
Christie
,
M. P.
,
Saez
,
N. J.
,
Latham
,
C. F.
,
Jarrott
,
R.
,
Lua
,
L. H.
,
Collins
,
B. M.
and
Martin
,
J. L.
(
2010
).
Possible roles for Munc18-1 domain 3a and Syntaxin1 N-peptide and C-terminal anchor in SNARE complex formation
.
Proc. Natl. Acad. Sci. USA
108
,
1040
-
1045
.
Kim
,
K.
,
Petrova
,
Y. M.
,
Scott
,
B. L.
,
Nigam
,
R.
,
Agrawal
,
A.
,
Evans
,
C. M.
,
Azzegagh
,
Z.
,
Gomez
,
A.
,
Rodarte
,
E. M.
,
Olkkonen
,
V. M.
, et al. 
(
2012
).
Munc18b is an essential gene in mice whose expression is limiting for secretion by airway epithelial and mast cells
.
Biochem. J.
446
,
383
-
394
.
Ma
,
C.
,
Su
,
L.
,
Seven
,
A. B.
,
Xu
,
Y.
and
Rizo
,
J.
(
2013
).
Reconstitution of the vital functions of Munc18 and Munc13 in neurotransmitter release
.
Science
339
,
421
-
425
.
Malintan
,
N. T.
,
Nguyen
,
T. H.
,
Han
,
L.
,
Latham
,
C. F.
,
Osborne
,
S. L.
,
Wen
,
P. J.
,
Lim
,
S. J.
,
Sugita
,
S.
,
Collins
,
B. M.
and
Meunier
,
F. A.
(
2009
).
Abrogating Munc18-1-SNARE complex interaction has limited impact on exocytosis in PC12 cells
.
J. Biol. Chem.
284
,
21637
-
21646
.
McEwen
,
J. M.
and
Kaplan
,
J. M.
(
2008
).
UNC-18 promotes both the anterograde trafficking and synaptic function of syntaxin
.
Mol. Biol. Cell
19
,
3836
-
3846
.
Medine
,
C. N.
,
Rickman
,
C.
,
Chamberlain
,
L. H.
and
Duncan
,
R. R.
(
2007
).
Munc18-1 prevents the formation of ectopic SNARE complexes in living cells
.
J. Cell Sci.
120
,
4407
-
4415
.
Meeths
,
M.
,
Entesarian
,
M.
,
Al-Herz
,
W.
,
Chiang
,
S.
,
Wood
,
S.
,
Al-Ateeqi
,
W.
,
Almazan
,
F.
,
Boelens
,
J.
,
Hasle
,
H.
,
Ifversen
,
M.
, et al. 
(
2010
).
Spectrum of clinical presentations in familial hemophagocytic lymphohistiocytosis (FHL) type 5 patients with mutations in STXBP2
.
Blood
116
,
2635
-
2643
.
Meijer
,
M.
,
Burkhardt
,
P.
,
de Wit
,
H.
,
Toonen
,
R. F.
,
Fasshauer
,
D.
and
Verhage
,
M.
(
2012
).
Munc18-1 mutations that strongly impair SNARE-complex binding support normal synaptic transmission
.
EMBO J.
31
,
2156
-
2168
.
Misura
,
K. M.
,
Scheller
,
R. H.
and
Weis
,
W. I.
(
2000
).
Three-dimensional structure of the neuronal-Sec1-syntaxin 1a complex
.
Nature
404
,
355
-
362
.
Okamoto
,
M.
and
Südhof
,
T. C.
(
1997
).
Mints, Munc18-interacting proteins in synaptic vesicle exocytosis
.
J. Biol. Chem.
272
,
31459
-
31464
.
Parisotto
,
D.
,
Pfau
,
M.
,
Scheutzow
,
A.
,
Wild
,
K.
,
Mayer
,
M. P.
,
Malsam
,
J.
,
Sinning
,
I.
and
Söllner
,
T. H.
(
2014
).
An extended helical conformation in domain 3a of Munc18-1 provides a template for SNARE (soluble N-ethylmaleimide-sensitive factor attachment protein receptor) complex assembly
.
J. Biol. Chem.
289
,
9639
-
9650
.
Richmond
,
J. E.
,
Weimer
,
R. M.
and
Jorgensen
,
E. M.
(
2001
).
An open form of syntaxin bypasses the requirement for UNC-13 in vesicle priming
.
Nature
412
,
338
-
341
.
Riento
,
K.
,
Galli
,
T.
,
Jansson
,
S.
,
Ehnholm
,
C.
,
Lehtonen
,
E.
and
Olkkonen
,
V. M.
(
1998
).
Interaction of Munc-18-2 with syntaxin 3 controls the association of apical SNAREs in epithelial cells
.
J. Cell Sci.
111
,
2681
-
2688
.
Rodkey
,
T. L.
,
Liu
,
S.
,
Barry
,
M.
and
McNew
,
J. A.
(
2008
).
Munc18a scaffolds SNARE assembly to promote membrane fusion
.
Mol. Biol. Cell
19
,
5422
-
5434
.
Rowe
,
J.
,
Corradi
,
N.
,
Malosio
,
M. L.
,
Taverna
,
E.
,
Halban
,
P.
,
Meldolesi
,
J.
and
Rosa
,
P.
(
1999
).
Blockade of membrane transport and disassembly of the Golgi complex by expression of syntaxin 1A in neurosecretion-incompetent cells: prevention by rbSEC1
.
J. Cell Sci.
112
,
1865
-
1877
.
Rowe
,
J.
,
Calegari
,
F.
,
Taverna
,
E.
,
Longhi
,
R.
and
Rosa
,
P.
(
2001
).
Syntaxin 1A is delivered to the apical and basolateral domains of epithelial cells: the role of munc-18 proteins
.
J. Cell Sci.
114
,
3323
-
3332
.
Schiestl
,
R. H.
and
Gietz
,
R. D.
(
1989
).
High efficiency transformation of intact yeast cells using single stranded nucleic acids as a carrier
.
Curr. Genet.
16
,
339
-
346
.
Shen
,
J.
,
Tareste
,
D. C.
,
Paumet
,
F.
,
Rothman
,
J. E.
and
Melia
,
T. J.
(
2007
).
Selective activation of cognate SNAREpins by Sec1/Munc18 proteins
.
Cell
128
,
183
-
195
.
Südhof
,
T. C.
and
Rothman
,
J. E.
(
2009
).
Membrane fusion: grappling with SNARE and SM proteins
.
Science
323
,
474
-
477
.
Tadokoro
,
S.
,
Kurimoto
,
T.
,
Nakanishi
,
M.
and
Hirashima
,
N.
(
2007
).
Munc18-2 regulates exocytotic membrane fusion positively interacting with syntaxin-3 in RBL-2H3 cells
.
Mol. Immunol.
44
,
3427
-
3433
.
Tamori
,
Y.
,
Kawanishi
,
M.
,
Niki
,
T.
,
Shinoda
,
H.
,
Araki
,
S.
,
Okazawa
,
H.
and
Kasuga
,
M.
(
1998
).
Inhibition of insulin-induced GLUT4 translocation by Munc18c through interaction with syntaxin4 in 3T3-L1 adipocytes
.
J. Biol. Chem.
273
,
19740
-
19746
.
Tareste
,
D.
,
Shen
,
J.
,
Melia
,
T. J.
and
Rothman
,
J. E.
(
2008
).
SNAREpin/Munc18 promotes adhesion and fusion of large vesicles to giant membranes
.
Proc. Natl. Acad. Sci. USA
105
,
2380
-
2385
.
Verhage
,
M.
,
Maia
,
A. S.
,
Plomp
,
J. J.
,
Brussaard
,
A. B.
,
Heeroma
,
J. H.
,
Vermeer
,
H.
,
Toonen
,
R. F.
,
Hammer
,
R. E.
,
van den Berg
,
T. K.
,
Missler
,
M.
, et al. 
(
2000
).
Synaptic assembly of the brain in the absence of neurotransmitter secretion
.
Science
287
,
864
-
869
.
Vojtek
,
A. B.
,
Hollenberg
,
S. M.
and
Cooper
,
J. A.
(
1993
).
Mammalian Ras interacts directly with the serine/threonine kinase Raf
.
Cell
74
,
205
-
214
.
Xie
,
L.
,
Zhu
,
D.
,
Kang
,
Y.
,
Liang
,
T.
,
He
,
Y.
and
Gaisano
,
H. Y.
(
2013
).
Exocyst sec5 regulates exocytosis of newcomer insulin granules underlying biphasic insulin secretion
.
PLoS ONE
8
,
e67561
.
Xu
,
Y.
,
Su
,
L.
and
Rizo
,
J.
(
2010
).
Binding of Munc18-1 to synaptobrevin and to the SNARE four-helix bundle
.
Biochemistry
49
,
1568
-
1576
.
Zhu
,
D.
,
Zhou
,
W.
,
Liang
,
T.
,
Yang
,
F.
,
Zhang
,
R. Y.
,
Wu
,
Z. X.
and
Xu
,
T.
(
2007
).
Synaptotagmin I and IX function redundantly in controlling fusion pore of large dense core vesicles
.
Biochem. Biophys. Res. Commun.
361
,
922
-
927
.
zur Stadt
,
U.
,
Rohr
,
J.
,
Seifert
,
W.
,
Koch
,
F.
,
Grieve
,
S.
,
Pagel
,
J.
,
Strauss
,
J.
,
Kasper
,
B.
,
Nürnberg
,
G.
,
Becker
,
C.
, et al. 
(
2009
).
Familial hemophagocytic lymphohistiocytosis type 5 (FHL-5) is caused by mutations in Munc18-2 and impaired binding to syntaxin 11
.
Am. J. Hum. Genet.
85
,
482
-
492
.

Competing interests

The authors declare no competing or financial interests.

Supplementary information