Intracellular membrane fusion occurs with exquisite coordination and specificity. Each fusion event requires three basic components: Rab-GTPases organize the fusion site; SNARE proteins act during fusion; and N-ethylmaleimide-sensitive factor (NSF) plus its cofactor α-SNAP are required for recycling or activation of the fusion machinery. Whereas Rab-GTPases seem to mediate the initial membrane contact, SNAREs appear to lie at the center of the fusion process. It is known that formation of complexes between SNAREs from apposed membranes is a prerequisite for lipid bilayer mixing; however, the biophysics and many details of SNARE function are still vague. Nevertheless, recent observations are shedding light on the role of SNAREs in membrane fusion. Structural studies are revealing the mechanisms by which SNARES form complexes and interact with other proteins. Furthermore, it is now apparent that the SNARE transmembrane segment not only anchors the protein but engages in SNARE-SNARE interactions and plays an active role in fusion. Recent work indicates that the fusion process itself may comprise two stages and proceed via a hemifusion intermediate.

Eukaryotic cells contain numerous compartments that are separated from the cytosol by a lipid bilayer. Exchange of lipids, metabolites or proteins across the lipid bilayer can occur through integral membrane proteins - as demonstrated for peroxisomes, mitochondria or chloroplasts. Within the endomembrane system, which includes the endoplasmic reticulum (ER), the Golgi, endosomes and lysosomes, vesicles transport cargo between organelles and thus mediate exchange (Bonifacino and Glick, 2004; Mellman and Warren, 2000). Generally, vesicle formation requires a conserved set of coat proteins that bind to cargo and induce membrane curvature. At the same time, a vesicle must incorporate proteins that target it to the right compartment and enable it to fuse with the target compartment (Fig. 1A). Thus, vesicle formation requires the uptake of targeting and fusion factors, including Rab-GTPases, tethering factors and SNAREs [soluble NSF (N-ethylmaleimide-sensitive factor) attachment protein receptors]. Fusion seems to involve a cascade in which a Rab-GTPase, together with tethering factors, mediates membrane contact, which is followed by SNARE pairing and lipid bilayer mixing. SNAREs seem to operate at all stages. They need to be incorporated into the right vesicle, interact with tethers, and are essential for tight membrane docking and bilayer mixing. Here, we focus primarily on their role in exocytosis and organelle fusion.

SNAREs are central to the fusion of endomembranes (Burri and Lithgow, 2004; Chen and Scheller, 2001; Jahn and Grubmüller, 2002; Söllner, 2004). Upon membrane contact, they form defined trans-SNARE complexes, also known as `SNAREpins' (Weber et al., 1998). In general, SNAREs consist of a central `SNARE domain' that is flanked by a variable N-terminal domain and a C-terminal single α-helical transmembrane anchor (Fig. 1B). SNAREs were initially classified according to their preferential localization as vesicle-localized (v)- or target-membrane-bound (t)-SNAREs (Rothman, 1994). This nomenclature turned out to be somewhat ambiguous, since t-SNAREs are also found on vesicles and v-SNAREs can be found on target membranes. A systematic sequence analysis revealed that most v-SNAREs have an arginine residue in the centre of the SNARE domain (R-SNAREs), whereas a glutamine (or aspartate) residue is found in syntaxins and SNAP-25-like proteins (Q-SNAREs) (Weimbs et al., 1997, Fasshauer et al., 1998). Further refinement of this analysis lead to the classification of three Q-SNARE groups: Qa, Qb and Qc (Bock et al., 2001). The crystal structures of complexes formed by fragments of synaptic and endosomal SNAREs have been solved (Antonin et al., 2002b; Sutton et al., 1998). In each case, the complex consists of four SNARE domains forming a coiled-coil: one contributed by the R-SNARE and one each by Qa-, Qb- and Qc-SNAREs. This composition of SNARE complexes is conserved: synaptic and yeast vacuole SNARE complexes each contain four helices (Sutton et al., 1998; Fukuda et al., 2000; Dietrich et al., 2005a). A closer analysis shows that the coiled-coil structure is composed of 15 hydrophobic layers, and a central hydrophilic zero-layer containing the R- and Q-residues (Sutton et al., 1998). The positions of the R and Q residues can be swapped among the different helices as long as the 3Q:1R ratio is maintained (Katz and Brennwald, 2000; Graf et al., 2005), which indicates that it is a hallmark of specificity in SNARE function.

The N-terminal domains of SNAREs show some variability (Duetruch et al., 2003). Two predominant folds are found. The N-terminus of Q-SNAREs, as far as it has been clarified structurally, consists of a three-helix bundle (Fig. 1C) (Fernandez et al., 1998; Nicholson et al., 1998; Dulubova et al., 2001; Antonin et al., 2002a). This can interact with the coiled-coil domain to control its interaction with other SNAREs and Sec1/Munc18 proteins (Fernandez et al., 1998; Nicholson et al., 1998; Rizo and Südhof, 2002). A subgroup of the R-SNAREs has a more complex longin fold, as discussed below (Fig. 1D) (Fernandez et al., 1998; Nicholson et al., 1998; Rizo and Südhof, 2002).

Fig. 1.

SNARE function in trafficking and SNARE domain structure. (A) Vesicle trafficking. The stages of vesicle budding and fusion are shown. During SNARE activation, SNAREs become disassembled with the help of NSF, α-SNAP and ATP. Budding includes uptake of the v-SNARE, Rab proteins and tethers, which assemble before or during tethering into a complete complex. Fusion requires SNARE complex assembly and lipid mixing. (B) Domain structure of selected SNAREs. Lipid anchors on Ykt6 and SNAP-25 are indicated. (C) The HABC N-terminal domain of syntaxin (PDB accession number 1BR0) (Fernandez et al., 1998). (D) The Ykt6 N-terminal longin domain (1H8M) (Tochio et al., 2001).

Fig. 1.

SNARE function in trafficking and SNARE domain structure. (A) Vesicle trafficking. The stages of vesicle budding and fusion are shown. During SNARE activation, SNAREs become disassembled with the help of NSF, α-SNAP and ATP. Budding includes uptake of the v-SNARE, Rab proteins and tethers, which assemble before or during tethering into a complete complex. Fusion requires SNARE complex assembly and lipid mixing. (B) Domain structure of selected SNAREs. Lipid anchors on Ykt6 and SNAP-25 are indicated. (C) The HABC N-terminal domain of syntaxin (PDB accession number 1BR0) (Fernandez et al., 1998). (D) The Ykt6 N-terminal longin domain (1H8M) (Tochio et al., 2001).

Among R-SNAREs, the absence of the longin-fold distinguishes brevins (synaptobrevin-like R-SNAREs) from longins (Filippini et al., 2001). Brevins have short, presumably unstructured, N-terminal extensions. The longin fold consists of five central β-sheets sandwiched by two α-helices on one face and one α-helix on the opposite face (Gonzalez et al., 2001; Tochio et al., 2001). Three longin proteins have been characterized in detail: Sec22 is required for trafficking between the ER and the Golgi, and for transport within the Golgi; Ykt6 is required along the Golgi, in endosomes and in the vacuole/lysosome system; and Ti-VAMP is required in neuronal cells (Rossi et al., 2004). The distribution of Sec22 and Ykt6 across the endomembrane system in yeast suggests that longins have a general role along the secretory pathway and in fusion reactions. The longin-fold is not unique to the three SNAREs, and has also been identified in the SEDL (Trs31) subunit of the TRAPP tethering complex (Jang et al., 2002), the σ2 and μ2 subunits of the adaptor protein (AP)-2 coat (Collins et al., 2002), the SRX domain of the α-subunit of the signal recognition particle (Schwartz and Blobel, 2003), and two endosome-associated mitogen-activated protein (MAP) kinase scaffold proteins, p15 and MP-1 (Kurzbauer et al., 2004; Lunin et al., 2004). This suggests that it is a general signature of proteins operating in trafficking reactions (for a more detailed discussion, see Rossi et al., 2004).

SNAREs are anchored to membranes by transmembrane segments (TMSs) that not only anchor them, but also contribute to SNARE-SNARE interactions and appear to play an active role in the fusion process. The synaptic R-SNARE synaptobrevin II and the Q-SNARE syntaxin 1A homo- and heterodimerize in vitro through a conserved motif within their respective α-helical TMSs (Laage and Langosch, 1997; Margittai et al., 1999; Laage et al., 2000; Roy et al., 2004) that forms a tightly packed interface (Fleming and Engelman, 2001). The long axes of these transmembrane helices are proposed to cross each other at a negative angle; by contrast, the soluble helices of the cytoplasmic SNARE domains assume a positive packing angle upon formation of the coiled-coil structure of the assembled SNARE complex. The sequence linking both domains is unstructured in syntaxin (Kim et al., 2002; Knecht and Grubmüller, 2003; Margittai et al., 2003). Homodimerization through TMS-TMS interactions, although not involving the same motif, is conserved in a yeast vacuolar Q-SNARE (R. Roy, J. Rohde, K. Peplowska, C.U. and D.L., unpublished).

A few SNAREs lack a TMS. The Q-SNAREs SNAP-25 and SNAP-23 bind to membranes through palmitate anchors (Hess et al., 1992; Vogel and Roche, 1999). Ykt6 has a palmitate and farnesyl anchor for membrane binding (Fukasawa et al., 2004; Hasegawa et al., 2004; Dietrich et al., 2005b), and the yeast Q-SNARE Vam7p has an N-terminal PX domain that binds to phosphoinositides on vacuoles (Cheever et al., 2001; Boeddinghaus et al., 2002). These alternative anchors might facilitate regulation of membrane association and the fusion reaction. Indeed, cytosolic intermediates of these SNAREs have been described. However, SNARE-mediated fusion can occur only if at least one SNARE on each bilayer is anchored by a TMS (McNew et al., 1999; Grote et al., 2000; Rohde et al., 2003).

Mechanism of membrane fusion by SNAREs SNARE pairing

The SNARE hypothesis proposed that complex formation by SNARE proteins can explain the specificity of vesicle fusion (Rothman, 1994). SNAREs residing on the target organelle and vesicle would bind to each other, and fusion would be triggered by NSF-driven disassembly of the SNARE complex. Despite being a major breakthrough, many aspects of this hypothesis have been challenged and modified. For example, the specificity of SNARE complex formation is lower than originally thought: the cytoplasmic domains of isoforms that correspond to organelles that do not fuse to each other in vivo efficiently co-assemble in vitro (Fasshauer et al., 1999; Yang et al., 1999; Tsui and Banfield, 2000). However, the specificity of SNARE pairing appears to be higher in the cell (Scales et al., 2000) and upon reconstitution of full-length SNAREs into liposomes (McNew et al., 2000a). Nevertheless, SNAREs can functionally replace each other in vivo (Götte and Gallwitz, 1997; Liu and Barlowe, 2002; Borisovska et al., 2005) and participate in multiple reactions (Fasshauer, 2003). The specificity of organelle-organelle recognition in the cell is at least in part due to Rab proteins and tethering factors that are likely to act upstream of SNARE complex formation to establish contact between cognate membranes (Fig. 1A). Proteins discussed as tethers include Uso1p, the VFT complex, TRAPP, the exocyst, the HOPS complex and EEA1 (Gillingham and Munro, 2003; Whyte and Munro, 2002; Zerial and McBride, 2001). Tethers can bind to SNARE proteins (Shorter et al., 2002; Collins et al., 2005), and complexes including SNAREs and proposed tethers have been identified (Whyte and Munro, 2002). Since such proteins contain large coiled-coil domains, they may direct assembly of SNARE complexes at fusion sites.

Some of the reported discrepancies might be due to inherent differences in the fusion reactions of evolutionarily distant organisms such as yeast, invertebrates and mammals. Moreover, different types of fusion - such as organelle-organelle fusion or neurotransmitter release - might have different requirements. It is quite likely that evolution has fine-tuned individual SNAREs for specific functions. For example, fusion of fragmented organelles upon cell division is symmetrical since identical membranes fuse (Wickner and Haas, 2000; Shorter and Warren, 2002). Symmetrical fusion has a slow time course, involves (at least for some organelles) membranes of low intrinsic curvature and may not need a sophisticated level of regulation. By contrast, fusion of secretory vesicles to the plasma membrane is asymmetric and might occur constitutively. An extreme case is the regulated release of neurotransmitter from presynaptic nerve terminals: this is extraordinarily fast, involves small, highly curved synaptic vesicles and is precisely regulated by Ca2+ (Südhof, 2004).

Nature has provided a number of solutions to these different requirements. One obvious difference between fusion at or between organelles and fast secretion is that the former exclusively employs SNAREs with single SNARE motifs, whereas the latter requires SNAP-25 orthologs, which combine two SNARE domains in one protein (Fig. 1B). In organelle-organelle fusion, cis complexes of R- and Q-SNAREs exist prior to fusion (Wickner and Haas, 2000). Their dissociation by the AAA-ATPase NSF and its cofactor α-SNAP is thus required prior to assembly of trans complexes that link the membranes (Mayer et al., 1996). This is exemplified by fusion between endosomes (McBride et al., 1999), ER-derived vesicles (Barlowe, 1997) and yeast vacuoles (Nichols et al., 1997; Ungermann et al., 1998; Dietrich et al., 2005a). In neurotransmitter release, only part of synaptobrevin complexes with Q-SNAREs on the vesicle (Kretzschmar et al., 1996). The rest appears either to associate with the synaptic vesicle protein synaptophysin or to homodimerize (Calakos and Scheller, 1994; Edelmann et al., 1995; Washbourne et al., 1995; Pennuto et al., 2002; Roy et al., 2004). At the plasma membrane, syntaxin and SNAP-25 seem to assemble in cis prior to trans complex formation with the R-SNARE (Pabst et al., 2002; An and Almers, 2004; Fasshauer and Margittai, 2004).

The precise mechanism of trans complex formation might also depend on the type of fusion and on its level of regulation. The SNARE domains of exocytotic SNAREs from yeast appear to associate in a one-step reaction over their entire length (Zhang et al., 2004). By contrast, trans assembly of neuronal SNARE motifs probably starts at their N-termini and proceeds towards the TMSs in a `zippering-up' process (Chen et al., 1999; Melia et al., 2002). Since the assembly occurs in a sequential pathway, populations of vesicles that exocytose with different kinetics might reflect different assembly stages (Xu et al., 1999). Surprisingly, trans assembly of synaptic SNAREs reconstituted in liposomal membranes at low density is prohibited by close association of the R-SNARE domain with the bilayer through membrane-proximal tryptophan residues (Hu et al., 2002; Kweon et al., 2003). Both, zippering-up and the availability of the R-SNARE domain might thus be subject to regulation by as-yet-unknown mechanisms.

SNARE regulation

Several proteins that control SNARE complex assembly have been identified (Fig. 2). Complexins, for example, can bind to assembled SNARE complexes and modulate exocytosis (Fig. 2E) (Chen et al., 2002; McMahon et al., 1995; Pabst et al., 2002). Tomosyn and amisyn interact with syntaxin, SNAP-25 and synaptotagmin and might facilitate assembly of SNARE complexes (Fujita et al., 1998; Pobbati et al., 2004; Scales et al., 2002). Furthermore, Sec1/Munc18 (SM) proteins regulate the availability of syntaxins for SNARE complex formation at the fusion site (Rizo and Südhof, 2002). At the synapse, Munc18 binds to the closed conformation of syntaxin (Fig. 2A) (Yang et al., 2000). Displacement of Munc18 is then required prior to SNARE complex assembly (Fig. 2D) (Sutton et al., 1998; Rizo and Südhof, 2002). How this transition is mediated is still unknown. Removing the N-terminal domain of syntaxin or mutating the linker between the N-terminus and the coiled-coil domain accelerates SNARE complex assembly dramatically (Nicholson et al., 1998; Parlati et al., 1999). In yeast, the SM protein Sec1 can bind to the assembled exocytic SNARE complex and stimulate liposome fusion (Carr et al., 1999; Scott et al., 2004), providing a variation of SM protein function. Several studies have identified a small peptide at the N-terminus of syntaxins as the binding site for SM proteins (Fig. 2B) (Bracher and Weissenhorn, 2002; Dulubova et al., 2002; Dulubova et al., 2003). Surprisingly, mutations that abolish the interaction of the SM protein Sly1 and the Golgi SNARE Sed5 in vitro do not seem to result in an obvious in vivo defect (Peng and Gallwitz, 2004). Thus, establishing the precise function of SM proteins during fusion will require additional studies.

Fig. 2.

SNARE complex assembly and its control. (A) Interaction of syntaxin with nSec1, also known as Munc18 (1DN1) (Yang et al., 2000). nSec1 binds to the closed conformation of syntaxin, regulating its availability. (B) Binding of Sly1 to the N-terminal peptide of the SNARE Sed5 may function similarly (1MQS) (Bracher and Weissenhorn, 2002). (C) The open conformation of the syntaxin molecule (as shown in Fig. 1C). (D) The assembled SNARE complex modeled between two membranes (1SFC) (Sutton et al., 1998). (E) Interaction of complexin with the SNARE complex might regulate exocytosis (1KIL) (Chen et al., 2002).

Fig. 2.

SNARE complex assembly and its control. (A) Interaction of syntaxin with nSec1, also known as Munc18 (1DN1) (Yang et al., 2000). nSec1 binds to the closed conformation of syntaxin, regulating its availability. (B) Binding of Sly1 to the N-terminal peptide of the SNARE Sed5 may function similarly (1MQS) (Bracher and Weissenhorn, 2002). (C) The open conformation of the syntaxin molecule (as shown in Fig. 1C). (D) The assembled SNARE complex modeled between two membranes (1SFC) (Sutton et al., 1998). (E) Interaction of complexin with the SNARE complex might regulate exocytosis (1KIL) (Chen et al., 2002).

Regulation by Ca2+ is crucial for several fusion reactions. Calmodulin and synaptotagmin can act as Ca2+ sensors in several systems, although their precise roles are still debated (Südhof, 2002; Burgoyne and Clague, 2003). The role of Ca2+ in regulated exocytosis at synapses has been analyzed in detail. Regulation of neurotransmitter release by presynaptic Ca2+ influx is likely to be mediated by synaptotagmin in the vesicular membrane (Fernandez-Chacon et al., 2001; Tucker et al., 2004; Südhof, 2004), which is physically coupled to the SNAREs (Bai et al., 2004). Synaptotagmin is anchored to membranes by a transmembrane domain and has two Ca2+-binding domains (Südhof, 2004). Synaptotagmin binds to SNAREs and lipids in a Ca2+-dependent manner, which might trigger fusion (Tucker et al., 2003; Bai et al., 2004). Interestingly, in vitro reconstitution of synaptic SNAREs without synaptotagmin leads to rapid docking and slow transition to full fusion (Weber et al., 1998), whereas addition of synaptotagmin accelerates fusion, adds Ca2+ sensitivity and allows reduction of SNARE density on liposomes (Tucker et al., 2004).

Release of neurotransmitter might not require complete fusion under all circumstances. Rather, the `kiss-and-run' mode could be favored (Aravanis et al., 2003), depending on the intracellular Ca2+ concentration (Ales et al., 1999), the synaptotagmin subtype (Wang et al., 2003) or the release probability of the respective type of nerve terminal (Gandhi and Stevens, 2003). During kiss-and-run events, neurotransmitter molecules are thought to be channeled through fusion pores (Fig. 3) (Bruns and Jahn, 1995; Borisovska et al., 2005) that extend through the vesicular and plasma membranes. These fusion pores might contain the syntaxin TMS since mutations in it alter neurotransmitter flux and pore conductance (Han et al., 2004). These mutations lie on the same face of a helix, outside the homodimerization motif (Laage et al., 2000). Such mutational analysis provides the first indication that the syntaxin TMS functions in fusion pore formation, but additional experimentation is clearly needed to clarify its precise contribution (Szule and Coorssen, 2004). In the kiss-and-run mode, the vesicles might remain tightly docked or even hemifused (see below) to the plasma membrane (Fig. 3). It is not known whether this stage is an intermediate before full fusion or whether it represents a separate pathway (Söllner, 2004). It will be interesting to see whether the synaptobrevin TMS also contributes to the fusion pore and whether TMS-TMS interactions are involved. Elucidating how the transition from hemifusion to full fusion is controlled and how this relates to kiss-and run will certainly be a major focus of future investigation.

Fig. 3.

Membrane fusion. (A) Putative model of SNARE-mediated membrane fusion. At stage I, Q- and R-SNAREs are separate from each other and the Q-SNARE syntaxin exists in its closed conformation. SNARE complex formation releases the Q-SNARE HABC N-terminal domain and is associated with vesicle docking (stage II). Several SNARE complexes may associate at the fusion site via their TMSs. Docking may result in hemifusion (III), which is followed by formation of lipidic (IV) or proteinaceous (V) fusion pores prior to full fusion (V). Note that proteins and membranes are not drawn to scale. (B) Lipid topology in hemifusion.

Fig. 3.

Membrane fusion. (A) Putative model of SNARE-mediated membrane fusion. At stage I, Q- and R-SNAREs are separate from each other and the Q-SNARE syntaxin exists in its closed conformation. SNARE complex formation releases the Q-SNARE HABC N-terminal domain and is associated with vesicle docking (stage II). Several SNARE complexes may associate at the fusion site via their TMSs. Docking may result in hemifusion (III), which is followed by formation of lipidic (IV) or proteinaceous (V) fusion pores prior to full fusion (V). Note that proteins and membranes are not drawn to scale. (B) Lipid topology in hemifusion.

Mechanism of lipid mixing

SNAREs probably constitute the minimal machinery necessary for membrane fusion (Fig. 3D and Fig. 4). This is underscored by their ability to mediate liposome-liposome fusion in vitro (Schütte et al., 2004; Tucker et al., 2004; Weber et al., 1998; Zhang et al., 2004). According to the original SNARE hypothesis (Rothman, 1994), trans SNARE pairing would pull the membranes together as a prerequisite of complete bilayer fusion. This simple scenario was based upon the following concept: high-curvature bilayers, such as those of secretory vesicles, are more strained than those of low-curvature bilayers that form upon fusion, and thus fusion is expected to be a downhill process. However, a shell of water molecules bound to the lipid head groups repels membranes at short distances and thus prohibits spontaneous fusion (Leikin et al., 1987). Therefore, close membrane apposition mediated by SNARE complex formation was thought to trigger fusion by local dehydration. However, it became evident that complete bilayer merging requires at least two major steps and that mere membrane apposition is unlikely to suffice.

Sequential steps of bilayer mixing

The first step of fusion corresponds to the mixing of the contacting monolayers while the distal monolayers stay intact. This generates the `hemifusion' intermediate (Chernomordik et al., 1995; Chernomordik and Kozlov, 2003). The energy barrier separating non-fused membranes and the hemifusion intermediate is probably relatively small (Cohen and Melikyan, 2004) and may be overcome by the binding enthalpy released upon SNARE complex formation. Viral fusion proteins such as influenza hemagglutinin (HA) may play a similar role. Viral fusogens share basic architectural features with SNAREs, such as their single-span transmembrane topology and oligomerization through formation of membrane-extrinsic coiled-coil domains (Skehel and Wiley, 1998). A global pH-driven conformational change in the ectodomain pushes an amphipathic fusion peptide towards the target bilayer to establish initial contact between both membranes. This is followed by further conformational changes, resulting in bilayer merging. If one considers the viral fusion peptide to be the functional correlate of one SNARE TMS, the fusion mechanism elicited by viral fusogens and SNAREs is likely to share crucial features (Jahn and Südhof, 1999; Söllner, 2004). Whereas SNAREs might initiate fusion by assembly, viral fusogens achieve the same goal by conformational rearrangement (Jahn et al., 2003; Tamm et al., 2003; Söllner, 2004).

The second main step of fusion is the transition of the hemifusion intermediate to complete bilayer merging, and the corresponding energy barrier might significantly exceed that of the first step (Cohen and Melikyan, 2004). For several viral fusion proteins, it has been shown that fusion is arrested at the hemifusion stage if the membrane anchor is altered. For example, influenza HA loses its ability to mediate complete bilayer fusion when a glycosylphosphatidylinositol (GPI) membrane anchor replaces its TMS; by contrast, hemifusion (Kemble et al., 1994; Nüssler et al., 1997) and small, nonenlarging fusion pores (Markosyan et al., 2000) can still be detected. A hemifusion phenotype is also seen upon shortening of the HA TMS by more than 12 residues (Armstrong et al., 2000). Likewise, substituting the TMS of the vesicular stomatitis virus G-protein for a GPI anchor (Odell et al., 1997) or mutating it (Cleverley and Lenard, 1998) attenuates full fusion but still allows hemifusion. Several studies indicate that a hemifusion intermediate is also an authentic stage of SNARE-driven membrane fusion.

First, replacing the TMSs of synaptic SNAREs by lipid anchors (di-oleoyl phosphatidylethanolamine or C15 and C20 isoprenoids) that cannot span the lipid bilayer blocks lipid mixing of liposomes but not SNARE complex formation. Long isoprenoid chains that can span both bilayers (C45 and C55 isoprenoids) restore lipid mixing (McNew et al., 2000b). McNew et al. have therefore proposed that the SNARE complex exerts force on the anchors but these have to be of sufficient length to initiate fusion (McNew et al., 2000b). Because these long lipid anchors render the bilayers too leaky to ascertain complete fusion (McNew et al., 2000b), these results are also compatible with a hemifused state. Second, replacement of the TMSs of exocytotic and vacuolar SNAREs by C16 geranyl-geranyl moieties abolishes content mixing of yeast vacuolar membranes (Grote et al., 2000; Rohde et al., 2003) without affecting trans SNARE complex formation. The isoprenylated SNAREs might elicit hemifusion, since addition of lysolecithin (an agent that induces positive curvature in membranes) to the cells rescues the functional defect (Grote et al., 2000). Third, shortening the TMS of a yeast SNARE protein to approximately half of its original length or using low protein-to-lipid ratios in the fusion assay arrests liposomes at the hemifusion stage (Xu et al., 2005). Fourth, examining vacuole fusion, Reese et al. showed that lipid mixing can occur in the absence of content mixing in the presence of GTP-γS (Reese et al., 2005). Similarly, it has recently been demonstrated that some SNARE-mediated cell-cell fusion events, as well as in vitro liposome fusion induced by neuronal SNAREs, proceed via a hemifusion intermediate (Giraudo et al., 2005; Lu et al., 2005).

Reconstitution of fusion by TMS mimics

As outlined above, several studies indicate that altering the TMS can block fusion by showing loss of content mixing (Grote et al., 2000; Rohde et al., 2003) paralleled by arrest at the hemifusion stage (Reese et al., 2005; Xu et al., 2005; Giraudo et al., 2005; Lu et al., 2005). It is thus conceivable that one role of the TMS is to promote the hemifusion-to-fusion transition. This is supported by in vitro studies in which synthetic peptides representing synaptic SNARE TMSs can drive complete liposome fusion (Langosch et al., 2001b). Moreover, the fusogenic activity of TMS peptides that have mutated sequences decreases with increasing stability of their α-helical conformations. Thus, structural flexibility of the TMSs might support their fusogenicity, and this is consistent with their unusual amino acid compositions [i.e. over-representation of β-sheet-promoting β-branched residues (valine and isoleucine) (Langosch et al., 2001b)]. Peptides representing the TMS of the vesicular stomatitis virus G-protein yield similar results. There, mutations that diminish fusogenicity of the full-length protein in vivo (Cleverley and Lenard, 1998) also diminish fusogenicity of the TMS peptide in vitro and stabilize its helical structure (Dennison et al., 2002; Langosch et al., 2001a).

In the absence of the ectodomains, fusion by TMS peptides is probably initiated by random collisions of the liposomes and is strongly enhanced by Ca2+-mediated aggregation (Langosch et al., 2001a, Langosch et al., 2001b). The suggestion that structurally flexible TMSs might contribute to lipid mixing is supported by the results of recent studies with de novo designed fusogenic TMS peptides. Depending on the ratio of helix-promoting leucine, sheet-promoting valine and helix-destabilizing proline and glycine residues, these peptides also display fusogenicities that are paralleled by various degrees of conformational flexibility (Hofmann et al., 2004). How flexibility of the TMS could translate into lipid mixing is presently unclear.

Lipid-protein interactions

Although little is known about the role of lipids in natural fusion reactions, the interdependence of lipids and fusion proteins has been addressed in some model systems. For example, liposome fusion in vitro is promoted by cone-shaped lipids, which are thought to stabilize the hemifusion intermediate, since they favor negative membrane curvature (Chernomordik et al., 1995; Chernomordik and Kozlov, 2003). Furthermore, lipids have recently been shown to have a role in yeast vacuole fusion. Fusion of yeast vacuoles starts at the vertices of apposed bilayers, which are enriched in the Q-SNARE Vam3p and regulatory proteins including Vac8p, the GTPase Ypt7p, the HOPS/Vps-C effector complex and protein phosphatase 1 (Wang et al., 2002). In addition, several regulatory lipids, including ergosterol, diacylglycerol and 3- and 4-phosphoinositides, accumulate at the vertices with SNAREs and other fusion factors in a mutually interdependent manner (Fratti et al., 2004). The phosphoinositides appear to form acceptor sites for the soluble SNARE Vam7p and actin. Ergosterol regulates Sec17p release and diacylglycerol might accelerate fusion owing to its cone shape (Jun et al., 2004). Palmitoylated proteins such as Vac8p might facilitate fusion by intercalating into the lipids at the fusion site (Veit et al., 2001; Wang et al., 2001; Wang et al., 2002). Likewise, transmembrane proteins such as the V0 subunit of the ATPase (Peters et al., 2001) and the associated Vtc proteins (Müller et al., 2002) might support the organization of lipids and SNAREs at the fusion site. Indeed, lipid and content mixing do not occur in vacuoles lacking the V0 subunit Vph1p or Vac8p (Reese et al., 2005).

Clustering of SNAREs has also been observed in other model systems (Lang et al., 2001; Salaun et al., 2004), which suggests that lipid-SNARE communication is a ubiquitous phenomenon in vivo. Interestingly, SNAREs prefer the liquid-disordered phase and not the `raft' phase when reconstituted into liposomes (Bacia et al., 2004). This indicates that the interaction of lipids and SNARE TMSs can support lateral sorting in membranes. It will be important to analyze the functional consequences of this observation for membrane fusion.

SNARE recycling

In contrast to viral fusion proteins, SNAREs can be reused after fusion. This requires the action of NSF and α/γ-SNAPs (Söllner et al., 1993), which dissociate SNARE complexes either prior to fusion (as in the case of in vitro organelle fusion and ER-Golgi transport reactions) (Barlowe, 1997; Mayer et al., 1996) or after fusion (as in regulated secretion) (Jahn et al., 2003). α-SNAP might serve as a folding sensor that binds to assembled SNARE complexes, mediates recruitment of NSF and supports NSF-directed disassembly. Recent studies have highlighted potential mechanisms for regulation of NSF, including its nitrosylation and phosphorylation (Huynh et al., 2004; Matsushita et al., 2003). How α-SNAP is able to bind to different SNARE subtypes is still mysterious. Interestingly, blocking the disassembly of SNAREs also interferes with vesicle budding, which suggests that the fusion and fission reactions are coupled (Littleton et al., 2001; Deak et al., 2004). This concept of SNARE function and SNARE sorting has recently been addressed in a mathematical model of the organization of the endomembrane system (Heinrich and Rapoport, 2005). The authors show that organelle identity can be explained by the specificity of the interaction of SNAREs on vesicle and target membranes and the affinity of a SNARE for a specific vesicle coat. Thus, only certain SNAREs should be packaged into vesicles and therefore be available for fusion. In turn, the nonhomogeneous distribution of SNAREs results in coat-specific vesicle flux and thus in nonidentical compartments.

Research on SNAREs will need to address several crucial questions that are only touched on here. Dissociation of SNARE complexes by NSF and α-SNAP is well established, but has not been resolved on a mechanistic level. The interplay of SNAREs and tethers is likely to be regulated by a variety of factors, including kinases, phosphatases and other modifying enzymes, but the details are unknown. We also know little about what determines the specificity of SNARE-mediated fusion, although it is clear that distinct combinations of SNAREs are required in vitro to fuse liposomes. It is also unclear whether SNARE assembly can be stalled at intermediate stages and, if so, how stalling is mediated or reversed. The interplay of synaptotagmin and SM proteins with SNAREs is established, but the temporal sequence of interactions at the fusion site is still mysterious. Similarly, the roles of SNARE TMSs in assembly and lipid mixing are apparent but their place within the fusion mechanism has yet to be defined. Proteins and lipids might form fusion-competent zones on membranes and support recruitment of SNAREs. Only a few studies have addressed the nature of such zones, and this area also requires further research.

We thank Francis Barr, Dieter Bruns, Ed Chapman, Andreas Mayer, Karolina Peplowska and William Wickner for helpful comments on the manuscript, and apologize to the many researchers whose work could not be cited owing to space constraints. Preparation of this article was supported by the Deutsche Forschungsgemeinschaft (grant LA699/8-1,2 to D.L. and C.U.).

Ales, E., Tabares, L., Poyato, J. M., Valero, V., Lindau, M. and Alvarez de Toledo, G. (
1999
). High calcium concentrations shift the mode of exocytosis to the kiss-and-run mechanism.
Nat. Cell Biol.
1
,
40
-44.
An, S. J. and Almers, W. (
2004
). Tracking SNARE complex formation in live endocrine cells.
Science
306
,
1042
-1046.
Antonin, W., Dulubova, I., Arac, D., Pabst, S., Plitzner, J., Rizo, J. and Jahn, R. (
2002a
). The N-terminal domains of syntaxin 7 and vti1b form three-helix bundles that differ in their ability to regulate SNARE complex assembly.
J. Biol. Chem.
277
,
36449
-36456.
Antonin, W., Fasshauer, D., Becker, S., Jahn, R. and Schneider, T. R. (
2002b
). Crystal structure of the endosomal SNARE complex reveals common structural principles of all SNAREs.
Nat. Struct. Biol.
9
,
107
-111.
Aravanis, A. M., Pyle, J. L. and Tsien, R. W. (
2003
). Single synaptic vesicles fusing transiently and successively without loss of identity.
Nature
423
,
643
-647.
Armstrong, R. T., Kushnir, A. S. and White, J. M. (
2000
). The transmembrane domain of influenza hemagglutinin exhibits a stringent length requirement to support the hemifusion to fusion transition.
J. Cell Biol.
151
,
425
-438.
Bacia, K., Schütte, C. G., Kahya, N., Jahn, R. and Schwille, P. (
2004
). SNAREs prefer liquid-disordered over `raft' (liquid-ordered) domains when reconstituted into giant unilamellar vesicles.
J. Biol. Chem.
279
,
37951
-37955.
Bai, J., Wang, C. T., Richards, D. A., Jackson, M. B. and Chapman, E. R. (
2004
). Fusion pore dynamics are regulated by synaptotagmin*t-SNARE interactions.
Neuron
41
,
929
-942.
Barlowe, C. (
1997
). Coupled ER to Golgi transport reconstituted with purified cytosolic proteins.
J. Cell Biol.
139
,
1097
-1108.
Bock, J. B., Matern, H. T., Peden, A. A. and Scheller, R. H. (
2001
). A genomic perspective on membrane compartment organization.
Nature
409
,
839
-841.
Boeddinghaus, C., Merz, A. J., Laage, R. and Ungermann, C. (
2002
). A cycle of Vam7p release from and PtdIns 3-P-dependent rebinding to the yeast vacuole is required for homotypic vacuole fusion.
J. Cell Biol.
157
,
79
-90.
Bonifacino, J. S. and Glick, B. S. (
2004
). The mechanisms of vesicle budding and fusion.
Cell
116
,
153
-166.
Borisovska, M., Zhao, Y., Tsytsyura, Y., Glyvuk, N., Takamori, S., Matti, U., Rettig, J., Südhof, T. and Bruns, D. (
2005
). v-SNAREs control exocytosis of vesicles from priming to fusion.
EMBO J.
24
,
2114
-2126.
Bracher, A. and Weissenhorn, W. (
2002
). Structural basis for the Golgi membrane recruitment of Sly1p by Sed5p.
EMBO J.
21
,
6114
-6124.
Bruns, D. and Jahn, R. (
1995
). Real-time measurement of transmitter release from single synaptic vesicles.
Nature
377
,
62
-65.
Burgoyne, R. D. and Clague, M. J. (
2003
). Calcium and calmodulin in membrane fusion.
Biochem. Biophys. Acta
1641
,
137
-143.
Burri, L. and Lithgow, T. (
2004
). A complete set of SNAREs in yeast.
Traffic
5
,
45
-52.
Calakos, N. and Scheller, R. H. (
1994
). Vesicle-associated membrane protein and synaptophysin are associated on the synaptic vesicle.
J. Biol. Chem.
269
,
24534
-24537.
Carr, C. M., Grote, E., Munson, M., Hughson, F. M. and Novick, P. J. (
1999
). Sec1p binds to SNARE complexes and concentrates at sites of secretion.
J. Cell Biol.
146
,
333
-344.
Cheever, M. L., Sato, T. K., de Beer, T., Kutateladze, T. G., Emr, S. D. and Overduin, M. (
2001
). Phox domain interaction with PtdIns(3)P targets the Vam7 t-SNARE to vacuole membranes.
Nat. Cell Biol.
3
,
613
-618.
Chen, X., Tomchick, D. R., Kovrigin, E., Arac, D., Machius, M., Südhof, T. C. and Rizo, J. (
2002
). Three-dimensional structure of the complexin/SNARE complex.
Neuron
33
,
397
-409.
Chen, Y. A. and Scheller, R. H. (
2001
). SNARE-mediated membrane fusion.
Nat. Rev. Mol. Cell Biol.
2
,
98
-106.
Chen, Y. A., Scales, S. J., Patel, S. M., Doung, Y. C. and Scheller, R. H. (
1999
). SNARE complex formation is triggered by Ca2+ and drives membrane fusion.
Cell
97
,
165
-174.
Chernomordik, L. V. and Kozlov, M. M. (
2003
). Protein-lipid interplay in fusion and fission of biological membranes.
Annu. Rev. Biochem.
72
,
175
-207.
Chernomordik, L., Kozlov, M. M. and Zimmerberg, J. (
1995
). Lipids in biological membrane fusion.
J. Membr. Biol.
146
,
1
-14.
Cleverley, D. Z. and Lenard, J. (
1998
). The transmembrane domain in viral fusion: essential role for a conserved glycine residue in vesicular stomatitis virus G protein.
Proc. Natl. Acad. Sci. USA
95
,
3425
-3430.
Cohen, F. S. and Melikyan, G. B. (
2004
). The energetics of membrane fusion from binding, through hemifusion, pore formation, and pore enlargement.
J. Membr. Biol.
199
,
1
-14.
Collins, B. M., McCoy, A. J., Kent, H. M., Evans, P. R. and Owen, D. J. (
2002
). Molecular architecture and functional model of the endocytic AP2 complex.
Cell
109
,
523
-535.
Collins, K. M., Thorngren, N. L., Fratti, R. A. and Wickner, W. T. (
2005
). Sec17p and HOPS, in distinct SNARE complexes, mediate SNARE complex disruption or assembly for fusion.
EMBO J.
24
,
1775
-1786.
Deak, F., Schoch, S., Liu, X., Südhof, T. C. and Kavalali, E. T. (
2004
). Synaptobrevin is essential for fast synaptic-vesicle endocytosis.
Nat. Cell. Biol.
6
,
1102
-1108.
Dennison, S. M., Greenfield, N., Lenard, J. and Lentz, B. R. (
2002
). VSV transmembrane domain (TMD) peptide promotes PEG-mediated fusion of liposomes in a conformationally sensitive fashion.
Biochemistry
41
,
14925
-14934.
Dietrich, L. E., Boeddinghaus, C., LaGrassa, T. J. and Ungermann, C. (
2003
). Control of eukaryotic membrane fusion by N-terminal domains of SNARE proteins.
Biochim. Biophys. Acta
1641
,
111
-119.
Dietrich, L. E., LaGrassa, T. J., Rohde, J., Cristodero, M., Meiringer, C. T. and Ungermann, C. (
2005a
). ATP-independent control of Vac8 Palmitoylation by a SNARE subcomplex on yeast vacuoles.
J. Biol. Chem.
280
,
15348
-15355.
Dietrich, L. E., Peplowska, K., LaGrassa, T. J., Hou, H., Rohde, J. and Ungermann, C. (
2005b
). The SNARE Ykt6 is released from yeast vacuoles during an early stage of fusion.
EMBO Rep.
6
,
245
-250.
Dulubova, I., Yamaguchi, T., Wang, Y., Südhof, T. C. and Rizo, J. (
2001
). Vam3p structure reveals conserved and divergent properties of syntaxins.
Nat. Struct. Biol.
8
,
258
-264.
Dulubova, I., Yamaguchi, T., Gao, Y., Min, S. W., Huryeva, I., Südhof, T. C. and Rizo, J. (
2002
). How Tlg2p/syntaxin 16 `snares' Vps45.
EMBO J.
21
,
3620
-3631.
Dulubova, I., Yamaguchi, T., Arac, D., Li, H., Huryeva, I., Min, S. W., Rizo, J. and Südhof, T. C. (
2003
). Convergence and divergence in the mechanism of SNARE binding by Sec1/Munc18-like proteins.
Proc. Natl. Acad. Sci. USA
100
,
32
-37.
Edelmann, L., Hanson, P. I., Chapman, E. R. and Jahn, R. (
1995
). Synaptobrevin binding to synaptophysin: a potential mechanism for controlling the exocytotic fusion machine.
EMBO J.
14
,
224
-231.
Fasshauer, D. (
2003
). Structural insights into the SNARE mechanism.
Biochem. Biophys. Acta
1641
,
87
-97.
Fasshauer, D. and Margittai, M. (
2004
). A transient N-terminal interaction of SNAP-25 and syntaxin nucleates SNARE assembly.
J. Biol. Chem.
279
,
7613
-7621.
Fasshauer, D., Sutton, R. B., Brunger, A. T. and Jahn, R. (
1998
). Conserved structural features of the synaptic fusion complex: SNARE proteins reclassified as Q- and R-SNAREs.
Proc. Natl. Acad. Sci. USA
95
,
15781
-15786.
Fasshauer, D., Antonin, W., Margittai, M., Pabst, S. and Jahn, R. (
1999
). Mixed and non-cognate SNARE complexes. Characterization of assembly and biophysical properties.
J. Biol. Chem.
274
,
15440
-15446.
Fernandez, I., Ubach, J., Dulubova, I., Zhang, X., Südhof, T. C. and Rizo, J. (
1998
). Three-dimensional structure of an evolutionarily conserved N-terminal domain of syntaxin 1A.
Cell
94
,
841
-849.
Fernandez-Chacon, R., Konigstorfer, A., Gerber, S. H., Garcia, J., Matos, M. F., Stevens, C. F., Brose, N., Rizo, J., Rosenmund, C. and Südhof, T. C. (
2001
). Synaptotagmin I functions as a calcium regulator of release probability.
Nature
410
,
41
-49.
Filippini, F., Rossi, V., Galli, T., Budillon, A., D'Urso, M. and D'Esposito, M. (
2001
). Longins: a new evolutionary conserved VAMP family sharing a novel SNARE domain.
Trends Biochem. Sci.
26
,
407
-409.
Fleming, K. G. and Engelman, D. M. (
2001
). Computation and mutagenesis suggest a right-handed structure for the synaptobrevin transmembrane dimer.
Proteins
45
,
313
-317.
Fratti, R. A., Jun, Y., Merz, A. J., Margolis, N. and Wickner, W. (
2004
). Interdependent assembly of specific regulatory lipids and membrane fusion proteins into the vertex ring domain of docked vacuoles.
J. Cell Biol.
167
,
1087
-1098.
Fujita, Y., Shirataki, H., Sakisaka, T., Asakura, T., Ohya, T., Kotani, H., Yokoyama, S., Nishioka, H., Matsuura, Y., Mizoguchi, A. et al. (
1998
). Tomosyn: a syntaxin-1-binding protein that forms a novel complex in the neurotransmitter release process.
Neuron
20
,
905
-915.
Fukasawa, M., Varlamov, O., Eng, W. S., Söllner, T. H. and Rothman, J. E. (
2004
). Localization and activity of the SNARE Ykt6 determined by its regulatory domain and palmitoylation.
Proc. Natl. Acad. Sci. USA
101
,
4815
-4820.
Fukuda, R., McNew, J. A., Weber, T., Parlati, F., Engel, T., Nickel, W., Rothman, J. E. and Söllner, T. H. (
2000
). Functional architecture of an intracellular membrane t-SNARE.
Nature
407
,
198
-202.
Gandhi, S. P. and Stevens, C. F. (
2003
). Three modes of synaptic vesicular recycling revealed by single-vesicle imaging.
Nature
423
,
607
-613.
Gillingham, A. K. and Munro, S. (
2003
). Long coiled-coil proteins and membrane traffic.
Biochem. Biophys. Acta
1641
,
71
-85.
Giraudo, C. G., Hu, C., You, D., Slovic, A. M., Mosharov, E. V., Sulzer, D., Melia, T. J. and Rothman, J. E. (
2005
). SNAREs can promote complete fusion and hemifusion as alternative outcomes.
J. Cell Biol.
170
,
249
-260.
Gonzalez, L. C., Jr, Weis, W. I. and Scheller, R. H. (
2001
). A novel snare n-terminal domain revealed by the crystal structure of sec22b.
J. Biol. Chem.
276
,
24203
-24211.
Götte, M. and Gallwitz, D. (
1997
). High expression of the yeast syntaxin-related Vam3 protein suppresses the protein transport defects of a pep12 null mutant.
FEBS Lett.
411
,
48
-52.
Graf, C. T., Riedel, D., Schmitt, H. D. and Jahn, R. (
2005
). Identification of functionally interacting SNAREs by using complementary substitutions in the conserved `0' layer.
Mol. Biol. Cell
16
,
2263
-2274.
Grote, E., Baba, M., Ohsumi, Y. and Novick, P. J. (
2000
). Geranylgeranylated SNAREs are dominant inhibitors of membrane fusion.
J. Cell Biol.
151
,
453
-466.
Han, X., Wang, C. T., Bai, J., Chapman, E. R. and Jackson, M. B. (
2004
). Transmembrane segments of syntaxin line the fusion pore of Ca2+-triggered exocytosis.
Science
304
,
289
-292.
Hasegawa, H., Yang, Z., Oltedal, L., Davanger, S. and Hay, J. C. (
2004
). Intramolecular protein-protein and protein-lipid interactions control the conformation and subcellular targeting of neuronal Ykt6.
J. Cell Sci.
117
,
4495
-4508.
Heinrich, R. and Rapoport, T. A. (
2005
). Generation of nonidentical compartments in vesicular transport systems.
J. Cell Biol.
168
,
271
-280.
Hess, D. T., Slater, T. M., Wilson, M. C. and Skene, J. H. (
1992
). The 25 kDa synaptosomal-associated protein SNAP-25 is the major methionine-rich polypeptide in rapid axonal transport and a major substrate for palmitoylation in adult CNS.
J. Neurosci.
12
,
4634
-4641.
Hofmann, M. W., Weise, K., Ollesch, J., Agrawal, P., Stalz, H., Stelzer, W., Hulsbergen, F., de Groot, H., Gerwert, K., Reed, J. et al. (
2004
). De novo design of conformationally flexible transmembrane peptides driving membrane fusion.
Proc. Natl. Acad. Sci. USA
101
,
14776
-14781.
Hu, K., Carroll, J., Fedorovich, S., Rickman, C., Sukhodub, A. and Davletov, B. (
2002
). Vesicular restriction of synaptobrevin suggests a role for calcium in membrane fusion.
Nature
415
,
646
-650.
Huynh, H., Bottini, N., Williams, S., Cherepanov, V., Musumeci, L., Saito, K., Bruckner, S., Vachon, E., Wang, X., Kruger, J. et al. (
2004
). Control of vesicle fusion by a tyrosine phosphatase.
Nat. Cell Biol.
6
,
831
-839.
Jahn, R. and Südhof, T. C. (
1999
). Membrane fusion and exocytosis.
Annu. Rev. Biochem.
68
,
863
-911.
Jahn, R. and Grubmüller, H. (
2002
). Membrane fusion.
Curr. Opin. Cell Biol.
14
,
488
-495.
Jahn, R., Lang, T. and Südhof, T. C. (
2003
). Membrane fusion.
Cell
112
,
519
-533.
Jang, S. B., Kim, Y. G., Cho, Y. S., Suh, P. G., Kim, K. H. and Oh, B. H. (
2002
). Crystal structure of SEDL and its implications for a genetic disease spondyloepiphyseal dysplasia tarda.
J. Biol. Chem.
277
,
49863
-49869.
Jun, Y., Fratti, R. A. and Wickner, W. (
2004
). Diacylglycerol and its formation by phospholipase C regulate Rab- and SNARE-dependent yeast vacuole fusion.
J. Biol. Chem.
279
,
53186
-53195.
Katz, L. and Brennwald, P. (
2000
). Testing the 3Q:1R `Rule': mutational analysis of the ionic `Zero' layer in the yeast exocytic SNARE complex reveals no requirement for arginine.
Mol. Biol. Cell
11
,
3849
-3858.
Kemble, G. W., Danieli, T. and White, J. M. (
1994
). Lipid-anchored influenza hemagglutinin promotes hemifusion, not complete fusion.
Cell
76
,
383
-391.
Kim, C. S., Kweon, D. H. and Shin, Y. K. (
2002
). Membrane topologies of neuronal SNARE folding intermediates.
Biochemistry
41
,
10928
-10933.
Knecht, V. and Grubmüller, H. (
2003
). Mechanical coupling via the membrane fusion SNARE protein syntaxin 1A: a molecular dynamics study.
Biophys. J.
84
,
1527
-1547.
Kretzschmar, S., Volknandt, W. and Zimmermann, H. (
1996
). Colocalization on the same synaptic vesicles of syntaxin and SNAP-25 with synaptic vesicle proteins: a re-evaluation of functional models required?
Neurosci. Res.
26
,
141
-148.
Kurzbauer, R., Teis, D., de Araujo, M. E., Maurer-Stroh, S., Eisenhaber, F., Bourenkov, G. P., Bartunik, H. D., Hekman, M., Rapp, U. R., Huber, L. A. et al. (
2004
). Crystal structure of the p14/MP1 scaffolding complex: how a twin couple attaches mitogen-activated protein kinase signaling to late endosomes.
Proc. Natl. Acad. Sci. USA
101
,
10984
-10989.
Kweon, D. H., Kim, C. S. and Shin, Y. K. (
2003
). Regulation of neuronal SNARE assembly by the membrane.
Nat. Struct. Biol.
10
,
440
-447.
Laage, R. and Langosch, D. (
1997
). Dimerization of the synaptic vesicle protein synaptobrevin (vesicle-associated membrane protein) II depends on specific residues within the transmembrane segment.
Eur. J. Biochem.
249
,
540
-546.
Laage, R., Rohde, J., Brosig, B. and Langosch, D. (
2000
). A conserved membrane-spanning amino acid motif drives homomeric and supports heteromeric assembly of presynaptic SNARE proteins.
J. Biol. Chem.
275
,
17481
-17487.
Lang, T., Bruns, D., Wenzel, D., Riedel, D., Holroyd, P., Thiele, C. and Jahn, R. (
2001
). SNAREs are concentrated in cholesterol-dependent clusters that define docking and fusion sites for exocytosis.
EMBO J.
20
,
2202
-2213.
Langosch, D., Brosig, B. and Pipkorn, R. (
2001a
). Peptide mimics of the vesicular stomatitis virus G-protein transmembrane segment drive membrane fusion in vitro.
J. Biol. Chem.
276
,
32016
-32021.
Langosch, D., Crane, J. M., Brosig, B., Hellwig, A., Tamm, L. K. and Reed, J. (
2001b
). Peptide mimics of snare transmembrane segments drive membrane fusion depending on their conformational plasticity.
J. Mol. Biol.
311
,
709
-721.
Leikin, S. L., Kozlov, M. M., Chernomordik, L. V., Markin, V. S. and Chizmadzhev, Y. A. (
1987
). Membrane fusion: overcoming of the hydration barrier and local restructuring.
J. Theor. Biol.
129
,
411
-425.
Littleton, J. T., Barnard, R. J., Titus, S. A., Slind, J., Chapman, E. R. and Ganetzky, B. (
2001
). SNARE-complex disassembly by NSF follows synaptic-vesicle fusion.
Proc. Natl. Acad. Sci. USA
98
,
12233
-12238.
Liu, Y. and Barlowe, C. (
2002
). Analysis of Sec22p in endoplasmic reticulum/Golgi transport reveals cellular redundancy in SNARE protein function.
Mol. Biol. Cell
13
,
3314
-3324.
Lu, X., Zhang, F., McNew, J. A. and Shin, Y. K. (
2005
). Membrane fusion induced by neuronal SNAREs transits through hemifusion.
J. Biol. Chem.
(in press).
Lunin, V. V., Munger, C., Wagner, J., Ye, Z., Cygler, M. and Sacher, M. (
2004
). The structure of the MAPK scaffold, MP1, bound to its partner, p14. A complex with a critical role in endosomal map kinase signaling.
J. Biol. Chem.
279
,
23422
-23430.
Margittai, M., Otto, H. and Jahn, R. (
1999
). A stable interaction between syntaxin 1a and synaptobrevin 2 mediated by their transmembrane domains.
FEBS Lett.
446
,
40
-44.
Margittai, M., Fasshauer, D., Jahn, R. and Langen, R. (
2003
). The Habc domain and the SNARE core complex are connected by a highly flexible linker.
Biochemistry
42
,
4009
-4014.
Markosyan, R. M., Cohen, F. S. and Melikyan, G. B. (
2000
). The lipid-anchored ectodomain of influenza virus hemagglutinin (GPI-HA) is capable of inducing nonenlarging fusion pores.
Mol. Biol. Cell
11
,
1143
-1152.
Matsushita, K., Morrell, C. N., Cambien, B., Yang, S. X., Yamakuchi, M., Bao, C., Hara, M. R., Quick, R. A., Cao, W., O'Rourke, B. et al. (
2003
). Nitric oxide regulates exocytosis by S-nitrosylation of N-ethylmaleimide-sensitive factor.
Cell
115
,
139
-150.
Mayer, A., Wickner, W. and Haas, A. (
1996
). Sec18p (NSF)-driven release of Sec17p (alpha-SNAP) can precede docking and fusion of yeast vacuoles.
Cell
85
,
83
-94.
McBride, H. M., Rybin, V., Murphy, C., Giner, A., Teasdale, R. and Zerial, M. (
1999
). Oligomeric complexes link Rab5 effectors with NSF and drive membrane fusion via interactions between EEA1 and syntaxin 13.
Cell
98
,
377
-386.
McMahon, H. T., Missler, M., Li, C. and Südhof, T. C. (
1995
). Complexins: cytosolic proteins that regulate SNAP receptor function.
Cell
83
,
111
-119.
McNew, J. A., Weber, T., Engelman, D. M., Söllner, T. H. and Rothman, J. E. (
1999
). The length of the flexible SNAREpin juxtamembrane region is a critical determinant of SNARE-dependent fusion.
Mol. Cell
4
,
415
-421.
McNew, J. A., Parlati, F., Fukuda, R., Johnston, R. J., Paz, K., Paumet, F., Söllner, T. H. and Rothman, J. E. (
2000a
). Compartmental specificity of cellular membrane fusion encoded in SNARE proteins.
Nature
407
,
153
-159.
McNew, J. A., Weber, T., Parlati, F., Johnston, R. J., Melia, T. J., Söllner, T. H. and Rothman, J. E. (
2000b
). Close is not enough: SNARE-dependent membrane fusion requires an active mechanism that transduces force to membrane anchors.
J. Cell Biol.
150
,
105
-117.
Melia, T. J., Weber, T., McNew, J. A., Fisher, L. E., Johnston, R. J., Parlati, F., Mahal, L. K., Söllner, T. H. and Rothman, J. E. (
2002
). Regulation of membrane fusion by the membrane-proximal coil of the t-SNARE during zippering of SNAREpins.
J. Cell Biol.
158
,
929
-940.
Mellman, I. and Warren, G. (
2000
). The road taken: past and future foundations of membrane traffic.
Cell
100
,
99
-112.
Müller, O., Bayer, M. J., Peters, C., Andersen, J. S., Mann, M. and Mayer, A. (
2002
). The Vtc proteins in vacuole fusion: coupling NSF activity to V(0) trans-complex formation.
EMBO J.
21
,
259
-269.
Nichols, B. J., Ungermann, C., Pelham, H. R., Wickner, W. T. and Haas, A. (
1997
). Homotypic vacuolar fusion mediated by t- and v-SNAREs.
Nature
387
,
199
-202.
Nicholson, K. L., Munson, M., Miller, R. B., Filip, T. J., Fairman, R. and Hughson, F. M. (
1998
). Regulation of SNARE complex assembly by an N-terminal domain of the t-SNARE Sso1p.
Nat. Struct. Biol.
5
,
793
-802.
Nüssler, F., Clague, M. J. and Herrmann, A. (
1997
). Meta-stability of the hemifusion intermediate induced by glycosylphosphatidylinositol-anchored influenza hemagglutinin.
Biophys. J.
73
,
2280
-2291.
Odell, D., Wanas, E., Yan, J. and Ghosh, H. P. (
1997
). Influence of membrane anchoring and cytoplasmic domains on the fusogenic activity of vesicular stomatitis virus glycoprotein G.
J. Virol.
71
,
7996
-8000.
Pabst, S., Margittai, M., Vainius, D., Langen, R., Jahn, R. and Fasshauer, D. (
2002
). Rapid and selective binding to the synaptic SNARE complex suggests a modulatory role of complexins in neuroexocytosis.
J. Biol. Chem.
277
,
7838
-7848.
Parlati, F., Weber, T., McNew, J. A., Westermann, B., Söllner, T. H. and Rothman, J. E. (
1999
). Rapid and efficient fusion of phospholipid vesicles by the alpha-helical core of a SNARE complex in the absence of an N-terminal regulatory domain.
Proc. Natl. Acad. Sci. USA
96
,
12565
-12570.
Peng, R. and Gallwitz, D. (
2004
). Multiple SNARE interactions of an SM protein: Sed5p/Sly1p binding is dispensable for transport.
EMBO J.
23
,
3939
-3949.
Pennuto, M., Dunlap, D., Contestabile, A., Benfenati, F. and Valtorta, F. (
2002
). Fluorescence resonance energy transfer detection of synaptophysin I and vesicle-associated membrane protein 2 interactions during exocytosis from single live synapses.
Mol. Biol. Cell
13
,
2706
-2717.
Peters, C., Bayer, M. J., Buhler, S., Andersen, J. S., Mann, M. and Mayer, A. (
2001
). Trans-complex formation by proteolipid channels in the terminal phase of membrane fusion.
Nature
409
,
581
-588.
Pobbati, A. V., Razeto, A., Boddener, M., Becker, S. and Fasshauer, D. (
2004
). Structural basis for the inhibitory role of tomosyn in exocytosis.
J. Biol. Chem.
279
,
47192
-47200.
Reese, C., Heise, F. and Mayer, A. (
2005
). Trans-SNARE pairing can precede a hemifusion intermediate in intracellular membrane fusion.
Nature
(in press).
Rizo, J. and Südhof, T. C. (
2002
). Snares and munc18 in synaptic vesicle fusion.
Nat. Rev. Neurosci.
3
,
641
-653.
Rohde, J., Dietrich, L., Langosch, D. and Ungermann, C. (
2003
). The transmembrane domain of Vam3 affects the composition of cis- and trans-SNARE complexes to promote homotypic vacuole fusion.
J. Biol. Chem.
278
,
1656
-1662.
Rossi, V., Banfield, D. K., Vacca, M., Dietrich, L. E., Ungermann, C., D'Esposito, M., Galli, T. and Filippini, F. (
2004
). Longins and their longin domains: regulated SNAREs and multifunctional SNARE regulators.
Trends Biochem. Sci.
29
,
682
-688.
Rothman, J. E. (
1994
). Mechanisms of intracellular protein transport.
Nature
372
,
55
-63.
Roy, R., Laage, R. and Langosch, D. (
2004
). Synaptobrevin transmembrane domain dimerization - revisited.
Biochemistry
43
,
4964
-4970.
Salaun, C., James, D. J. and Chamberlain, L. H. (
2004
). Lipid rafts and the regulation of exocytosis.
Traffic
5
,
255
-264.
Scales, S. J., Chen, Y. A., Yoo, B. Y., Patel, S. M., Doung, Y. C. and Scheller, R. H. (
2000
). SNAREs contribute to the specificity of membrane fusion.
Neuron
26
,
457
-464.
Scales, S. J., Hesser, B. A., Masuda, E. S. and Scheller, R. H. (
2002
). Amisyn, a novel syntaxin-binding protein that may regulate SNARE complex assembly.
J. Biol. Chem.
277
,
28271
-28279.
Schütte, C. G., Hatsuzawa, K., Margittai, M., Stein, A., Riedel, D., Kuster, P., Konig, M., Seidel, C. and Jahn, R. (
2004
). Determinants of liposome fusion mediated by synaptic SNARE proteins.
Proc. Natl. Acad. Sci. USA
101
,
2858
-2863.
Schwartz, T. and Blobel, G. (
2003
). Structural basis for the function of the beta subunit of the eukaryotic signal recognition particle receptor.
Cell
112
,
793
-803.
Scott, B. L., Van Komen, J. S., Irshad, H., Liu, S., Wilson, K. A. and McNew, J. A. (
2004
). Sec1p directly stimulates SNARE-mediated membrane fusion in vitro.
J. Cell Biol.
167
,
75
-85.
Shorter, J. and Warren, G. (
2002
). Golgi architecture and inheritance.
Annu. Rev. Cell Dev. Biol.
18
,
379
-420.
Shorter, J., Beard, M. B., Seemann, J., Dirac-Svejstrup, A. B. and Warren, G. (
2002
). Sequential tethering of Golgins and catalysis of SNAREpin assembly by the vesicle-tethering protein p115.
J. Cell Biol.
157
,
45
-62.
Skehel, J. J. and Wiley, D. C. (
1998
). Coiled coils in both intracellular vesicle and viral membrane fusion.
Cell
95
,
871
-874.
Söllner, T. H. (
2004
). Intracellular and viral membrane fusion: a uniting mechanism.
Curr. Opin. Cell Biol.
16
,
429
-435.
Südhof, T. C. (
2002
). Synaptotagmins: why so many?
J. Biol. Chem.
277
,
7629
-7632.
Südhof, T. C. (
2004
). The synaptic vesicle cycle.
Annu. Rev. Neurosci.
27
,
509
-547.
Sutton, R. B., Fasshauer, D., Jahn, R. and Brünger, A. T. (
1998
). Crystal structure of a SNARE complex involved in synaptic exocytosis at 2.4 Å resolution.
Nature
395
,
347
-353.
Szule, J. A. and Coorssen, J. R. (
2004
). Comment on `Transmembrane segments of syntaxin line the fusion pore of Ca2+-triggered exocytosis'.
Science
306
,
813
.
Tamm, L. K., Crane, J. and Kiessling, V. (
2003
). Membrane fusion: a structural perspective on the interplay of lipids and proteins.
Curr. Opin. Struct. Biol.
13
,
453
-466.
Tochio, H., Tsui, M. M., Banfield, D. K. and Zhang, M. (
2001
). An autoinhibitory mechanism for nonsyntaxin SNARE proteins revealed by the structure of Ykt6p.
Science
293
,
698
-702.
Tsui, M. M. and Banfield, D. K. (
2000
). Yeast Golgi SNARE interactions are promiscuous.
J. Cell Sci.
113
,
145
-152.
Tucker, W. C., Weber, T. and Chapman, E. R. (
2004
). Reconstitution of Ca2+-regulated membrane fusion by synaptotagmin and SNAREs.
Science
304
,
435
-438.
Ungermann, C., Sato, K. and Wickner, W. (
1998
). Defining the functions of trans-SNARE pairs.
Nature
396
,
543
-548.
Veit, M., Laage, R., Dietrich, L., Wang, L. and Ungermann, C. (
2001
). Vac8p release from the SNARE complex and its palmitoylation are coupled and essential for vacuole fusion.
EMBO J.
20
,
3145
-3155.
Vogel, K. and Roche, P. A. (
1999
). SNAP-23 and SNAP-25 are palmitoylated in vivo.
Biochem. Biophys. Res. Commun.
258
,
407
-410.
Wang, C. T., Lu, J. C., Bai, J., Chang, P. Y., Martin, T. F., Chapman, E. R. and Jackson, M. B. (
2003
). Different domains of synaptotagmin control the choice between kiss-and-run and full fusion.
Nature
424
,
943
-947.
Wang, L., Seeley, E. S., Wickner, W. and Merz, A. J. (
2002
). Vacuole fusion at a ring of vertex docking sites leaves membrane fragments within the organelle.
Cell
108
,
357
-369.
Wang, Y. X., Kauffman, E. J., Duex, J. E. and Weisman, L. S. (
2001
). Fusion of docked membranes requires the armadillo repeat protein Vac8p.
J. Biol. Chem.
276
,
35133
-35140.
Washbourne, P., Schiavo, G. and Montecucco, C. (
1995
). Vesicle-associated membrane protein-2 (synaptobrevin-2) forms a complex with synaptophysin.
Biochem. J.
305
,
721
-724.
Weber, T., Zemelman, B. V., McNew, J. A., Westermann, B., Gmachl, M., Parlati, F., Söllner, T. H. and Rothman, J. E. (
1998
). SNAREpins: minimal machinery for membrane fusion.
Cell
92
,
759
-772.
Weimbs, T., Low, S. H., Chapin, S. J., Mostov, K. E., Bucher, P. and Hofmann, K. (
1997
). A conserved domain is present in different families of vesicular fusion proteins: a new superfamily.
Proc. Natl. Acad. Sci. USA
94
,
3046
-3051.
Whyte, J. R. and Munro, S. (
2002
). Vesicle tethering complexes in membrane traffic.
J. Cell Sci.
115
,
2627
-2637.
Wickner, W. and Haas, A. (
2000
). Yeast homotypic vacuole fusion: a window on organelle trafficking mechanisms.
Annu. Rev. Biochem.
69
,
247
-275.
Xu, T., Rammner, B., Margittai, M., Artalejo, A. R., Neher, E. and Jahn, R. (
1999
). Inhibition of SNARE complex assembly differentially affects kinetic components of exocytosis.
Cell
99
,
713
-722.
Xu, Y., Zhang, F., Su, Z., McNew, J. A. and Shin, Y. K. (
2005
). Hemifusion in SNARE-mediated membrane fusion.
Nat. Struct. Mol. Biol.
12
,
417
-422.
Yang, B., Gonzalez, L., Prekeris, R., Steegmaier, M., Advani, R. J. and Scheller, R. H. (
1999
). SNARE interactions are not selective. Implications for membrane fusion specificity.
J. Biol. Chem.
274
,
5649
-5653.
Yang, B., Steegmaier, M., Gonzalez, L. C., Jr and Scheller, R. H. (
2000
). nSec1 binds a closed conformation of syntaxin1A.
J. Cell Biol.
148
,
247
-252.
Zerial, M. and McBride, H. (
2001
). Rab proteins as membrane organizers.
Nat. Rev. Mol. Cell Biol.
2
,
107
-117.
Zhang, F., Chen, Y., Su, Z. and Shin, Y. K. (
2004
). SNARE assembly and membrane fusion, a kinetic analysis.
J. Biol. Chem.
279
,
38668
-38672.