The tightly coordinated, spatiotemporal control of actin filament remodeling provides the basis of fundamental cellular processes, such as cell migration and adhesion. Specific protein assemblies, composed of various actin-binding proteins, are thought to operate in these processes to nucleate and elongate new filaments, arrange them into complex three-dimensional (3D) arrays and recycle them to replenish the actin monomer pool. Actin filament assembly is not only necessary to generate pushing forces against the leading edge membrane or to propel pathogens through the cytoplasm, but also coincides with the generation of stress fibers (SFs) and focal adhesions (FAs) that generate, transmit and sense mechanical tension. The only protein families known to date that directly enhance the elongation of actin filaments are formins and the family of Ena/VASP proteins. Their mechanisms of action, however, in enhancing processive filament elongation are distinct. The aim of this Review is to summarize our current knowledge on the molecular mechanisms of Ena/VASP-mediated actin filament assembly, and to discuss recent insights into the cell biological functions of Ena/VASP proteins in cell edge protrusion, migration and adhesion.

Members of the Ena/VASP protein family are structurally conserved and present in both multicellular animals and single-celled organisms such as Dictyostelium (Fig. 1A). Drosophila Enabled (Ena), the founding member of this family, was identified through its genetic interaction with Abl tyrosine kinase, and was later found to play an important role in axon guidance in the developing nervous system (Bashaw et al., 2000; Gertler et al., 1995). C. elegans also contains a single Enabled homolog known as UNC-34, which was shown to be crucial for the migration of cells and growth cones in the developing embryo (Fleming et al., 2010). Vertebrates express three Ena-related paralogs – vasodilator-stimulated phosphoprotein (VASP), mouse Ena (Mena, also known as ENAH), and Ena-VASP-like (Evl). VASP was originally isolated from human platelets (Reinhard et al., 1992) and recognized as major substrate for the cAMP- and cGMP-regulated protein kinases PKA and PKG (Lambrechts et al., 2000; Waldmann et al., 1987), but it is also widely expressed in other cells and tissues (Reinhard et al., 1992). Mena and Evl were subsequently identified by sequence similarity (Gertler et al., 1996).

Fig. 1.

Domain organization and structure of Ena/VASP proteins. (A) Representative Ena/VASP proteins from different organisms. All family members share a highly similar, tripartite domain organization. Drosophila Ena (Enabled) and vertebrate Mena (murine Ena) additionally contain a Q-rich region (green) or LERER repeats (yellow), respectively. Numbers indicate amino acid residues (aa). (B) Structure–function relationship of Ena/VASP proteins. The N-terminal EVH1 region contributes to subcellular targeting of Ena/VASP proteins predominantly through interaction with proteins containing FP4 or related proline-rich repeats, but this requires EVH1 to be combined with C-terminal tetramerization (also see Fig. 3). The central proline-rich domain (PRD) of Ena/VASP binds to profilin and profilin–actin complexes, as well as SH3- and WW-containing proteins. The EVH2 domain at the C-terminus encompasses a WH2 type G-actin binding (GAB) motif (blue), a WH2-like F-actin binding module (FAB, red) followed by the tetramerization (Tet) domain (green) mentioned above. (C) Transmission electron microscope images of murine VASP tetramers after rotary shadowing, illustrating its highly flexible architecture (Breitsprecher et al., 2008). Scale bar: 50 nm. Images courtesy of Günther Resch and John Victor Small, Institute of Molecular Biotechnology (IMBA), Austria. (D) Cartoon illustrating a highly flexible VASP tetramer.

Fig. 1.

Domain organization and structure of Ena/VASP proteins. (A) Representative Ena/VASP proteins from different organisms. All family members share a highly similar, tripartite domain organization. Drosophila Ena (Enabled) and vertebrate Mena (murine Ena) additionally contain a Q-rich region (green) or LERER repeats (yellow), respectively. Numbers indicate amino acid residues (aa). (B) Structure–function relationship of Ena/VASP proteins. The N-terminal EVH1 region contributes to subcellular targeting of Ena/VASP proteins predominantly through interaction with proteins containing FP4 or related proline-rich repeats, but this requires EVH1 to be combined with C-terminal tetramerization (also see Fig. 3). The central proline-rich domain (PRD) of Ena/VASP binds to profilin and profilin–actin complexes, as well as SH3- and WW-containing proteins. The EVH2 domain at the C-terminus encompasses a WH2 type G-actin binding (GAB) motif (blue), a WH2-like F-actin binding module (FAB, red) followed by the tetramerization (Tet) domain (green) mentioned above. (C) Transmission electron microscope images of murine VASP tetramers after rotary shadowing, illustrating its highly flexible architecture (Breitsprecher et al., 2008). Scale bar: 50 nm. Images courtesy of Günther Resch and John Victor Small, Institute of Molecular Biotechnology (IMBA), Austria. (D) Cartoon illustrating a highly flexible VASP tetramer.

Members of the Ena/VASP family of proteins encompass three highly conserved domains (Fig. 1B), an N- and C-terminal, homologous domain, thus termed Ena/VASP homology 1 (EVH1) and EVH2, respectively, and a central, proline-rich domain (PRD). The N-terminal EVH1 domain binds most commonly to the so-called FPPPP motif, found in cytoskeletal proteins such as vinculin (Brindle et al., 1996), zyxin (Drees et al., 2000), lamellipodin (Krause et al., 2004) and the Listeria surface protein ActA (Niebuhr et al., 1997; Pistor et al., 1995). However, other cytoskeletal proteins such as the WAVE regulatory complex (WRC) subunit Abi from Drosophila or the human testin LIM-domain protein (Tes) either bind to the EVH1 domain with proline-rich sequences that deviate from the core consensus sequence (W/F)PxϕP motif (x, any residue; ϕ, hydrophobic amino acid) or even utilize an unrelated LIM domain for interaction (Acevedo et al., 2017; Ball et al., 2000; Boëda et al., 2007; Chen et al., 2014). The EVH1 domain has been proposed to mediate subcellular targeting of Eva/VASP proteins (Krause et al., 2003). Nevertheless, the EVH1 domain alone does not appear to be sufficient for robust targeting to all subcellular sites, despite its essential function (see below). The central PRD domain is the least conserved and is rich in clusters of proline residues. It binds to the actin-monomer-binding regulator profilin, and is primarily used in the cellular context for recruitment of profilin–actin complexes to fuel actin assembly (Ferron et al., 2007) and for mediating interactions with signaling proteins containing SH3- or WW-domains, such as FE65 or Irsp53 (also known as APBB1 and BAIAP2, respectively) (Ermekova et al., 1997; Jonckheere et al., 1999; Krugmann et al., 2001; Reinhard et al., 1995). The C-terminal EVH2 domain mediates interactions with monomeric and filamentous actin, as well as tetramerization (Bachmann et al., 1999; Hüttelmaier et al., 1999). The G-actin-binding site (GAB) within the EVH2 domain represents a WASP homology 2 (WH2) motif found in many other actin regulators (Dominguez, 2016; Paunola et al., 2002). The adjacent F-actin-binding site (FAB) has also been proposed to possess WH2-like properties (Dominguez, 2009; Ferron et al., 2007). The VASP tetramer contains long stretches of intrinsically disordered amino acid sequences and is an exceptionally flexible molecule, as revealed by electron microscopy and analytical ultracentrifugation (Breitsprecher et al., 2008) (Fig. 1C,D). The only structured regions of the molecule are the globular N-terminal EVH1 domains and a short right-handed coiled-coil tetramerization domain (Tet) at the C-terminus of the protein (Ball et al., 2000; Kühnel et al., 2004).

In this Review, we start by discussing requirements for subcellular positioning of Ena/VASP proteins followed by a brief summary of the initial difficulties and controversies of the biochemical properties of this fascinating family of actin polymerases, and then sketch our current understanding of the molecular mechanisms of Ena/VASP-mediated actin filament assembly in vitro and in vivo. This will lead us to elaborate on the functions of Ena/VASP proteins in cell edge protrusion, adhesive two-dimensional (2D)-migration and cell adhesion. We will also briefly highlight the potential relevance of Ena/VASP member phosphorylation, although concerning this topic and further detail, readers are mostly referred to several, excellent reviews (Blanchoin et al., 2014; Drees and Gertler, 2008; Krause and Gautreau, 2014; Romero et al., 2020; Trichet et al., 2008).

Our body cells can harbor dozens of distinct actin filament-containing structures, the relative extent of formation of which will of course depend on state of differentiation or cell-specific functions (Rottner et al., 2017). This means that even non-muscle cells will harbor contractile pseudo-sarcomeric structures in the form of stress fibers (SFs) and their anchorage sites, focal adhesions (FAs), and at the same time frequently highly dynamic protrusive structures at their motile fronts (see model in Fig. 2A). The most prominent, protrusive structures include the sheet-like lamellipodia and the finger-like, actin filament bundle-containing filopodia. If bundled arrays remain embedded into lamellipodial actin networks, they are termed microspikes (Small et al., 2002).

Fig. 2.

Subcellular localization of Ena/VASP proteins in different cell types. (A) Schematic illustration of specific localizations of Ena/VASP proteins to most prominent actin-based structures of mesenchymal (left) and epithelial cells (right). Only ventral SFs are shown for clarity. (B) Endogenous VASP at the leading edge and tips of microspikes in B16-F1 mouse melanoma cells. (C) Endogenous VASP at the leading edge and tips of microspikes and filopodia in neuronal CAD cells. (D) Ectopically expressed Evl targeted to FAs and ventral SFs of NIH 3T3 fibroblasts. (E) VASP at the tip of a dorsal ruffle in an NIH 3T3 cell. (F) Mena accumulating in cell–cell contacts of an MDCK II monolayer. (G) VASP enriching at the rear surface of L. monocytogenes, which promotes its intracellular motility in infected U2OS ostersarcoma cells. Ena/VASP proteins are shown in green, F-actin in red and DAPI in magenta (in E). Scale bars: 20 µm (B–D,F); 10 µm (E,G).

Fig. 2.

Subcellular localization of Ena/VASP proteins in different cell types. (A) Schematic illustration of specific localizations of Ena/VASP proteins to most prominent actin-based structures of mesenchymal (left) and epithelial cells (right). Only ventral SFs are shown for clarity. (B) Endogenous VASP at the leading edge and tips of microspikes in B16-F1 mouse melanoma cells. (C) Endogenous VASP at the leading edge and tips of microspikes and filopodia in neuronal CAD cells. (D) Ectopically expressed Evl targeted to FAs and ventral SFs of NIH 3T3 fibroblasts. (E) VASP at the tip of a dorsal ruffle in an NIH 3T3 cell. (F) Mena accumulating in cell–cell contacts of an MDCK II monolayer. (G) VASP enriching at the rear surface of L. monocytogenes, which promotes its intracellular motility in infected U2OS ostersarcoma cells. Ena/VASP proteins are shown in green, F-actin in red and DAPI in magenta (in E). Scale bars: 20 µm (B–D,F); 10 µm (E,G).

Ena/VASP proteins are long known to accumulate at all these sites of active actin assembly (Fig. 2A). Specifically, VASP, Mena and Evl have been shown to localize to FAs and SFs in a dotty fashion, to the protruding lamellipodial leading edge as a continuous line, and the distal tips of microspikes and filopodia in the form of bright spots (Damiano-Guercio et al., 2020; Gertler et al., 1996; Lanier et al., 1999; Reinhard et al., 1992; Rottner et al., 1999). Moreover, these proteins accumulate at the surfaces of bacterial pathogens, such as Listeria monocytogenes (Chakraborty et al., 1995), as well as in phagosomes (Coppolino et al., 2001), dorsal ruffles (Michael et al., 2010), invadopodia (Philippar et al., 2008) and in epithelial cell–cell contacts (Vasioukhin et al., 2000). For examples for most of these distinct localization patterns, see Fig. 2B–G.

The EVH1 domain has been proposed to mediate subcellular targeting of Eva/VASP proteins (Bear et al., 2000; Krause et al., 2003). Despite its essential function, however, the EVH1 domain alone does not mediate robust targeting to all subcellular sites. Although the VASP-EVH1 domain, for instance, has been reported to be sufficient at least for targeting to filopodial tips, it failed to robustly accumulate at the rim of lamellipodia and in FAs (Fig. 3; Applewhite et al., 2007). This is true both in the presence and absence of endogenous Ena/VASP proteins. Thus, additional domains of Ena/VASP proteins mediating interactions with actin or other actin-binding factors were hitherto concluded to be necessary for appropriate, subcellular targeting. However, the availability of true Ena/VASP family knockout cell lines now allows defining the minimal domains required for subcellular Ena/VASP targeting. Indeed, Fig. 3 (bottom row) exemplifies that the EVH1 domain alone can mediate all aspects of VASP subcellular distribution, as long as it is combined with the tetramerization domain to increase its functional affinity and residence time. This shows that, as opposed to previous assumptions (Loureiro et al., 2002), the F-actin-binding EVH2 domain is dispensable for proper subcellular targeting.

Fig. 3.

The EVH1 and tetramerization domains are required for appropriate subcellular positioning of VASP. Representative images of B16-F1 and Ena/VASP-deficient EVM-KO cells (Damiano-Guercio et al., 2020) migrating on laminin, expressing indicated constructs (green) and counterstained for F-actin (red). The EVH1 domain alone fused to EGFP (EGFP–EVH1) accumulated only very weakly at the tips of microspikes (Ms) in B16-F1 wild-type cells (white arrowheads) and was not detectable at the tip of the lamellipodium (Lp) and stress fibers (SFs). Similar results were observed in Ena/VASP-deficient EVM-KO cells (EGFP–EVH1, middle panel), even though in addition, this cell line is virtually devoid of microspikes (also see Fig. 5). In stark contrast, the EVH1 domain combined with the short Tet motif (EGFP–EVH1–Tet, bottom panel), mediating tetramerization of the construct, was fully sufficient for its prominent accumulation at both leading edge and stress fibers in EVM-KO cells. Scale bars: 20 µm.

Fig. 3.

The EVH1 and tetramerization domains are required for appropriate subcellular positioning of VASP. Representative images of B16-F1 and Ena/VASP-deficient EVM-KO cells (Damiano-Guercio et al., 2020) migrating on laminin, expressing indicated constructs (green) and counterstained for F-actin (red). The EVH1 domain alone fused to EGFP (EGFP–EVH1) accumulated only very weakly at the tips of microspikes (Ms) in B16-F1 wild-type cells (white arrowheads) and was not detectable at the tip of the lamellipodium (Lp) and stress fibers (SFs). Similar results were observed in Ena/VASP-deficient EVM-KO cells (EGFP–EVH1, middle panel), even though in addition, this cell line is virtually devoid of microspikes (also see Fig. 5). In stark contrast, the EVH1 domain combined with the short Tet motif (EGFP–EVH1–Tet, bottom panel), mediating tetramerization of the construct, was fully sufficient for its prominent accumulation at both leading edge and stress fibers in EVM-KO cells. Scale bars: 20 µm.

Ena/VASP recruitment to FAs requires mostly zyxin (Hoffman et al., 2006), with contributions from additional FPPPP motif-harboring proteins, such as vinculin or palladin. The latter has also been implicated in the recruitment of VASP to non-contractile, so called dorsal SFs (Gateva et al., 2014; Hotulainen and Lappalainen, 2006). An additional level of complexity comes from the EVH1 ligand Tes, which is recruited to FAs by zyxin, but can compete with canonical FPPPP-containing ligands for the interaction with Mena (Boëda et al., 2007; Garvalov et al., 2003). Another EVH1 interactor, proline-rich EVH1 ligand (PREL1; also known as Rap1-GTP-interacting adaptor molecule, RIAM), appears to trigger Ena/VASP accumulation in protrusions and FAs downstream of Rap/Ras signaling (Jenzora et al., 2005; Lafuente et al., 2004). The PREL/RIAM-related lamellipodin (Lpd; also known as RAPH1) has been proposed to be obligatory for Ena/VASP targeting to the leading edge of lamellipodia (Krause et al., 2004), but its genetic disruption recently disproved this view (Dimchev et al., 2020).

The formation of lamellipodial actin networks occurs downstream of Rac-triggered activation of the Arp2/3 complex by the WRC (Rottner et al., 2021). In Dictyostelium, the latter was shown to operate upstream of Ena/VASP accumulation in pseudopods (Litschko et al., 2017). This finding was recently confirmed in mammalian cells lacking the WRC, but harboring non-canonical, lamellipodia-like structures (Kage et al., 2021 preprint). Interestingly, Dictyostelium and Drosophila Abi have been shown to bind to Ena/VASP through EVH1 (Chen et al., 2014; Litschko et al., 2017), thus likely mediating Ena/VASP accumulation in pseudopodia and lamellipodia. The precise recruitment features to additional subcellular structures, such as along SFs or to adherens junctions of epithelial or endothelial cells, remain to be established.

Previous biochemical analyses quickly established the function of individual domains of vertebrate Ena/VASP proteins in filament bundling, interaction with monomeric actin and tetramerization. However, owing to the rather low affinity of the vertebrate GAB motif for actin monomers (Breitsprecher et al., 2011), it is not surprising that earlier work reported conflicting results on the impact of Ena/VASP proteins on actin assembly, as relatively low concentrations of actin and Ena/VASP in buffers with different salt concentrations were used (Trichet et al., 2008). The situation changed upon analysis of Dictyostelium VASP, which harbors a high-affinity GAB (Breitsprecher et al., 2008, 2011). Using total internal reflection fluorescence (TIRF) microscopy with purified proteins, it could be shown that both human and Dictyostelium VASP are directly involved in accelerating filament elongation by delivering actin monomers to the growing barbed end (Breitsprecher et al., 2008). This study further revealed that, in bulk solution, Dictyostelium VASP accelerates actin filament elongation up to sevenfold as compared to the approximately twofold acceleration by human VASP, although both actin polymerases were readily inhibited by low concentrations of heterodimeric capping protein (CP; comprising e.g. human CAPZA1 and CAPZB), which binds tightly to barbed ends to terminate filament growth (Maun et al., 1996; Yamashita et al., 2003). In striking contrast, dense clustering of neighboring VASP tetramers in the range of 10 nm on functionalized beads enabled long-lasting processive filament elongation that became virtually insensitive even to high concentrations of CP (Breitsprecher et al., 2008) (Fig. 4A).

Fig. 4.

Molecular mechanism of VASP-mediated filament elongation in solution and in clustered arrays. (A) Representative images from time-lapse movies illustrating acceleration of filament elongation by Dictyostelium VASP and its inhibition by CP in solution. In contrast, on VASP-derivatized beads, filament elongation becomes long-lasting and virtually resistant to CP. Scale bar: 10 µm. Image courtesy of Thomas Pokrant, Hannover Medical School, Germany. (B) General mechanism of VASP-mediated actin assembly. The VASP tetramer is continuously attached to the filament-barbed end by at least one EVH2 domain during filament elongation. Free EVH2 domains can capture actin monomers from solution with an on-rate constant kon, and subsequently either transfer the monomers onto the barbed end with a transfer rate kt to drive filament assembly, or release them back into solution with an off-rate koff. Through this mode of action, the VASP tetramer is able to processively track and elongate growing barbed ends. (C) In solution, processive association of VASP with the barbed end is rather short lived, and therefore competes with the spontaneous addition of actin monomers to the barbed end through the direct pathway, with an independent transfer rate constant kf. Free barbed ends are then accessible for either other VASP tetramers, actin monomers or capping protein (CP). The average elongation rate governing filament growth is therefore composed of the rates of VASP-mediated filament elongation (rv) and of the direct pathway (rd). (D) VASP clustering on surfaces, such as beads or the plasma membrane, leads to processive association of VASP with growing filament ends, efficiently blocking barbed end access for CP and the spontaneous addition of monomers. Thus, in clustered arrays, filament elongation is exclusively driven by actin monomers recruited and transferred by VASP. For more details, see Breitsprecher et al. (2011).

Fig. 4.

Molecular mechanism of VASP-mediated filament elongation in solution and in clustered arrays. (A) Representative images from time-lapse movies illustrating acceleration of filament elongation by Dictyostelium VASP and its inhibition by CP in solution. In contrast, on VASP-derivatized beads, filament elongation becomes long-lasting and virtually resistant to CP. Scale bar: 10 µm. Image courtesy of Thomas Pokrant, Hannover Medical School, Germany. (B) General mechanism of VASP-mediated actin assembly. The VASP tetramer is continuously attached to the filament-barbed end by at least one EVH2 domain during filament elongation. Free EVH2 domains can capture actin monomers from solution with an on-rate constant kon, and subsequently either transfer the monomers onto the barbed end with a transfer rate kt to drive filament assembly, or release them back into solution with an off-rate koff. Through this mode of action, the VASP tetramer is able to processively track and elongate growing barbed ends. (C) In solution, processive association of VASP with the barbed end is rather short lived, and therefore competes with the spontaneous addition of actin monomers to the barbed end through the direct pathway, with an independent transfer rate constant kf. Free barbed ends are then accessible for either other VASP tetramers, actin monomers or capping protein (CP). The average elongation rate governing filament growth is therefore composed of the rates of VASP-mediated filament elongation (rv) and of the direct pathway (rd). (D) VASP clustering on surfaces, such as beads or the plasma membrane, leads to processive association of VASP with growing filament ends, efficiently blocking barbed end access for CP and the spontaneous addition of monomers. Thus, in clustered arrays, filament elongation is exclusively driven by actin monomers recruited and transferred by VASP. For more details, see Breitsprecher et al. (2011).

Based on thermodynamic data, the determination of the association rate constants of G-actin to human and Dictyostelium GAB, as well as TIRF microscopy, a quantitative mathematical model of VASP-mediated filament elongation could be developed (Breitsprecher et al., 2011). It is derived from the ‘actoclampin’ model of actin filament end-tracking proteins (Dickinson and Purich, 2002; Dickinson et al., 2004) in which processive elongation is achieved by alternating, multivalent and affinity-modulated interactions between the VASP tetramer and the filament barbed end (Fig. 4B–D). The model implies that, at steady state, only one polypeptide chain of the VASP tetramer is bound to the terminal filament subunit at any given time, leaving a number (N) of free GABs to recruit actin monomers from solution. Captured monomers can then be either released back into solution or are transferred onto the filament tip and irreversibly incorporated, so that the rate of VASP-mediated filament elongation scales with the saturation of GABs with actin monomers (Breitsprecher et al., 2011). This not only explains why Dictyostelium VASP carrying a high-affinity GAB is more efficient when compared with its human counterpart at the low actin concentrations used in in vitro assays, but also implies that, due to the much higher cellular actin concentration, all Ena/VASP proteins are fully saturated with actin in vivo and act as powerful actin polymerases (Breitsprecher et al., 2011). Global fitting of experimental data for filament elongation from beads coated with VASP mutants carrying different GAB domains yielded N=3, formally suggesting that a single VASP tetramer operates at a filament tip during filament elongation (Breitsprecher et al., 2011). In fact, processive, but relatively short-lived, filament elongation has also been shown with single vertebrate and Drosophila Ena/VASP tetramers (Hansen and Mullins, 2010; Winkelman et al., 2014). Since single VASP tetramers are only transiently associated with barbed ends, monomers can also be added spontaneously to filament ends in a parallel pathway (termed the direct pathway). Thus, the total elongation rates observed with VASP in solution constitute the sum of the rate of VASP-mediated filament elongation and that resulting from the direct pathway (Breitsprecher et al., 2011). However, the underlying mechanism in clustered arrays is profoundly different, since in the latter, processive VASP-mediated filament elongation becomes exceptionally long-lasting and resistant against CP (Breitsprecher et al., 2008, 2011). Experimental testing of the model by variation of oligomerization states and increase of GAB numbers on individual polypeptide chains indeed revealed that – within a broad range – the rates of VASP-mediated actin filament elongation are directly proportional to the number of free GABs in solution (Brühmann et al., 2017). In contrast, in surface-bound VASP clusters, formed by the different oligomerization mutants, N is not only fixed to three free GABs, but additionally shows that VASP molecules from distinct oligomers or VASP monomers can synergize in the processive elongation of single actin filaments (Brühmann et al., 2017). Notably, processivity, which additionally correlates with the affinity of the FAB for filament side binding, increases continuously with Ena/VASP oligomerization (Breitsprecher et al., 2011; Brühmann et al., 2017). Thus, the VASP tetramer appears to be the evolutionarily optimal form to drive processive actin assembly from clustered arrays in the presence of CP.

Actin bound to profilin is assumed to be the main polymerization-competent form of monomeric actin in cells (Kaiser et al., 1999; Pollard and Borisy, 2003). However, an open issue in the field is the question as to whether Ena/VASP utilizes profilin to accelerate actin assembly in vivo, as typically seen with formins (Kovar et al., 2006), or whether it might also operate with pure actin and if so to what extent. The co-crystallization of mammalian profilin–actin in complex with a peptide spanning the PRD and GAB of murine VASP actually suggests that this region has an important function in recruitment and subsequent delivery of monomeric actin to the GAB domain (Ferron et al., 2007). However, although one study found no positive effect of human or Dictyostelium profilin on VASP-mediated filament elongation in vitro (Breitsprecher et al., 2008), others have reported that VASP and profilin enhance the rate of actin filament elongation (Hansen and Mullins, 2010; Pasic et al., 2008). One important difference between these studies is how actin assembly rates were compared. While Breitsprecher et al. explored VASP-mediated actin assembly in the presence or absence of profilin, as typically done in the formin field, the latter studies only compared rates of actin filament elongation in the presence of VASP and profilin–actin with those observed with profilin–actin alone, precluding assessment of the direct effect of profilin on Ena/VASP-mediated actin assembly. In our more recent work, we also analyzed different profilin isoforms, and found that Ena/VASP proteins can indeed effectively assemble filaments from profilin–actin, albeit with approximately only half the rates than seen with pure actin alone (Brühmann et al., 2017). Unexpectedly, in a very recent study using profilin 1-deficient CAD cells, Ena/VASP was concluded to be non-functional despite its increased localization at the protruding cell edge (Skruber et al., 2020). However, given that at least in vitro, Ena/VASP proteins can use both pure actin and profilin-actin complexes for actin assembly, this conclusion, derived from mitochondrial Ena/VASP sequestration and lamellipodial F-actin intensity measurements, seems counterintuitive. Considering the established positive correlation between protrusion and Ena/VASP accumulation in wild-type lamellipodia (Rottner et al., 1999), the hypothesis put forward by Skruber et al. deserves to be revisited in more direct assays in the future.

Consistent with the positive regulatory role in both actin filament elongation in vitro and the efficiency of actin assembly-dependent motility of intracellular pathogens (see Box 1), Ena/VASP proteins were originally considered to be positive regulators and effectors of lamellipodia protrusion. Historically, for instance, VASP was the first actin-binding protein shown to dynamically accumulate at the edges of protruding lamellipodia in a fashion that linearly correlated with protrusion rates, and thus active actin polymerization (Rottner et al., 1999). This behavior had perfectly fitted earlier knowledge on the positive regulatory function of Ena/VASP proteins on the surface of Listeria (Sechi and Wehland, 2004). In other words, rapid protrusion perfectly correlated with maximal VASP accumulation at lamellipodial leading edges, and a halt in protrusion or lamellipodial retraction resulted in the complete vanishing of Ena/VASP from these sites (Rottner et al., 1999). Importantly, these activities are completely independent of association of lamellipodia with the substratum, which occurs through nascent adhesions or focal complexes that Ena/VASP members also associate with, or from their recruitment to other subcellular structures, such as cytoplasmic SFs (Rottner et al., 1999, 2001). Importantly, these Ena/VASP dynamics must also be distinguished from their general association with the entire plasma membrane, as mediated for instance by factors equipped with a C-terminal CAAX-box, originally identified in Ras or Rho GTPases, and destined for membrane association through prenylation (Roberts et al., 2008). Around the same time, in their seminal studies, Gertler and colleagues found, surprisingly, that effective fibroblast migration negatively correlated with Ena/VASP expression, mostly using Mena- and VASP-deficient cells that have been screened for lowest possible Evl expression, termed MVD7 (Bear et al., 2000). Aside from analyses using MVD7, Ena/VASP overexpression or their inhibition by non-productive mistargeting, for instance to mitochondrial surfaces was also used (Bear et al., 2000). Consistent with the above results, this particular study concluded that Ena/VASP also inhibits rates of membrane extension and retraction (Bear et al., 2000), contradicting the view of positive regulatory roles on Listeria motility and lamellipodia protrusion (Rottner et al., 1999; Sechi and Wehland, 2004). One complication in the interpretation of this study is that MVD7, unlike the original assumptions, later turned out not to be truly Ena/VASP-deficient, as they still expressed significant levels of Evl (Damiano-Guercio et al., 2020). However, the original hypothesis of a negative effect of Ena/VASP on actin assembly at the leading edges of lamellipodia was subsequently corrected by the discovery of a positive role. The latter positive regulatory function, was concluded to counteract the slow and persistent advancement of the cell front presumed to accompany effective fibroblast migration (Bear et al., 2002). More specifically, lamellipodia that protruded too rapidly as a result of high activities of Ena/VASP were now proposed to be prone to rapid and frequent collapse, thus being incapable of mediating the slow but efficient leading edge advance in fibroblast migration. This view seemingly solved the original discrepancies that had arisen when collectively considering all those observations. Hence, since then, the consensus view in the field was that actin polymerases that act at lamellipodial leading edges, such as VASP or, for instance the formin and Cdc42 effector FMNL2, promote actin filament elongation, thereby increasing the spacing between Arp2/3-dependent branches in such actin networks. Of note, the continuous branch formation by Arp2/3 complex is considered to be an inherent feature of the formation of actin networks such as lamellipodia, which must therefore be coordinated with the activities of various actin regulators during actin-dependent protrusion (Dimchev et al., 2021; Krause and Gautreau, 2014). The increase of branch spacing at the individual filament level as caused by Ena/VASP would thus decrease the stability and protrusion persistence of these structures. Conversely, reduction of the activity of Ena/VASP or FMNL2 would enhance the density of Arp2/3-dependent branches, the persistence of protrusion and lamellipodial force development, which could indeed make a lot of sense as beautifully summarized about one decade later (Krause and Gautreau, 2014). However, subsequent studies by various labs clearly established that the relative dependencies of all those parameters, that is lamellipodium protrusion versus actin assembly rate versus cell migration, are far more complex than originally anticipated (Damiano-Guercio et al., 2020; Kage et al., 2017; Law et al., 2021).

Box 1: Ena/VASP functions in Listeria motility

Listeria monocytogenes, a food-born human pathogen, can cause serious medical problems, such as meningitis or miscarriage. The gram-positive bacterium hides within the cytoplasm of infected cells and from there can spread into neighboring host cells, thereby effectively escaping the humoral immune response. Spreading occurs through active actin polymerization that pushes the bacterium forward, accompanied by formation of a transient, actin filament comet tail (Welch and Way, 2013). Interestingly, Listeria binds both to the Arp2/3 complex (Domann et al., 1992; Kocks et al., 1992) and Ena/VASP (Chakraborty et al., 1995). Binding to the latter requires interactions between their EVH1 domains and the ActA consensus motif E/DFPPPPXD/E, which is repeated four times in ActA and also present in host cell vinculin and zyxin (Niebuhr et al., 1997). Although this domain can redirect VASP and actin accumulation to various subcellular compartments, including mitochondria (Fradelizi et al., 2001; Pistor et al., 1994), this activity only constitutes half of the fascinating story of Listeria-mediated actin comet formation. In addition, seminal studies by the groups of Mitchison and Carlier have established ActA as a key determinant of Arp2/3 complex-mediated Listeria motility (Loisel et al., 1999; Welch et al., 1998). The in vitro reconstitution of actin-based Listeria motility using purified proteins confirmed that although the Arp2/3 complex is obligatory for actin assembly on the bacterial surface, Ena/VASP family members are not essential, but increase the efficiency of the process (Loisel et al., 1999), which is highly analogous to their role in lamellipodia (Damiano-Guercio et al., 2020). The relevance of Ena/VASP proteins in Listeria motility was also explored in MVD7 cells lacking Mena and VASP (Bear et al., 2000; Geese et al., 2002). Although these studies could clearly establish the effects of Ena/VASP gene dose on Listeria motility and successfully explore the relevance of specific Ena/VASP domains, the reported data may have to be recapitulated, at least partly, since MVD7 cells still express Evl (Auerbuch et al., 2003; Damiano-Guercio et al., 2020). In addition, suggestions for how precisely Ena/VASP contribute to Arp2/3 complex-mediated actin assembly at the Listeria surface range from the promotion of Arp2/3-mediated branched nucleation (Skoble et al., 2001) to inducing the dissociation of ActA from the branch junction (Samarin et al., 2003). Considering the recently described promotion of WRC- and Arp2/3-dependent actin assembly by the Ena/VASP antagonist CP (Funk et al., 2021), and extrapolating this mechanism to ActA-dependent processes, we feel it will certainly be instrumental to tease out the precise functions of both CP and Ena/VASP at the Listeria surface in the future.

Specifically, according to this simple model (Krause and Gautreau, 2014) and with respect to FMNL formin activities, the expected phenotypes of removal of FMNL2 and its related family member FMNL3, both of which accumulate at the tips of protruding lamellipodia (Kage et al., 2017), were not only a reduction in protrusion and polymerization rate of the network, but also an increase in actin filament densities, Arp2/3 complex incorporation and exerted protrusion strength. However, none of the expectations were correct, except for reduced protrusion rates; instead, FMNL2/3 KOs exhibited unchanged Arp2/3 complex densities and actin network polymerization rates, despite the decreased F-actin densities and exerted protrusion forces (Kage et al., 2017). Based on all these observations, the conclusion was that FMNL formins (as opposed to Ena/VASP) polymerize a subpopulation of lamellipodial actin filaments, thereby contributing to the overall F-actin concentration and thickness of the lamellipodium, as well as its protrusion efficiency and strength, without directly affecting the Arp2/3 complex-dependent branched actin network (Kage et al., 2017). These conclusions were also confirmed by mathematical modeling (Dolati et al., 2018). However, the story does not end here, as the complete removal of all three Ena/VASP family members, which was recently achieved through CRISPR/Cas9-mediated genome editing (Damiano-Guercio et al., 2020) led to a phenotype that significantly differed from that obtained upon formin depletion, and contradicted previous proposals and models (Bear et al., 2000; Krause and Gautreau, 2014). Upon complete loss of Ena/VASP, lamellipodia still formed and protruded and assembled actin more slowly and with increased Arp2/3 complex incorporation, as would be expected from previous models, but the cells displayed significantly reduced actin network densities with markedly perturbed actin filament architecture (Fig. 5), as well as an at best unchanged, but certainly not increased, lamellipodial persistence (Damiano-Guercio et al., 2020). In analogy, a more recent characterization of Nance–Horan syndrome-like 1 protein (NHSL1), an inhibitor of the WRC, the main Arp2/3 complex activator in canonical lamellipodia, revealed that increased lamellipodial Arp2/3 complex activities and F-actin intensities in lamellipodia do not necessarily give rise to reduced protrusion speed. More specifically, NHSL1-disrupted cell lines displayed unchanged protrusion rates in spite of increased Arp2/3 complex and F-actin densities, but compromised lamellipodia stability (Law et al., 2021). Likewise, removal of the prominent WRC antagonist CYRI (CYFIP-related Rac interactor) increased protrusion plasticity, i.e. the extent of protrusion and retraction (Fort et al., 2018), again distinct from expectations of a model, in which Arp2/3-dependent branching density would positively and negatively correlate with protrusion persistence and speed, respectively (Krause and Gautreau, 2014).

Fig. 5.

Loss of Ena/VASP perturbs lamellipodia and abrogates microspike formation in B16-F1 cells. (A) Representative Airyscan confocal images of wild-type B16-F1 and EVM-KO cells allowed to spread and migrate on laminin and stained for the actin cytoskeleton. Scale bar: 5 µm. (B) Electron tomography of ultrastructural changes in lamellipodial actin networks upon loss of Ena/VASP family members. Transmission electron micrographs of negatively stained cytoskeletons of protrusive edges of a representative wild-type B16-F1 versus an EVM-KO cell (left). 2D projections of digital 3D tomograms of the same fields showing actin filament trajectories in blue (right). Scale bar: 100 nm. For further details see Damiano-Guercio et al. (2020). Images in B courtesy of Jan Müller, Maria Nemethova and Michael Sixt, Institute of Science and Technology (IST), Austria.

Fig. 5.

Loss of Ena/VASP perturbs lamellipodia and abrogates microspike formation in B16-F1 cells. (A) Representative Airyscan confocal images of wild-type B16-F1 and EVM-KO cells allowed to spread and migrate on laminin and stained for the actin cytoskeleton. Scale bar: 5 µm. (B) Electron tomography of ultrastructural changes in lamellipodial actin networks upon loss of Ena/VASP family members. Transmission electron micrographs of negatively stained cytoskeletons of protrusive edges of a representative wild-type B16-F1 versus an EVM-KO cell (left). 2D projections of digital 3D tomograms of the same fields showing actin filament trajectories in blue (right). Scale bar: 100 nm. For further details see Damiano-Guercio et al. (2020). Images in B courtesy of Jan Müller, Maria Nemethova and Michael Sixt, Institute of Science and Technology (IST), Austria.

Together, these data show that the interrelationships of all these lamellipodial parameters are much more complex than previously anticipated. Indeed, despite the comparable activities of given actin-binding proteins at the single-filament level (such as the actin polymerase function of Ena/VASP and formins), the precise phenotypic outcome of interfering with the function of the different factors in cells can be surprisingly distinct. Thus, the mechanism of how Ena/VASP counteracts Arp2/3 activity and incorporation into lamellipodia deserves further in-depth analysis and elucidation.

Aside from all this, it should be additionally emphasized that increased protrusion activity, as mediated for instance by FMNL or Ena/VASP activity at lamellipodial edges, does not necessarily translate into abrogated 2D migration, as had been deduced from earlier observations (Bear et al., 2002). Furthermore, it is difficult to exclude cell-type-specific effects when studying highly complex processes such as migration, the efficiency of which depends on a myriad of parameters, but our recent studies have clearly established that reduced protrusion efficiency (e.g. upon Ena/VASP removal) can indeed positively correlate with reduced migration efficiency (Damiano-Guercio et al., 2020). As this positive correlation was observed in both B16-F1 melanoma cells and NIH 3T3 fibroblasts, it is unlikely that this discrepancy with earlier studies can solely be explained by cell-type-specific differences. Instead, we conclude that the efficiency of cell migration may differentially depend on the effectivity of lamellipodium protrusion in distinct cell types, so neither characteristic can be unambiguously deduced indirectly by just exploring the respective other. However, in highly motile cell types, such as the model cell line B16-F1 melanoma, we clearly and repeatedly see a positive correlation between capability or efficiency of lamellipodia protrusion and cell migration (Damiano-Guercio et al., 2020; Kage et al., 2017; Schaks et al., 2018).

Genetic disruption of all Ena/VASP family members in B16-F1 mouse melanoma cells strongly decreased lamellipodial F-actin content and the width of lamellipodia, hence overall actin filament mass (Damiano-Guercio et al., 2020). However, the most striking phenotype was the virtual elimination of microspike bundles within the remaining lamellipodia (see Figs 3 and 5); in wild-type cells, these consist of bundled lamellipodial filaments that appear to translocate laterally within lamellipodial actin networks driven by actin polymerization (Small et al., 2002). However, as other actin polymerases can also contribute to microspike formation, such as the FMNL subfamily of formins (Kage et al., 2017), it is not entirely clear why Ena/VASP removal had such a strong defect. One possibility is that microspike bundles, which can be distinguished from bona fide filopodia owing to their almost complete embedding into lamellipodial actin networks (see above), are triggered non-specifically by the sheer amount of lamellipodial filaments, so that a reduction of actin filament mass might already be sufficient for a decrease in microspike bundles. This would be consistent both with the correlation between reduced F-actin mass and microspike numbers upon FMNL2 and FMNL3 knockout (Kage et al., 2017), and explain the almost complete removal of these structures upon Ena/VASP disruption despite the assumed presence of FMNLs in the lamellipodia of these cells (Damiano-Guercio et al., 2020). However, even though it seems tempting to speculate that filament bundling into microspikes is abolished if F-actin mass in the lamellipodium falls below a certain threshold, this exclusive explanation is unlikely as FMNL overexpression – which is known to increase lamellipodial F-actin mass (Dimchev et al., 2017) – does not rescue microspike formation in EVM-knockout cells (EVM-KO) lacking all Ena/VASP family members (Damiano-Guercio et al., 2020). Thus, microspike formation appears to require specifically Ena/VASP activity, although the underlying molecular pathway has hitherto remained unclear.

On a different note, the loss of microspikes did not directly translate into similarly strong defects in the formation of canonical filopodia (Damiano-Guercio et al., 2020), which are capable of protruding beyond and independently of lamellipodial actin networks (Small et al., 2002). To our knowledge, this provides the first direct experimental evidence for the notion that filopodia and microspikes should indeed be considered as independent entities, and are not interchangeable, as proposed based on previous observations that suggested that filopodia can in principle emerge from microspikes (Svitkina et al., 2003). Yet, and partly consistent at least with previous observations of strongly diminished filopodia formation in VASP-null Dictyostelium mutants (Han et al., 2002), Ena/VASP-deficient cells displayed moderately reduced filopodia numbers as well, if induced, for instance, by Arp2/3 complex inhibition (Damiano-Guercio et al., 2020). In contrast, filopodia nucleated by constitutively active mDia2 (also known as DIAPH3), a Diaphanous-related formin, appeared to form completely independently of Ena/VASP activity (Damiano-Guercio et al., 2020). Of note, cells can be coerced to form filopodia by various means (see, for example, Rottner and Schaks, 2019), including the suppression of lamellipodia (as for instance by Arp2/3 complex inhibition, see above) or by employing constitutively active formins, in particular mDia2, which is particularly effective in this context (Block et al., 2008; Yang et al., 2007). It has hitherto remained unknown, however, what the obligatory factors are in filopodia formation or whether this means that filopodia could be formed by various redundant mechanisms. Notwithstanding this, Ena/VASP family members are surprisingly relevant for the formation of microspike bundles, and much less so for filopodia, irrespective of how they are formed. Given the fact that the literature uses the terms microspikes and filopodia in a largely synonymous fashion, the observation that VASP removal only affects microspikes but not filopodia, provides us with an experimental tool to distinguish between those related, yet clearly distinct, structures in the future.

FAs are the main cellular structures linking the intracellular cytoskeleton of adherent cells to the extracellular matrix (ECM) on rigid substrates. They are large macromolecular protein complexes, relevant for adhesion and cell migration by generating, transmitting and sensing mechanical tension (Burridge and Guilluy, 2016; Winograd-Katz et al., 2014; Zaidel-Bar et al., 2007). Proper FA formation and turnover is also crucial for efficient cell spreading, for example, after trypsinization of adherent cells. As revealed by super-resolution imaging, the architecture of FAs comprises stratified layers of distinct proteins that function together to transmit forces sensed by clustered, plasma membrane-spanning integrin receptors to actin filaments (Kanchanawong et al., 2010); here, VASP resides together with zyxin and α-actinin in the so-called ‘actin-regulatory layer’ in close proximity to SF ends. Although the prominent FA proteins zyxin and vinculin can both interact with Ena/VASP (Brindle et al., 1996; Drees et al., 2000), only zyxin-deficient fibroblasts were shown to exhibit severely reduced accumulation of VASP and Mena in FAs as compared to control (Hoffman et al., 2006). Notably, zyxin was also found to recruit Ena/VASP and α-actinin to sites of SF strain to mediate the repair of damaged bundles (Smith et al., 2010). Moreover, AMP-activated protein kinase (AMPK)-mediated phosphorylation of VASP has been proposed to terminate actin assembly in FAs, thereby increasing tension and driving conversion of transverse arcs and non-contractile, dorsal SF into mature contractile ventral SFs (Tojkander et al., 2015). Notwithstanding this, in spite of their prominent accumulation in FAs, the precise and specific roles of Ena/VASP proteins in cell substrate adhesion and FA formation have remained under debate, as the conclusions from multiple studies present an incoherent picture. For instance, VASP-deficient fibroblasts have been reported to have thicker and more stable SFs and enlarged FAs as compared to control cells (Galler et al., 2006), whereas others have reported that depletion of VASP impairs FA formation in LX2 liver myofibroblasts (Kang et al., 2010). Another study implicated VASP in the splitting of FAs during their maturation, which, interestingly, correlates with an increase in FA-associated tension (Young and Higgs, 2018). In spite of this, no conclusive evidence was found previously for changes in cell spreading or differences in the composition or appearance of FAs in fibroblasts that lacked Mena and VASP compared to controls (Bear et al., 2000). Thus, considering the high degree of functional redundancy within this protein family, the analyses of only partially depleted cells likely might have prevented the accurate assessment of the contributions of Ena/VASP proteins to integrin-dependent adhesion. Indeed, our analysis of CRISPR/Cas9-generated mutant cell lines devoid of all three Ena/VASP members first of all showed that Ena/VASP-deficient B16-F1 cells or fibroblasts exhibit severe spreading defects on laminin or fibronectin, respectively (Damiano-Guercio et al., 2020). Of note, in these experiments, Evl rescued the spreading defect more effectively than the other Ena/VASP family members. Complete loss of Ena/VASP in fibroblasts also noticeably perturbed FA formation and impaired integrin-mediated adhesion, as verified by markedly diminished traction forces (Damiano-Guercio et al., 2020). Interestingly, once again, Evl proved most effective in rescuing relevant adhesion parameters, such as adhesion size and the generation of traction forces exerted by these structures. The particularly important role of Evl in these processes is further supported by FRAP data showing that Evl is more stably associated with other FA components compared to VASP (Damiano-Guercio et al., 2020). This is also in line with earlier work demonstrating that Evl-mediated actin polymerization regulates cell matrix adhesion, maturation of FAs and mechanosensing during durotaxis (Puleo et al., 2019). By analysis of chimeric Ena/VASP proteins, the authors of this study further proposed that the EVH1 domain of Evl is uniquely required for its function at FAs. Even though the full range of molecular functions of Ena/VASP proteins in FAs is not yet sufficiently understood, the more robust attachment of Evl to zyxin could, at least in part, explain the specialized role of this family member in adhesion.

Adherens junctions (AJs) mediated by cadherin adhesion receptors constitute important molecular assemblies that initiate and stabilize adhesion between neighboring cells. Proper regulation of these structures is crucial for various processes in tissues, including cell polarization, cytoskeletal regulation, migration, mechanotransduction and establishment of endothelial barriers (Bachir et al., 2017; Charras and Yap, 2018). Given their functional similarity, it is not surprising that FAs and AJs share many components, including Ena/VASP proteins, vinculin and zyxin (Hansen and Beckerle, 2006; Scott et al., 2006; Vasioukhin et al., 2000). Using super-resolution imaging, the highly organized protein organization within planar cadherin-based adhesions could be mapped at nanoscale resolution (Bertocchi et al., 2017). This revealed a compartmentalized architecture (distal to proximal relative to the plasma membrane), in which the cadherin–catenin compartment is located at the plasma membrane and followed by a central structure that contains vinculin connected to the innermost layer, which contains actin filaments and actin-regulatory proteins, including VASP and zyxin (Bertocchi et al., 2017). Although zyxin has also been implicated in recruiting Ena/VASP to AJs (Hansen and Beckerle, 2006; Nguyen et al., 2010), tyrosine phosphorylated (Y822) and mechanically stretched vinculin appears to act as the primary factor in recruiting VASP and Mena to cadherin junctions (Bays et al., 2014; Leerberg et al., 2014). Notably, vinculin is not tyrosine phosphorylated in cell–matrix adhesions, possibly explaining why it has little role in the recruitment of Ena/VASP to FAs (Bays et al., 2014). Interestingly, loss-of-function studies have revealed that Ena/VASP activity is necessary for tension-dependent actin polymerization in cadherin junctions of various cell types (Leerberg et al., 2014; Scott et al., 2006). Correspondingly, Drosophila Ena was previously shown to be required for actin assembly in adherens junctions within the follicle cell epithelium (Baum and Perrimon, 2001). And more recently, adhesive interactions between neighboring cells have been reported to promote compaction of small cadherin precursor assemblies into larger adhesive clusters during AJ maturation (Charras and Yap, 2018; Wu et al., 2015). It is tempting to speculate therefore that this tension-sensitive pathway may ultimately lead to Ena/VASP clustering, which, in turn, initiates junctional actin assembly, and drives actin remodeling and shapes adhesion architecture. The critical role of Ena/VASP proteins in maintaining a functional endothelium is further highlighted in mice lacking all three family members. These animals exhibit defects in acto-myosin contractility and structural integrity of endothelial barriers, leading to edema, hemorrhage and lethality (Furman et al., 2007). Nevertheless, our current knowledge regarding the specific functions of Ena/VASP proteins in these fascinating structures is far from complete and will therefore require further research.

Aside from their impact on actin remodeling based on their subcellular distributions and actin assembly activities, a precise understanding of Ena/VASP functions will also have to include a comprehensive dissection of post-translational modifications (PTMs), such as phosphorylation and indirect effects based on interactions or competition with other actin regulators. As one example, heterodimeric CP has recently turned out to promote Arp2/3-dependent actin network formation, as for instance in lamellipodia by competing off the non-productive, continuous association of WAVE proteins (e.g. WAVE1, also known as WASF1) with the growing barbed ends of filaments (Funk et al., 2021). However, CP is also a well-established antagonist of Ena/VASP family members, potentially explaining both CP and Arp2/3 complex accumulation at the periphery of Ena/VASP-deficient cells (Damiano-Guercio et al., 2020). Thus, we anticipate that comprehensive interpretation of gene removal phenotypes will increasingly have to consider such mechanistic connections between connected actin regulators. This is also of particular relevance in the context of the multiple direct interactions that have previously been described for Ena/VASP proteins with various actin regulators, most directly in this context with Arp2/3 complex activators such as WAVE or WASP (Castellano et al., 2001; Chen et al., 2014; Havrylenko et al., 2015).

VASP has also been shown to be phosphorylated on multiple sites, including both on serine/threonine sites and tyrosine residues (for details see Döppler and Storz, 2013). As an example, a tyrosine (Y39) within the EVH1 domain that is relevant for subcellular positioning (see above) has been implicated in both zyxin binding and FA targeting (Maruoka et al., 2012), but experiments with true KO cells will certainly help to quantify the precise relevance of this effect. Phosphorylation of other, further C-terminal sites, that is S239 and T278 were shown to inhibit VASP-dependent actin assembly (Benz et al., 2009; Harbeck et al., 2000), which – if occurring in FAs – was subsequently connected to the maturation of SFs and establishment of contractility (Tojkander et al., 2015). Clearly, such events will likely add to the fine tuning of activities of all family members and to the complexity of their functions, so will have to be explored comprehensively in the context of all family members and their relative abundances in different cell types and tissues.

Despite recent progress, much remains to be learned concerning the intricate specific functions of Ena/VASP family members and their coordination in distinct cell types and tissues. The availability of CRISPR/Cas9 technology will allow the generation and characterization of true KO cells and tissues, as well as enable rescue experiments with specific family members to obtain a more complete picture of their potentially distinct activities. Furthermore, it does remain puzzling why the field for so many years has believed in negative regulatory functions of Ena/VASP members in migration. Future studies should therefore also aim at providing convincing data to reconcile any previous discrepancies among the former and more recent work. For instance, such studies should validate the recently observed positive correlations between Ena/VASP activity, lamellipodium protrusion and efficiency of cell migration (Damiano-Guercio et al., 2020) in additional cell types, including the previously employed Rat2 fibroblasts (Bear et al., 2002). Finally, many open questions remain, which can now be addressed with modern technology and novel approaches, including but not restricted to the precise functions of phosphorylations, and the impact of the manifold direct and indirect interactions of Ena/VASP family members (see previous section). Elucidating the answers to all those questions will constitute major steps forward in our understanding of the fine tuning of actin remodeling in distinct cell types and tissues, and might harbor the potential to compensate for defects in actin-dependent disease.

We would like to thank Thomas Pokrant, Günther Resch, John Victor Small, Jan Müller, Maria Nemethova and Michael Sixt for providing us with unpublished TIRF and EM images as well as tomograms.

Funding

Our work in this area was supported by the Deutsche Forschungsgemeinschaft (DFG) with grants to J.F. (FA330/9-2) and K.R. (Research Training Group GRK2223).

Acevedo
,
L. A.
,
Greenwood
,
A. I.
and
Nicholson
,
L. K
. (
2017
).
A noncanonical binding site in the EVH1 domain of vasodilator-stimulated phosphoprotein regulates its interactions with the proline rich region of zyxin
.
Biochemistry
56
,
4626
-
4636
.
Applewhite
,
D. A.
,
Barzik
,
M.
,
Kojima
,
S.-I.
,
Svitkina
,
T. M.
,
Gertler
,
F. B.
and
Borisy
,
G. G.
(
2007
).
Ena/VASP proteins have an anti-capping independent function in filopodia formation
.
Mol. Biol. Cell
18
,
2579
-
2591
.
Auerbuch
,
V.
,
Loureiro
,
J. J.
,
Gertler
,
F. B.
,
Theriot
,
J. A.
and
Portnoy
,
D. A
. (
2003
).
Ena/VASP proteins contribute to Listeria monocytogenes pathogenesis by controlling temporal and spatial persistence of bacterial actin-based motility
.
Mol. Microbiol
.
49
,
1361
-
1375
.
Bachir
,
A. I.
,
Horwitz
,
A. R.
,
Nelson
,
W. J.
and
Bianchini
,
J. M
. (
2017
).
Actin-based adhesion modules mediate cell interactions with the extracellular matrix and neighboring cells
.
Cold Spring Harb. Perspect. Biol
.
9
,
a023234
.
Bachmann
,
C.
,
Fischer
,
L.
,
Walter
,
U.
and
Reinhard
,
M
. (
1999
).
The EVH2 domain of the vasodilator-stimulated phosphoprotein mediates tetramerization, F-actin binding, and actin bundle formation
.
J. Biol. Chem
.
274
,
23549
-
23557
.
Ball
,
L. J.
,
Kühne
,
R.
,
Hoffmann
,
B.
,
Häfner
,
A.
,
Schmieder
,
P.
,
Volkmer-Engert
,
R.
,
Hof
,
M.
,
Wahl
,
M.
,
Schneider-Mergener
,
J.
,
Walter
,
U.
et al. 
(
2000
).
Dual epitope recognition by the VASP EVH1 domain modulates polyproline ligand specificity and binding affinity
.
EMBO J.
19
,
4903
-
4914
.
Bashaw
,
G. J.
,
Kidd
,
T.
,
Murray
,
D.
,
Pawson
,
T.
and
Goodman
,
C. S
. (
2000
).
Repulsive axon guidance: Abelson and enabled play opposing roles downstream of the roundabout receptor
.
Cell
101
,
703
-
715
.
Baum
,
B.
and
Perrimon
,
N
. (
2001
).
Spatial control of the actin cytoskeleton in Drosophila epithelial cells
.
Nat. Cell Biol
.
3
,
883
-
890
.
Bays
,
J. L.
,
Peng
,
X.
,
Tolbert
,
C. E.
,
Guilluy
,
C.
,
Angell
,
A. E.
,
Pan
,
Y.
,
Superfine
,
R.
,
Burridge
,
K.
and
DeMali
,
K. A
. (
2014
).
Vinculin phosphorylation differentially regulates mechanotransduction at cell-cell and cell-matrix adhesions
.
J. Cell Biol
.
205
,
251
-
263
.
Bear
,
J. E.
,
Loureiro
,
J. J.
,
Libova
,
I.
,
Fässler
,
R.
,
Wehland
,
J.
and
Gertler
,
F. B
. (
2000
).
Negative regulation of fibroblast motility by Ena/VASP proteins
.
Cell
101
,
717
-
728
.
Bear
,
J. E.
,
Svitkina
,
T. M.
,
Krause
,
M.
,
Schafer
,
D. A.
,
Loureiro
,
J. J.
,
Strasser
,
G. A.
,
Maly
,
I. V.
,
Chaga
,
O. Y.
,
Cooper
,
J. A.
,
Borisy
,
G. G.
et al. 
(
2002
).
Antagonism between Ena/VASP proteins and actin filament capping regulates fibroblast motility
.
Cell
109
,
509
-
521
.
Benz
,
P. M.
,
Blume
,
C.
,
Seifert
,
S.
,
Wilhelm
,
S.
,
Waschke
,
J.
,
Schuh
,
K.
,
Gertler
,
F. B.
,
Münzel
,
T.
and
Renné
,
T
. (
2009
).
Differential VASP phosphorylation controls remodeling of the actin cytoskeleton
.
J. Cell Sci
.
122
,
3954
-
3965
.
Bertocchi
,
C.
,
Wang
,
Y.
,
Ravasio
,
A.
,
Hara
,
Y.
,
Wu
,
Y.
,
Sailov
,
T.
,
Baird
,
M. A.
,
Davidson
,
M. W.
,
Zaidel-Bar
,
R.
,
Toyama
,
Y.
et al. 
(
2017
).
Nanoscale architecture of cadherin-based cell adhesions
.
Nat. Cell Biol
.
9
,
28
-
37
.
Blanchoin
,
L.
,
Boujemaa-Paterski
,
R.
,
Sykes
,
C.
and
Plastino
,
J
. (
2014
).
Actin dynamics, architecture, and mechanics in cell motility
.
Physiol. Rev
.
94
,
235
-
263
.
Block
,
J.
,
Stradal
,
T. E. B.
,
Hänisch
,
J.
,
Geffers
,
R.
,
Köstler
,
S. A.
,
Urban
,
E.
,
Small
,
J. V.
,
Rottner
,
K.
and
Faix
,
J
. (
2008
).
Filopodia formation induced by active mDia2/Drf3
.
J. Microsc
.
231
,
506
-
517
.
Boëda
,
B.
,
Briggs
,
D. C.
,
Higgins
,
T.
,
Garvalov
,
B. K.
,
Fadden
,
A. J.
,
McDonald
,
N. Q.
and
Way
,
M
. (
2007
).
Tes, a specific Mena interacting partner, breaks the rules for EVH1 binding
.
Mol. Cell
28
,
1071
-
1082
.
Breitsprecher
,
D.
,
Kiesewetter
,
A. K.
,
Linkner
,
J.
,
Urbanke
,
C.
,
Resch
,
G. P.
,
Small
,
J. V.
and
Faix
,
J
. (
2008
).
Clustering of VASP actively drives processive, WH2 domain-mediated actin filament elongation
.
EMBO J.
27
,
2943
-
2954
.
Breitsprecher
,
D.
,
Kiesewetter
,
A. K.
,
Linkner
,
J.
,
Vinzenz
,
M.
,
Stradal
,
T. E. B.
,
Small
,
J. V.
,
Curth
,
U.
,
Dickinson
,
R. B.
and
Faix
,
J
. (
2011
).
Molecular mechanism of Ena/VASP-mediated actin-filament elongation
.
EMBO J.
30
,
456
-
467
.
Brindle
,
N. P.
,
Holt
,
M. R.
,
Davies
,
J. E.
,
Price
,
C. J.
and
Critchley
,
D. R
. (
1996
).
The focal-adhesion vasodilator-stimulated phosphoprotein (VASP) binds to the proline-rich domain in vinculin
.
Biochem. J
.
318
,
753
-
757
.
Brühmann
,
S.
,
Ushakov
,
D. S.
,
Winterhoff
,
M.
,
Dickinson
,
R. B.
,
Curth
,
U.
and
Faix
,
J
. (
2017
).
Distinct VASP tetramers synergize in the processive elongation of individual actin filaments from clustered arrays
.
Proc. Natl. Acad. Sci
.
114
,
E5815
-
E5824
.
Burridge
,
K.
and
Guilluy
,
C
. (
2016
).
Focal adhesions, stress fibers and mechanical tension
.
Exp. Cell Res
.
343
,
10
-
20
.
Castellano
,
F.
,
Le Clainche
,
C.
,
Patin
,
D.
,
Carlier
,
M. F.
and
Chavrier
,
P.
(
2001
).
A WASp-VASP complex regulates actin polymerization at the plasma membrane
.
EMBO J.
20
,
5603
-
5614
.
Chakraborty
,
T.
,
Ebel
,
F.
,
Domann
,
E.
,
Niebuhr
,
K.
,
Gerstel
,
B.
,
Pistor
,
S.
,
Temm-Grove
,
C. J.
,
Jockusch
,
B. M.
,
Reinhard
,
M.
,
Walter
,
U.
et al. 
(
1995
).
A focal adhesion factor directly linking intracellularly motile Listeria monocytogenes and Listeria ivanovii to the actin-based cytoskeleton of mammalian cells
.
EMBO J.
14
,
1314
-
1321
.
Charras
,
G. T.
and
Yap
,
A. S
. (
2018
).
Tensile forces and mechanotransduction at cell–cell junctions
.
Curr. Biol
.
28
,
R445
-
R457
.
Chen
,
X. J.
,
Squarr
,
A. J.
,
Stephan
,
R.
,
Chen
,
B.
,
Higgins
,
T. E.
,
Barry
,
D. J.
,
Martin
,
M. C.
,
Rosen
,
M. K.
,
Bogdan
,
S.
and
Way
,
M
. (
2014
).
Ena/VASP proteins cooperate with the WAVE complex to regulate the actin cytoskeleton
.
Dev. Cell
30
,
569
-
584
.
Coppolino
,
M. G.
,
Krause
,
M.
,
Hagendorff
,
P.
,
Monner
,
D. A.
,
Trimble
,
W.
,
Grinstein
,
S.
,
Wehland
,
J.
and
Sechi
,
A. S
. (
2001
).
Evidence for a molecular complex consisting of Fyb/SLAP, SLP-76, Nck, VASP and WASP that links the actin cytoskeleton to Fcγ receptor signalling during phagocytosis
.
J. Cell Sci.
114
,
4307
-
4318
.
Damiano-Guercio
,
J.
,
Kurzawa
,
L.
,
Mueller
,
J.
,
Dimchev
,
G.
,
Schaks
,
M.
,
Nemethova
,
M.
,
Pokrant
,
T.
,
Brühmann
,
S.
,
Linkner
,
J.
,
Blanchoin
,
L.
et al. 
(
2020
).
Loss of ENA/VASP interferes with lamellipodium architecture, motility and integrin-dependent adhesion
.
Elife
9
,
e55351
.
Dickinson
,
R. B.
and
Purich
,
D. L
. (
2002
).
Clamped-filament elongation model for actin-based motors
.
Biophys. J
.
82
,
605
-
617
.
Dickinson
,
R. B.
,
Caro
,
L.
and
Purich
,
D. L
. (
2004
).
Force generation by cytoskeletal filament end-tracking proteins
.
Biophys. J
.
87
,
2838
-
2854
.
Dimchev
,
G. A.
,
Steffen
,
A.
,
Kage
,
F.
,
Dimchev
,
V.
,
Pernier
,
J.
,
Carlier
,
M. F.
and
Rottner
,
K
. (
2017
).
Efficiency of lamellipodia protrusion is determined by the extent of cytosolic actin assembly
.
Mol. Biol. Cell
28
,
1311
-
1325
.
Dimchev
,
G. A.
,
Amiri
,
B.
,
Humphries
,
A. C.
,
Schaks
,
M.
,
Dimchev
,
V.
,
Stradal
,
T. E. B.
,
Faix
,
J.
,
Krause
,
M.
,
Way
,
M.
,
Falcke
,
M.
et al. 
(
2020
).
Lamellipodin tunes cell migration by stabilizing protrusions and promoting adhesion formation
.
J. Cell Sci
.
133
,
jcs239020
.
Dimchev
,
V.
,
Lahmann
,
I.
,
Koestler
,
S. A.
,
Kage
,
F.
,
Dimchev
,
G.
,
Steffen
,
A.
,
Stradal
,
T. E. B.
,
Vauti
,
F.
,
Arnold
,
H. H.
and
Rottner
,
K
. (
2021
).
Induced Arp2/3 complex depletion increases FMNL2/3 formin expression and filopodia formation
.
Front. Cell Dev. Biol
.
9
,
634708
.
Dolati
,
S.
,
Kage
,
F.
,
Mueller
,
J.
,
Müsken
,
M.
,
Kirchner
,
M.
,
Dittmar
,
G.
,
Sixt
,
M.
,
Rottner
,
K.
and
Falckea
,
M
. (
2018
).
On the relation between filament density, force generation, and protrusion rate in mesenchymal cell motility
.
Mol. Biol. Cell
29
,
2674
-
2686
.
Domann
,
E.
,
Wehland
,
J.
,
Rohde
,
M.
,
Pistor
,
S.
,
Hartl
,
M.
,
Goebel
,
W.
,
Leimeister-Wachter
,
M.
,
Wuenscher
,
M.
and
Chakraborty
,
T
. (
1992
).
A novel bacterial virulence gene in Listeria monocytogenes required for host cell microfilament interaction with homology to the proline-rich region of vinculin
.
EMBO J.
11
,
1981
-
1990
.
Dominguez
,
R
. (
2009
).
Actin filament nucleation and elongation factors structure function relationships
.
Crit. Rev. Biochem. Mol. Biol
.
44
,
351
-
366
.
Dominguez
,
R
. (
2016
).
The WH2 domain and actin nucleation: necessary but insufficient
.
Trends Biochem. Sci
.
41
,
478
-
490
.
Döppler
,
H.
and
Storz
,
P
. (
2013
).
Regulation of VASP by phosphorylation. Consequences for cell migration
.
Cell Adhes. Migr
.
7
,
482
-
486
.
Drees
,
F.
and
Gertler
,
F. B
. (
2008
).
Ena/VASP: proteins at the tip of the nervous system
.
Curr. Opin. Neurobiol
.
18
,
53
-
59
.
Drees
,
B.
,
Friederich
,
E.
,
Fradelizi
,
J.
,
Louvard
,
D.
,
Beckerle
,
M. C.
and
Golsteyn
,
R. M
. (
2000
).
Characterization of the interaction between zyxin and members of the Ena/vasodilator-stimulated phosphoprotein family of proteins
.
J. Biol. Chem
.
275
,
22503
-
22511
.
Ermekova
,
K. S.
,
Zambrano
,
N.
,
Linn
,
H.
,
Minopoli
,
G.
,
Gertler
,
F. B.
,
Russo
,
T.
and
Sudol
,
M
. (
1997
).
The WW domain of neural protein FE65 interacts with proline-rich motifs in Mena, the mammalian homolog of Drosophila enabled
.
J. Biol. Chem
.
272
,
32869
-
32877
.
Ferron
,
F.
,
Rebowski
,
G.
,
Lee
,
S. H.
and
Dominguez
,
R
. (
2007
).
Structural basis for the recruitment of profilin-actin complexes during filament elongation by Ena/VASP
.
EMBO J.
26
,
4597
-
4606
.
Fleming
,
T.
,
Chien
,
S. C.
,
Vanderzalm
,
P. J.
,
Dell
,
M.
,
Gavin
,
M. K.
,
Forrester
,
W. C.
and
Garriga
,
G
. (
2010
).
The role of C. elegans Ena/VASP homolog UNC-34 in neuronal polarity and motility
.
Dev. Biol
.
344
,
94
-
106
.
Fort
,
L.
,
Batista
,
J. M.
,
Thomason
,
P. A.
,
Spence
,
H. J.
,
Whitelaw
,
J. A.
,
Tweedy
,
L.
,
Greaves
,
J.
,
Martin
,
K. J.
,
Anderson
,
K. I.
,
Brown
,
P.
et al. 
(
2018
).
Fam49/CYRI interacts with Rac1 and locally suppresses protrusions
.
Nat. Cell Biol
.
20
,
1159
-
1171
.
Fradelizi
,
J.
,
Noireaux
,
V.
,
Plastino
,
J.
,
Menichi
,
B.
,
Louvard
,
D.
,
Sykes
,
C.
,
Golsteyn
,
R. M.
and
Friederich
,
E
. (
2001
).
ActA and human zyxin harbour Arp2/3-independent actin-polymerization activity
.
Nat. Cell Biol
.
3
,
699
-
707
.
Funk
,
J.
,
Merino
,
F.
,
Schaks
,
M.
,
Rottner
,
K.
,
Raunser
,
S.
and
Bieling
,
P
. (
2021
).
A barbed end interference mechanism reveals how capping protein promotes nucleation in branched actin networks
.
Nat. Commun
.
12
,
5329
.
Furman
,
C.
,
Sieminski
,
A. L.
,
Kwiatkowski
,
A. V.
,
Rubinson
,
D. A.
,
Vasile
,
E.
,
Bronson
,
R. T.
,
Fässler
,
R.
and
Gertler
,
F. B
. (
2007
).
Ena/VASP is required for endothelial barrier function in vivo
.
J. Cell Biol
.
179
,
761
-
775
.
Galler
,
A. B.
,
García Arguinzonis
,
M. I.
,
Baumgartner
,
W.
,
Kuhn
,
M.
,
Smolenski
,
A.
,
Simm
,
A.
and
Reinhard
,
M
. (
2006
).
VASP-dependent regulation of actin cytoskeleton rigidity, cell adhesion, and detachment
.
Histochem. Cell Biol
.
125
,
457
-
474
.
Garvalov
,
B. K.
,
Higgins
,
T. E.
,
Sutherland
,
J. D.
,
Zettl
,
M.
,
Scaplehorn
,
N.
,
Köcher
,
T.
,
Piddini
,
E.
,
Griffiths
,
G.
and
Way
,
M
. (
2003
).
The conformational state of Tes regulates its zyxin-dependent recruitment to focal adhesions
.
J. Cell Biol
.
161
,
33
-
39
.
Gateva
,
G.
,
Tojkander
,
S.
,
Koho
,
S.
,
Carpen
,
O.
and
Lappalainen
,
P
. (
2014
).
Palladin promotes assembly of non-contractile dorsal stress fibers through VASP recruitment
.
J. Cell Sci
.
127
,
1887
-
1898
,
Geese
,
M.
,
Loureiro
,
J. J.
,
Bear
,
J. E.
,
Wehland
,
J.
,
Gertler
,
F. B.
and
Sechi
,
A. S
. (
2002
).
Contribution of Ena/VASP proteins to intracellular motility of Listeria requires phosphorylation and proline-rich core but not F-actin binding or multimerization
.
Mol. Biol. Cell
13
,
2383
-
2396
.
Gertler
,
F. B.
,
Comer
,
A. R.
,
Juang
,
J. L.
,
Ahern
,
S. M.
,
Clark
,
M. J.
,
Liebl
,
E. C.
and
Hoffmann
,
F. M
. (
1995
)
. enabled, A dosage-sensitive suppressor of mutations in the Drosophila Abl tyrosine kinase, encodes an Abl substrate with SH3 domain-binding properties
.
Genes Dev.
9
,
521
-
533
.
Gertler
,
F. B.
,
Niebuhr
,
K.
,
Reinhard
,
M.
,
Wehland
,
J.
and
Soriano
,
P
. (
1996
).
Mena, a relative of VASP and Drosophila enabled, is implicated in the control of microfilament dynamics
.
Cell
87
,
227
-
239
.
Han
,
Y.-H.
,
Chung
,
C. Y.
,
Wessels
,
D.
,
Stephens
,
S.
,
Titus
,
M. A.
,
Soll
,
D. R.
and
Firtel
,
R. A
. (
2002
).
Requirement of a vasodilator-stimulated phosphoprotein family member for cell adhesion, the formation of filopodia, and chemotaxis in Dictyostelium
.
J. Biol. Chem
.
277
,
49877
-
49887
.
Hansen
,
M. D. H.
and
Beckerle
,
M. C
. (
2006
).
Opposing roles of Zyxin/LPP ACTA repeats and the LIM domain region in cell-cell adhesion
.
J. Biol. Chem
.
281
,
16178
-
16188
.
Hansen
,
S. D.
and
Mullins
,
R. D
. (
2010
).
VASP is a processive actin polymerase that requires monomeric actin for barbed end association
.
J. Cell Biol
.
191
,
571
-
584
.
Harbeck
,
B.
,
Hüttelmaier
,
S.
,
Schlüter
,
K.
,
Jockusch
,
B. M.
and
Illenberger
,
S
. (
2000
).
Phosphorylation of the vasodilator-stimulated phosphoprotein regulates its interaction with actin
.
J. Biol. Chem
.
275
,
30817
-
30825
.
Havrylenko
,
S.
,
Noguera
,
P.
,
Abou-Ghali
,
M.
,
Manzi
,
J.
,
Faqir
,
F.
,
Lamora
,
A.
,
Guérin
,
C.
,
Blanchoin
,
L.
and
Plastino
,
J
. (
2015
).
WAVE binds Ena/VASP for enhanced Arp2/3 complex-based actin assembly
.
Mol. Biol. Cell
26
,
55
-
65
.
Hoffman
,
L. M.
,
Jensen
,
C. C.
,
Kloeker
,
S.
,
Wang
,
C. L. A.
,
Yoshigi
,
M.
and
Beckerle
,
M. C
. (
2006
).
Genetic ablation of zyxin causes Mena/VASP mislocalization, increased motility, and deficits in actin remodeling
.
J. Cell Biol
.
172
,
771
-
782
.
Hotulainen
,
P.
and
Lappalainen
,
P
. (
2006
).
Stress fibers are generated by two distinct actin assembly mechanisms in motile cells
.
J. Cell Biol
.
173
,
383
-
394
.
Hüttelmaier
,
S.
,
Harbeck
,
B.
,
Steffens
,
N. O.
,
Meßerschmidt
,
T.
,
Illenberger
,
S.
and
Jockusch
,
B. M
. (
1999
).
Characterization of the actin binding properties of the vasodilator-stimulated phosphoprotein VASP
.
FEBS Lett.
451
,
68
-
74
.
Jenzora
,
A.
,
Behrendt
,
B.
,
Small
,
J. V.
,
Wehland
,
J.
and
Stradal
,
T. E. B
. (
2005
).
PREL1 provides a link from Ras signalling to the actin cytoskeleton via Ena/VASP proteins
.
FEBS Lett.
579
,
455
-
463
.
Jonckheere
,
V.
,
Lambrechts
,
A.
,
Vandekerckhove
,
J.
and
Ampe
,
C
. (
1999
).
Dimerization of profilin II upon binding the (GP5)3 peptide from VASP overcomes the inhibition of actin nucleation by profilin II and thymosin β4
.
FEBS Lett.
447
,
257
-
263
.
Kage
,
F.
,
Winterhoff
,
M.
,
Dimchev
,
V.
,
Mueller
,
J.
,
Thalheim
,
T.
,
Freise
,
A.
,
Brühmann
,
S.
,
Kollasser
,
J.
,
Block
,
J.
,
Dimchev
,
G. A.
et al. 
(
2017
).
FMNL formins boost lamellipodial force generation
.
Nat. Commun
.
8
,
14832
.
Kage
,
F.
,
Döring
,
H.
,
Mietkowska
,
M.
,
Schaks
,
M.
,
Grüner
,
F.
,
Stahnke
,
S.
,
Steffen
,
A.
,
Müsken
,
M.
,
Stradal
,
T. E. B.
and
Rottner
,
K
. (
2021
).
Lamellipodia-like actin networks in cells lacking WAVE Regulatory Complex
.
bioRxiv
,
2021.06.18.449030
.
Kaiser
,
D. A.
,
Vinson
,
V. K.
,
Murphy
,
D. B.
and
Pollard
,
T. D
. (
1999
).
Profilin is predominantly associated with monomeric actin in Acanthamoeba
.
J. Cell Sci
.
112
,
3779
-
3790
.
Kanchanawong
,
P.
,
Shtengel
,
G.
,
Pasapera
,
A. M.
,
Ramko
,
E. B.
,
Davidson
,
M. W.
,
Hess
,
H. F.
and
Waterman-Storer
,
C. M
. (
2010
).
Nanoscale architecture of integrin-based cell adhesions
.
Nature
468
,
580
-
584
.
Kang
,
N.
,
Yaqoob
,
U.
,
Geng
,
Z.
,
Bloch
,
K.
,
Liu
,
C.
,
Gomez
,
T.
,
Billadeau
,
D.
and
Shah
,
V
. (
2010
).
Focal Adhesion Assembly in Myofibroblasts Fosters a Microenvironment that Promotes Tumor Growth
.
Am. J. Pathol
.
177
,
1888
-
1900
.
Kocks
,
C.
,
Gouin
,
E.
,
Tabouret
,
M.
,
Berche
,
P.
,
Ohayon
,
H.
and
Cossart
,
P
. (
1992
).
L. monocytogenes-induced actin assembly requires the actA gene product, a surface protein
.
Cell
68
,
521
-
531
.
Kovar
,
D. R.
,
Harris
,
E. S.
,
Mahaffy
,
R.
,
Higgs
,
H. N.
and
Pollard
,
T. D
. (
2006
).
Control of the Assembly of ATP- and ADP-Actin by Formins and Profilin
.
Cell
124
,
423
-
435
.
Krause
,
M.
and
Gautreau
,
A
. (
2014
).
Steering cell migration: lamellipodium dynamics and the regulation of directional persistence
.
Nat. Rev. Mol. Cell Biol
.
15
,
577
-
590
.
Krause
,
M.
,
Dent
,
E. W.
,
Bear
,
J. E.
,
Loureiro
,
J. J.
and
Gertler
,
F. B
. (
2003
).
Ena/VASP proteins: regulators of the actin cytoskeleton and cell migration
.
Annu. Rev. Cell Dev. Biol
.
19
,
541
-
564
.
Krause
,
M.
,
Leslie
,
J. D.
,
Stewart
,
M.
,
Lafuente
,
E. M.
,
Valderrama
,
F.
,
Jagannathan
,
R.
,
Strasser
,
G. A.
,
Rubinson
,
D. A.
,
Liu
,
H.
,
Way
,
M.
et al. 
(
2004
).
Lamellipodin, an Ena/VASP ligand, is implicated in the regulation of lamellipodial dynamics
.
Dev. Cell
7
,
571
-
583
.
Krugmann
,
S.
,
Jordens
,
I.
,
Gevaert
,
K.
,
Driessens
,
M.
,
Vandekerckhove
,
J.
and
Hall
,
A
. (
2001
).
Cdc42 induces filopodia by promoting the formation of an IRSp53:Mena complex
.
Curr. Biol
.
11
,
1645
-
1655
.
Kühnel
,
K.
,
Jarchau
,
T.
,
Wolf
,
E.
,
Schlichting
,
I.
,
Walter
,
U.
,
Wittinghofer
,
A.
and
Strelkov
,
S. V
. (
2004
).
The VASP tetramerization domain is a right-handed coiled coil based on a 15-residue repeat
.
Proc. Natl. Acad. Sci. USA
101
,
17027
-
17032
.
Lafuente
,
E. M.
,
van Puijenbroek
,
A. A. F. L.
,
Krause
,
M.
,
Carman
,
C. V.
,
Freeman
,
G. J.
,
Berezovskaya
,
A.
,
Constantine
,
E.
,
Springer
,
T. A.
,
Gertler
,
F. B.
and
Boussiotis
,
V. A
. (
2004
).
RIAM, an Ena/VASP and profilin ligand, interacts with Rap1-GTP and mediates Rap1-induced adhesion
.
Dev. Cell
7
,
585
-
595
.
Lambrechts
,
A.
,
Kwiatkowski
,
A. V.
,
Lanier
,
L. M.
,
Bear
,
J. E.
,
Vandekerckhove
,
J.
,
Ampe
,
C.
and
Gertler
,
F. B
. (
2000
).
cAMP-dependent protein kinase phosphorylation of EVL, a Mena/VASP relative, regulates its interaction with actin and SH3 domains
.
J. Biol. Chem
.
275
,
36143
-
36151
.
Lanier
,
L. M.
,
Gates
,
M. A.
,
Witke
,
W.
,
Menzies
,
A. S.
,
Wehman
,
A. M.
,
Macklis
,
J. D.
,
Kwiatkowski
,
D.
,
Soriano
,
P.
and
Gertler
,
F. B
. (
1999
).
Mena is required for neurulation and commissure formation
.
Neuron
22
,
313
-
325
.
Law
,
A. L.
,
Jalal
,
S.
,
Pallett
,
T.
,
Mosis
,
F.
,
Guni
,
A.
,
Brayford
,
S.
,
Yolland
,
L.
,
Marcotti
,
S.
,
Levitt
,
J. A.
,
Poland
,
S. P.
et al. 
(
2021
).
Nance-Horan Syndrome-like 1 protein negatively regulates Scar/WAVE-Arp2/3 activity and inhibits lamellipodia stability and cell migration
.
Nat. Commun
.
12
,
5687
.
Leerberg
,
J. M.
,
Gomez
,
G. A.
,
Verma
,
S.
,
Moussa
,
E. J.
,
Wu
,
S. K.
,
Priya
,
R.
,
Hoffman
,
B. D.
,
Grashoff
,
C.
,
Schwartz
,
M. A.
and
Yap
,
A. S
. (
2014
).
Tension-sensitive actin assembly supports contractility at the epithelial zonula adherens
.
Curr. Biol
.
24
,
1689
-
1699
.
Litschko
,
C.
,
Linkner
,
J.
,
Brühmann
,
S.
,
Stradal
,
T. E. B.
,
Reinl
,
T.
,
Jänsch
,
L.
,
Rottner
,
K.
and
Faix
,
J
. (
2017
).
Differential functions of WAVE regulatory complex subunits in the regulation of actin-driven processes
.
Eur. J. Cell Biol
.
96
,
715
-
727
.
Loisel
,
T. P.
,
Boujemaa
,
R.
,
Pantaloni
,
D.
and
Carlier
,
M.-F
. (
1999
).
Reconstitution of actin-based motility of Listeria and Shigella using pure proteins
.
Nature
401
,
613
-
616
.
Loureiro
,
J. J.
,
Rubinson
,
D. A.
,
Bear
,
J. E.
,
Baltus
,
G. A.
,
Kwiatkowski
,
A. V.
and
Gertler
,
F. B
. (
2002
).
Critical roles of phosphorylation and actin binding motifs, but not the central proline-rich region, for Ena/vasodilator-stimulated phosphoprotein (VASP) function during cell migration
.
Mol. Biol. Cell
13
,
2533
-
2546
.
Maruoka
,
M.
,
Sato
,
M.
,
Yuan
,
Y.
,
Ichiba
,
M.
,
Fujii
,
R.
,
Ogawa
,
T.
,
Ishida-Kitagawa
,
N.
,
Takeya
,
T.
and
Watanabe
,
N
. (
2012
).
Abi-1-bridged tyrosine phosphorylation of VASP by Abelson kinase impairs association of VASP to focal adhesions and regulates leukaemic cell adhesion
.
Biochem. J
.
441
,
889
-
899
.
Maun
,
N. A.
,
Speicher
,
D. W.
,
DiNubile
,
M. J.
and
Southwick
,
F. S
. (
1996
).
Purification and properties of a Ca2+-independent barbed-end actin filament capping protein, CapZ, from human polymorphonuclear leukocytes
.
Biochemistry
35
,
3518
-
3524
.
Michael
,
M.
,
Vehlow
,
A.
,
Navarro
,
C.
and
Krause
,
M
. (
2010
)
. c-Abl, Lamellipodin, and Ena/VASP proteins cooperate in dorsal ruffling of fibroblasts and axonal morphogenesis
.
Curr. Biol
.
20
,
783
-
791
.
Nguyen
,
T. N.
,
Uemura
,
A.
,
Shih
,
W.
and
Yamada
,
S
. (
2010
).
Zyxin-mediated actin assembly is required for efficient wound closure
.
J. Biol. Chem
.
285
,
35439
-
35445
.
Niebuhr
,
K.
,
Ebel
,
F.
,
Frank
,
R.
,
Reinhard
,
M.
,
Domann
,
E.
,
Carl
,
U. D.
,
Walter
,
U.
,
Gertler
,
F. B.
,
Wehland
,
J.
and
Chakraborty
,
T
. (
1997
).
A novel proline-rich motif present in ActA of Listeria monocytogenes and cytoskeletal proteins is the ligand for the EVH1 domain, a protein module present in the Ena/VASP family
.
EMBO J.
16
,
5433
-
5444
.
Pasic
,
L.
,
Kotova
,
T.
and
Schafer
,
D. A
. (
2008
).
Ena/VASP proteins capture actin filament barbed ends
.
J. Biol. Chem
.
283
,
9814
-
9819
.
Paunola
,
E.
,
Mattila
,
P. K.
and
Lappalainen
,
P
. (
2002
).
WH2 domain: A small, versatile adapter for actin monomers
.
FEBS Lett.
513
,
92
-
97
.
Philippar
,
U.
,
Roussos
,
E. T.
,
Oser
,
M.
,
Yamaguchi
,
H.
,
Kim
,
H. D.
,
Giampieri
,
S.
,
Wang
,
Y.
,
Goswami
,
S.
,
Wyckoff
,
J. B.
,
Lauffenburger
,
D. A.
et al. 
(
2008
).
A Mena Invasion Isoform Potentiates EGF-Induced Carcinoma Cell Invasion and Metastasis
.
Dev. Cell
15
,
813
-
828
.
Pistor
,
S.
,
Chakraborty
,
T.
,
Niebuhr
,
K.
,
Domann
,
E.
and
Wehland
,
J
. (
1994
).
The ActA protein of Listeria monocytogenes acts as a nucleator inducing reorganization of the actin cytoskeleton
.
EMBO J.
13
,
758
-
763
.
Pistor
,
S.
,
Chakraborty
,
T.
,
Walter
,
U.
and
Wehland
,
J
. (
1995
).
The bacterial actin nucleator protein ActA of Listeria monocytogenes contains multiple binding sites for host microfilament proteins
.
Curr. Biol
.
5
,
517
-
525
.
Pollard
,
T. D.
and
Borisy
,
G. G
. (
2003
).
Cellular motility driven by assembly and disassembly of actin filaments
.
Cell
112
,
453
-
465
.
Puleo
,
J. I.
,
Parker
,
S. S.
,
Roman
,
M. R.
,
Watson
,
A. W.
,
Eliato
,
K. R.
,
Peng
,
L.
,
Saboda
,
K.
,
Roe
,
D. J.
,
Ros
,
R.
,
Gertler
,
F. B.
et al. 
(
2019
).
Mechanosensing during directed cell migration requires dynamic actin polymerization at focal adhesions
.
J. Cell Biol
.
218
,
4215
-
4235
.
Reinhard
,
M.
,
Halbrügge
,
M.
,
Scheer
,
U.
,
Wiegand
,
C.
,
Jockusch
,
B. M.
and
Walter
,
U
. (
1992
).
The 46/50 kDa phosphoprotein VASP purified from human platelets is a novel protein associated with actin filaments and focal contacts
.
EMBO J.
11
,
2063
-
2070
.
Reinhard
,
M.
,
Jouvenal
,
K.
,
Tripier
,
D.
and
Walter
,
U
. (
1995
).
Identification, purification, and characterization of a zyxin-related protein that binds the focal adhesion and microfilament protein VASP (vasodilator-stimulated phosphoprotein)
.
Proc. Natl. Acad. Sci
.
92
,
7956
-
7960
.
Roberts
,
P. J.
,
Mitin
,
N.
,
Keller
,
P. J.
,
Chenette
,
E. J.
,
Madigan
,
J. P.
,
Currin
,
R. O.
,
Cox
,
A. D.
,
Wilson
,
O.
,
Kirschmeier
,
P.
and
Der
,
C. J
. (
2008
).
Rho family GTPase modification and dependence on CAAX motif-signaled posttranslational modification
.
J. Biol. Chem
.
283
,
25150
-
25163
.
Romero
,
S.
,
Le Clainche
,
C.
and
Gautreau
,
A. M
. (
2020
).
Actin polymerization downstream of integrins: Signaling pathways and mechanotransduction
.
Biochem. J
.
477
,
1
-
21
.
Rottner
,
K.
and
Schaks
,
M
. (
2019
).
Assembling actin filaments for protrusion
.
Curr. Opin. Cell Biol
.
56
,
53
-
63
.
Rottner
,
K.
,
Behrendt
,
B.
,
Small
,
J. V.
and
Wehland
,
J
. (
1999
).
VASP dynamics during lamellipodia protrusion
.
Nat. Cell Biol
.
1
,
321
-
322
.
Rottner
,
K.
,
Krause
,
M.
,
Gimona
,
M.
,
Small
,
J. V.
and
Wehland
,
J
. (
2001
).
Zyxin is not colocalized with vasodilator-stimulated phosphoprotein (VASP) at lamellipodial tips and exhibits different dynamics to vinculin, paxillin, and vasp in focal adhesions
.
Mol. Biol. Cell
12
,
3103
-
3113
.
Rottner
,
K.
,
Faix
,
J.
,
Bogdan
,
S.
,
Linder
,
S.
and
Kerkhoff
,
E
. (
2017
).
Actin assembly mechanisms at a glance
.
J. Cell Sci
.
130
,
3427
-
3435
.
Rottner
,
K.
,
Stradal
,
T. E. B.
and
Chen
,
B
. (
2021
).
WAVE regulatory complex
.
Curr. Biol
.
31
,
R512
-
R517
.
Samarin
,
S.
,
Romero
,
S.
,
Kocks
,
C.
,
Didry
,
D.
,
Pantaloni
,
D.
and
Carlier
,
M.-F
. (
2003
).
How VASP enhances actin-based motility
.
J. Cell Biol
.
163
,
131
-
142
.
Schaks
,
M.
,
Singh
,
S. P.
,
Kage
,
F.
,
Thomason
,
P.
,
Klünemann
,
T.
,
Steffen
,
A.
,
Blankenfeldt
,
W.
,
Stradal
,
T. E.
,
Insall
,
R. H.
and
Rottner
,
K
. (
2018
).
Distinct Interaction Sites of Rac GTPase with WAVE Regulatory Complex Have Non-redundant Functions in Vivo
.
Curr. Biol
.
28
,
3674
-
3684.e6
.
Scott
,
J. A.
,
Shewan
,
A. M.
,
den Elzen
,
N. R.
,
Loureiro
,
J. J.
,
Gertler
,
F. B.
and
Yap
,
A. S.
(
2006
).
Ena/VASP proteins can regulate distinct modes of actin organization at cadherin-adhesive contacts
.
Mol. Biol. Cell
17
,
1085
-
1095
.
Sechi
,
A. S.
and
Wehland
,
J
. (
2004
).
ENA/VASP proteins: multifunctional regulators of actin cytoskeleton dynamics
.
Front. Biosci
.
9
,
1294
-
1310
.
Skoble
,
J.
,
Auerbuch
,
V.
,
Goley
,
E. D.
,
Welch
,
M. D.
and
Portnoy
,
D. A
. (
2001
).
Pivotal role of VASP in Arp2/3 complex–mediated actin nucleation, actin branch-formation, and Listeria monocytogenes motility
.
J. Cell Biol
.
155
,
89
-
100
.
Skruber
,
K.
,
Warp
,
P. V.
,
Shklyarov
,
R.
,
Thomas
,
J. D.
,
Swanson
,
M. S.
,
Henty-Ridilla
,
J. L.
,
Read
,
T. A.
and
Vitriol
,
E. A
. (
2020
).
Arp2/3 and Mena/VASP require profilin 1 for actin network assembly at the leading edge
.
Curr. Biol
.
30
,
2651
-
2664.e5
.
Small
,
J. V.
,
Stradal
,
T. E. B.
,
Vignal
,
E.
and
Rottner
,
K
. (
2002
).
The lamellipodium: where motility begins
.
Trends Cell Biol.
12
,
112
-
120
.
Smith
,
M. A.
,
Blankman
,
E.
,
Gardel
,
M. L.
,
Luettjohann
,
L.
,
Waterman
,
C. M.
and
Beckerle
,
M. C
. (
2010
).
A Zyxin-mediated mechanism for actin stress fiber maintenance and repair
.
Dev. Cell
19
,
365
-
376
.
Svitkina
,
T. M.
,
Bulanova
,
E. A.
,
Chaga
,
O. Y.
,
Vignjevic
,
D. M.
,
Kojima
,
S.
,
Vasiliev
,
J. M.
and
Borisy
,
G. G
. (
2003
).
Mechanism of filopodia initiation by reorganization of a dendritic network
.
J. Cell Biol
.
160
,
409
-
421
.
Tojkander
,
S.
,
Gateva
,
G.
,
Husain
,
A.
,
Krishnan
,
R.
and
Lappalainen
,
P
. (
2015
).
Generation of contractile actomyosin bundles depends on mechanosensitive actin filament assembly and disassembly
.
Elife
4
,
e06126
.
Trichet
,
L.
,
Sykes
,
C.
and
Plastino
,
J
. (
2008
).
Relaxing the actin cytoskeleton for adhesion and movement with Ena/VASP
.
J. Cell Biol
.
181
,
19
-
25
.
Vasioukhin
,
V.
,
Bauer
,
C.
,
Yin
,
M.
and
Fuchs
,
E
. (
2000
).
Directed actin polymerization is the driving force for epithelial cell–cell adhesion
.
Cell
100
,
209
-
219
.
Waldmann
,
R.
,
Nieberding
,
M.
and
Walter
,
U
. (
1987
).
Vasodilator-stimulated protein phosphorylation in platelets is mediated by cAMP- and cGMP-dependent protein kinases
.
Eur. J. Biochem
.
167
,
441
-
448
.
Welch
,
M. D.
and
Way
,
M
. (
2013
).
Arp2/3-mediated actin-based motility: A tail of pathogen abuse
.
Cell Host Microbe
14
,
242
-
255
.
Welch
,
M. D.
,
Rosenblatt
,
J.
,
Skoble
,
J.
,
Portnoy
,
D. A.
and
Mitchison
,
T. J
. (
1998
).
Interaction of human Arp2/3 complex and the Listeria monocytognes ActA protein in actin filament nucleation
.
Science
281
,
105
-
108
.
Winkelman
,
J. D.
,
Bilancia
,
C. G.
,
Peifer
,
M.
and
Kovar
,
D. R
. (
2014
).
Ena/VASP Enabled is a highly processive actin polymerase tailored to self-assemble parallel-bundled F-actin networks with Fascin
.
Proc. Natl. Acad. Sci. USA
111
,
4121
-
4126
.
Winograd-Katz
,
S. E.
,
Fässler
,
R.
,
Geiger
,
B.
and
Legate
,
K. R
. (
2014
).
The integrin adhesome: From genes and proteins to human disease
.
Nat. Rev. Mol. Cell Biol
.
15
,
273
-
288
.
Wu
,
Y.
,
Kanchanawong
,
P.
and
Zaidel-Bar
,
R
. (
2015
).
Actin-delimited adhesion-independent clustering of E-cadherin forms the nanoscale building blocks of adherens junctions
.
Dev. Cell
32
,
139
-
154
.
Yamashita
,
A.
,
Maeda
,
K.
and
Maéda
,
Y
. (
2003
).
Crystal structure of CapZ: Structural basis for actin filament barbed end capping
.
EMBO J.
22
,
1529
-
1538
.
Yang
,
C.
,
Czech
,
L.
,
Gerboth
,
S.
,
Kojima
,
S. I.
,
Scita
,
G.
and
Svitkina
,
T. M
. (
2007
).
Novel roles of formin mDia2 in lamellipodia and filopodia formation in motile cells
.
PLoS Biol.
5
,
e317
.
Young
,
L. E.
and
Higgs
,
H. N
. (
2018
).
Focal adhesions undergo longitudinal splitting into fixed-width units
.
Curr. Biol
.
28
,
2033
-
2045.e5
.
Zaidel-Bar
,
R.
,
Itzkovitz
,
S.
,
Ma'ayan
,
A.
,
Iyengar
,
R.
and
Geiger
,
B
. (
2007
).
Functional atlas of the integrin adhesome
.
Nat. Cell Biol
.
9
,
858
-
867
.

Competing interests

The authors declare no competing or financial interests.