The major fibronectin (FN)-binding α5β1 and αvβ3 integrins exhibit cooperativity during cell adhesion, migration and mechanosensing, through mechanisms that are not yet fully resolved. Exploiting mechanically tunable nano-patterned substrates, and peptidomimetic ligands designed to selectively bind corresponding integrins, we report that focal adhesions (FAs) of endothelial cells assembled on α5β1 integrin-selective substrates rapidly recruit αvβ3 integrins, but not vice versa. Blocking of αvβ3 integrin hindered FA maturation and cell spreading on α5β1 integrin-selective substrates, indicating a mechanism dependent on extracellular ligand binding and highlighting the requirement of αvβ3 integrin engagement for efficient adhesion. Recruitment of αvβ3 integrins additionally occurred on hydrogel substrates of varying mechanical properties, above a threshold stiffness that supports FA formation. Mechanistic studies revealed the need for soluble factors present in serum to allow recruitment, and excluded exogenous, or endogenous, FN as the ligand responsible for αvβ3 integrin accumulation to adhesion clusters. Our findings highlight a novel mechanism of integrin cooperation and a critical role for αvβ3 integrins in promoting cell adhesion on α5β1 integrin-selective substrates.

A layer of endothelial cells (ECs) constitutes the endothelium that lines the entire circulatory system, adhering to the underlying vessels or heart tissue. Normal vascular development, homeostasis and remodeling necessitates continuous and extensive crosstalk between ECs and their extracellular matrix (ECM) (Eming and Hubbell, 2011; Hahn and Schwartz, 2009). Under physiological conditions, ECs attach to a basement membrane, rich in collagen IV, laminin and proteoglycans, while injury and inflammation bring about changes in ECM composition. The presence of fibronectin (FN) in the basement membrane is a key element during the dynamic processes accompanying vascular remodeling (Hahn and Schwartz, 2009; Zhou et al., 2008). Indeed, mice lacking FN die before birth due to impaired vascular development (George et al., 1993). The presence of FN in the endothelium has also been correlated with tumorigenesis (Wang et al., 2016) and the progression of atherosclerosis (Feaver et al., 2010; Rohwedder et al., 2012).

ECs are able to both secrete FN and bind circulating plasma FN, which they can then reorganize into an extracellular fibrillar network. In its fibrillar, bioactive form, FN delivers essential mechanical and chemical polarity cues that allow endothelial cells to align correctly during vascular tubulogenesis (Hielscher et al., 2016; Zovein et al., 2010). Endothelial cell adhesion to FN is predominantly mediated by α5β1 and αvβ3 integrins, members of the integrin family of transmembrane heterodimeric adhesion receptors (Plow et al., 2014). The phenotype of mice lacking the major FN-binding α5β1 integrin (also known as the fibronectin receptor) is less severe compared to that of FN-knockout mice (Yang et al., 1993), presumably due to the ability of αv integrins to compensate for α5β1 integrin loss (Yang and Hynes, 1996). By contrast, endothelial cell-specific deletion of α5 had no obvious effects on developmental vasculogenesis, while a double knockdown of both α5 and αv integrins in ECs led to some developmental defects, which were significantly different compared to total α5 integrin- or fibronectin-null embryos (van der Flier et al., 2010). These findings call for a clearer understanding of the individual contribution and cooperation of these integrins on EC physiology.

Integrins cluster in focal adhesions (FAs) at points of contact with the ECM, providing the essential physical and chemical link necessary for cell survival, proliferation and migration (Geiger et al., 2009). Recent studies have verified FA formation under ECs in vivo with characteristics corresponding to their location in the circulatory system (van Geemen et al., 2014). While FN is mostly absent from healthy EC basal membranes, it is involved in the processes of new vessel formation (Zhou et al., 2008) and atherosclerosis (Feaver et al., 2010). α5β1 and αvβ3 integrins both bind the RGD amino acid sequence at the FN type III10 repeat; α5 integrin additionally interacts with a synergy site located in the adjacent FN type III9 repeat, making α5β1 integrin specific for FN, compared to other RGD-containing ECM proteins (Aota et al., 1994; Mould et al., 1997). Importantly, following their initial clustering, only α5β1 is able to relocate from FAs in a contractility-dependent manner to fibrillar adhesions, and mediate FN fibrillogenesis (Pankov et al., 2000; Zamir et al., 2000; Zhong et al., 1998).

A substantial degree of cooperation and crosstalk between α5β1 and αvβ3 integrins has emerged over the years. Both integrins are required for efficient mechanosensing (Schiller et al., 2013) and directional migration (Missirlis et al., 2016). However, beyond their overlapping roles and sharing FN as a common ligand, α5β1 and αvβ3 integrins exhibit distinct intracellular binding partners and signaling (Danen et al., 2002; Morgan et al., 2009). These differences manifest in altered cytoskeleton organization, force transmission and traction orientation as a function of the integrin type engaged (Balcioglu et al., 2015). In particular, studies have shown that high matrix forces are primarily supported by clustered α5β1 integrins, while less stable links to αvβ3 integrins initiate mechanotransduction (Lin et al., 2013; Rahmouni et al., 2013; Roca-Cusachs et al., 2009). The synergy between α5β1 and αvβ3 integrins thus enables the cell to sense and adapt to its mechanical microenvironment (Schiller et al., 2013), a trait that is relevant when tissues go through extensive physical remodeling during embryonic development, angiogenesis and cancer progression.

One aspect of integrin-mediated adhesion to FN expected to influence cell behavior is the temporal control of engagement for each integrin. Integrins transfer between intracellular compartments and the plasma membrane in a process known as integrin trafficking (Paul et al., 2015); interestingly, each heterodimer appears to dictate the recycling pathways of the other (Moreno-Layseca et al., 2019). Initial binding to FN seems to be mediated by α5β1 integrins, since blocking this integrin interferes with cell adhesion (Dejana et al., 1988; Schwartz and Denninghoff, 1994), and α5β1, but not αvβ3 integrins, initially recognize and bind FN (Huveneers et al., 2008; Sun et al., 2005). Following initial binding of α5β1 integrins to FN, Strohmeyer et al. (2017) found that force loading of these bonds leads to rapid αvβ3 recruitment and adhesion reinforcement; however, these data are in apparent contrast with an atomic force microscopy (AFM)-based study from the same group suggesting that αvβ3 out-competes α5β1 for initial binding to FN and is responsible for priming α5β1 for subsequent engagement (Bharadwaj et al., 2017). It is conceivable that the discrepancy arises from experimental details, or rapid switching of integrins; nevertheless, it is clear that more work is required to clarify this issue.

Overall, despite significant progress, a clear mechanistic picture of how α5β1 and αvβ3 integrins mediate distinct functions by binding the same ECM ligand is lacking. Such information is critical when designing biomaterials or drug delivery systems that target specific integrins. Many of the aforementioned studies were based on genetic manipulations, which are potentially associated with side-effects, adaptation through compensatory mechanisms or deregulation of integrin trafficking. Moreover, cells were typically presented with ECM ligands that do not enable specific and selective integrin engagement. In this study, we combined the use of engineered substrates and non-peptidic ligands (Mas-Moruno et al., 2016) to drive integrin-selective adhesion and reveal the contribution of each integrin on the same cell population. The αvβ3- and α5β1-selective integrin antagonists (peptidomimetics) we employed exhibit very potent integrin-binding affinities (IC50 values in the sub- to nanomolar range) towards the respective integrin, combined with high selectivity of at least two orders of magnitude lower affinity towards the other integrin, or αvβ5 integrin (Table S1) (Kapp et al., 2017). These antagonists have previously been used to examine the role of these integrins on cell adhesion and migration (Guasch et al., 2015; Missirlis et al., 2016; Rechenmacher et al., 2012; Schaufler et al., 2016). In particular, the α5β1 integrin peptidomimetics are unique in this respect, since selective α5β1 ligands are still lacking, despite efforts to modulate integrin affinity through mutations of FN fragments (Martino et al., 2009). The specificity and selectivity of substrates functionalized with αvβ3- and α5β1-selective peptidomimetics have been validated by showing that only cells expressing the corresponding integrin adhere to the substrates (Rechenmacher et al., 2012). Despite the high ligand specificity, FAs of ECs formed on α5β1 integrin-selective substrates stained positive for αvβ3 integrins, but not vice versa. Hence, we set out to investigate the apparent αvβ3 integrin recruitment to α5β1 integrin-based adhesions on ECs, describe its functional effects and reveal the underlying mechanism.

ECs spread efficiently on nano-patterned substrates above a threshold in ligand density

The use of nano-patterned gold substrates permits the selective presentation of ligands to cells, with additional control over ligand density and type (Arnold et al., 2004; Guasch et al., 2015). The immobilized gold nanoparticles have a size (5–10 nm) comparable to that of integrins, and the space between nanoparticles is passivated with a poly(ethylene glycol) (PEG) layer. As a result, cells can interact exclusively with immobilized ligands using one integrin per nanoparticle owing to steric properties. The threshold in ligand density for efficient cell adhesion appears to differ between cell types (Arnold et al., 2004; Muth et al., 2013); therefore, we initially examined EC spreading on quasi-hexagonal, nano-patterned substrates as a function of inter-particle spacing to determine the minimal ligand density required for efficient cell spreading (Fig. 1A; Fig. S1). Human umbilical vein ECs spread efficiently on substrates functionalized with a cyclic-RGD peptide (c-RGDfK) (Haubner et al., 1996) only when the inter-particle distance was ≤50 nm, corresponding to a threshold density of ∼400 particles/μm2 (Fig. 1B,C; Fig. S1). This value is higher compared to what has previously been reported for a fibroblast cell line (Arnold et al., 2004), but lower compared to that reported for hematopoietic cells (Muth et al., 2013), suggesting that the threshold distance between integrins for promoting efficient cell spreading is indeed cell-type specific. Of note, cells were unable to adhere to the non-patterned, PEG-coated area of the substrates, even in presence of serum, validating the specificity of the interactions with the functionalized ligands (Fig. S1). For ECs, maximal cell adhesion and projected cell area were observed on the shorter spacing examined (Fig. 1D,E), while the cell aspect ratio did not largely depend on ligand density (Fig. 1F). The high aspect ratio on substrates with inter-particle spacing of 70 nm most likely reflects high cell protrusive activity as the cell explores for permissive adhesion sites, and not effective cell polarization (Fig. 1F). Similar trends in cell adhesion and spreading as a function of ligand density were observed for ECs on nano-patterned substrates functionalized with the selective α5β1 or αvβ3 integrin antagonists (Fig. 1D–F). Despite the slightly lower efficiency of adhesion observed for ECs on integrin-selective substrates compared to c-RGDfK for the 30 nm inter-particle spacing (Fig. 1D), cell spreading and morphology were similar (Fig. 1E,F). Taking into account these results, we selected nano-patterned substrates with inter-particle spacing of 30 nm to investigate EC adhesion as a function of specific integrin engagement.

Fig. 1.

Impact of ligand type and density on endothelial cell adhesion. (A) SEM images of nano-patterned glass substrates with varying gold nanoparticle densities. The nominal inter-particle distances are shown; the measured inter-particle distance are presented in Fig. S1. (B) Epifluorescence microscopy images of phalloidin-stained ECs 4 h after seeding on nano-patterned substrates functionalized with c-RGDfK in complete medium. ECs showed efficient cell adhesion and spreading on substrates with 30 nm and 50 nm inter-particle spacing. Scale bars: 50 μm. (C) ECs cultured for 4 h in complete medium on substrates with varying inter-particle spacings and functionalized with c-RGDfK were immunostained against paxillin to visualize FAs. ECs on substrates with inter-particle distances of 30 nm and 50 nm assembled FAs, whereas paxillin fluorescence was diffuse at higher distances. Scale bars: 20 μm. (D) The number of adherent cells on substrates 4 h after seeding in complete medium, decreased with increasing inter-particle distance for all cell-adhesive ligands tested (data from three independent experiments). (E,F) Projected cell area decreased with increasing inter-particle distance (E) while aspect ratio of ECs was less dependent on ligand density (F). Data are from the indicated number of cells and from three independent experiments for 30 nm and 50 nm spacings and two independent experiments for 70 nm spacing. The middle line in box plots indicates the median, the box indicates the interquartile range and the whiskers the 5th and 95th percentiles.

Fig. 1.

Impact of ligand type and density on endothelial cell adhesion. (A) SEM images of nano-patterned glass substrates with varying gold nanoparticle densities. The nominal inter-particle distances are shown; the measured inter-particle distance are presented in Fig. S1. (B) Epifluorescence microscopy images of phalloidin-stained ECs 4 h after seeding on nano-patterned substrates functionalized with c-RGDfK in complete medium. ECs showed efficient cell adhesion and spreading on substrates with 30 nm and 50 nm inter-particle spacing. Scale bars: 50 μm. (C) ECs cultured for 4 h in complete medium on substrates with varying inter-particle spacings and functionalized with c-RGDfK were immunostained against paxillin to visualize FAs. ECs on substrates with inter-particle distances of 30 nm and 50 nm assembled FAs, whereas paxillin fluorescence was diffuse at higher distances. Scale bars: 20 μm. (D) The number of adherent cells on substrates 4 h after seeding in complete medium, decreased with increasing inter-particle distance for all cell-adhesive ligands tested (data from three independent experiments). (E,F) Projected cell area decreased with increasing inter-particle distance (E) while aspect ratio of ECs was less dependent on ligand density (F). Data are from the indicated number of cells and from three independent experiments for 30 nm and 50 nm spacings and two independent experiments for 70 nm spacing. The middle line in box plots indicates the median, the box indicates the interquartile range and the whiskers the 5th and 95th percentiles.

Focal adhesion assembly depends on the type of engaged integrin

We next examined the assembly of focal adhesions (FAs) on ECs at the initial stages of cell spreading (3 h post-seeding), when contributions from cell-secreted matrix and substrate remodeling are minimal. ECs on nano-patterned substrates with an inter-particle distance of 30 nm exhibited differences in FA morphology and distribution as a function of the presented ligands, and hence of the corresponding integrins used by ECs to adhere (Fig. 2A). Substrates presenting a c-RGDfK peptide, which primarily binds αvβ3 integrins (García et al., 2002; Petrie et al., 2006), or functionalized with αvβ3 integrin antagonists, promoted the formation of larger and more peripheral FAs, compared to substrates functionalized with α5β1 integrin-selective antagonists, on which ECs assembled smaller but more elongated adhesions, as well as central adhesions (Fig. 2A–C). On control FN-coated glass coverslips, ECs assembled a high number of FAs and fibrillar adhesions, as expected. The differences in EC adhesion cluster assembly as a function of integrin specificity are consistent with previous reports (Danen et al., 2002; Missirlis et al., 2016; Schiller et al., 2013) and indicate that ECs can distinguish the presented ligands on the integrin-selective substrates.

Fig. 2.

Focal adhesion morphology depends on the type of cell adhesive ligands presented on the substrate. (A) Inverted epifluorescence microscopy images of paxillin-immunostained ECs, 3 h after seeding in complete medium on nano-patterned substrates with 30 nm inter-particle distance and functionalized with indicated ligands or FN-coated glass coverslips. Peripheral FAs were observed on all substrates. α5β1 integrin-selective ligands and FN additionally promoted the formation of fibrillar adhesions under the cell body. Scale bars: 20 μm; 5 μm, detail panels. (B) FA area quantified from paxillin immunofluorescence staining on ECs that were allowed to spread for 3 h on nano-patterned substrates (30 nm inter-ligand spacing). FAs were similar on αvβ3- and c-RGDfK-functionalized substrates and larger than FAs assembled on α5β1-selective substrates and FN-coated glass. (C) FAs exhibited a higher aspect ratio and an elongated shape on α5β1-selective compared to αvβ3-selective substrates. Data for B and C were acquired from >10 cells and three independent experiments (n-values given on graph). The middle line in box plots indicates the median, the box indicates the interquartile range, the whiskers the 5th and 95th percentiles and the black, dashed line the mean. A Kruskal–Wallis test was used for statistical analysis (only significant differences shown): *P<0.05; ***P<0.001.

Fig. 2.

Focal adhesion morphology depends on the type of cell adhesive ligands presented on the substrate. (A) Inverted epifluorescence microscopy images of paxillin-immunostained ECs, 3 h after seeding in complete medium on nano-patterned substrates with 30 nm inter-particle distance and functionalized with indicated ligands or FN-coated glass coverslips. Peripheral FAs were observed on all substrates. α5β1 integrin-selective ligands and FN additionally promoted the formation of fibrillar adhesions under the cell body. Scale bars: 20 μm; 5 μm, detail panels. (B) FA area quantified from paxillin immunofluorescence staining on ECs that were allowed to spread for 3 h on nano-patterned substrates (30 nm inter-ligand spacing). FAs were similar on αvβ3- and c-RGDfK-functionalized substrates and larger than FAs assembled on α5β1-selective substrates and FN-coated glass. (C) FAs exhibited a higher aspect ratio and an elongated shape on α5β1-selective compared to αvβ3-selective substrates. Data for B and C were acquired from >10 cells and three independent experiments (n-values given on graph). The middle line in box plots indicates the median, the box indicates the interquartile range, the whiskers the 5th and 95th percentiles and the black, dashed line the mean. A Kruskal–Wallis test was used for statistical analysis (only significant differences shown): *P<0.05; ***P<0.001.

αvβ3 integrin is recruited to FAs assembled on α5β1 integrin-selective substrates

We next examined the effects of selective integrin engagement on integrin localization using the peptidomimetic integrin antagonists (Table S1). ECs express both α5β1 and αvβ3 integrins, among others (Fig. S2A). The integrin selectivity of substrates was verified by blocking studies: EC adhesion on α5β1 integrin-selective substrates was inhibited by soluble α5β1 integrin, but not αvβ3 integrin, peptidomimetics and vice versa for αvβ3 integrin-selective substrates (Fig. S2B). On control FN-coated glass substrates, EC adhesion was reduced by 15% and 60% following blocking of αvβ3 and α5β1 integrin, respectively, indicating that α5β1 is the principal receptor for initial FN binding (Fig. S2C), confirming previous reports on ECs (Dejana et al., 1988; Schwartz and Denninghoff, 1994) and fibroblasts (Missirlis et al., 2016). Blocking both integrins abolished EC adhesion on FN, confirming that αvβ3 and α5β1 integrins are the principal FN-binding receptors of ECs (Fig. S2C). Accordingly, αvβ3 and α5β1 integrins co-localized in FAs of ECs spreading on FN (Fig. S2D).

On α5β1 integrin-selective substrates, immunostaining of non-transformed ECs with an antibody against the α5 integrin subunit (clone MAB11) revealed clustering of α5β1 integrins as expected (note that α5 pairs exclusively with β1) (Fig. 3A). Surprisingly, integrin αvβ3 co-clustered in the same adhesions, indicating their recruitment, despite the absence of presented αvβ3 ligands (Fig. 3A). This result aligns with preliminary microscopic observations on transformed osteosarcoma cells showing αvβ3 integrins co-clustering with α5β1 integrins on α5β1-selective substrates (Guasch et al., 2015; Schaufler et al., 2016). Immunostaining with monoclonal antibodies against the activated form of integrin α5β1 (clones SNAKA51 and 9EG7) and αvβ3 (clone Ab62) demonstrated that integrin αvβ3 was present in α5β1-based adhesion in its active form (Fig. S2E,F). The finding that αvβ3 is in its extended (active) form further argues against the possibility that it binds a ligand (e.g. Thy1) anchored on the plasma membrane in a bent conformation (Fiore et al., 2015). The co-clustering of integrins was not observed on αvβ3 integrin-selective substrates, on which α5β1 integrin was not enriched in FAs abundant in αvβ3 integrins (Fig. 3A).

Fig. 3.

αvβ3 integrin is recruited to α5β1 integrin-based focal adhesions but not vice versa. (A) Inverted confocal microscopy images of ECs seeded on α5β1 or αvβ3 integrin-selective substrates for 3 h in complete medium, fixed and immunostained against the α5 integrin subunit (clone mAb11) and αvβ3 integrins (clone LM609). Clustering of both α5β1 and αvβ3 integrins was observed on α5β1 integrin-selective, but not on αvβ3 integrin-selective substrates. Line scans were performed along the green and orange lines shown in the detail panels; line scan intensity profiles highlight the presence or absence of integrin co-localization. Scale bars: 10 μm; 5 μm, detail panels. (B) The enrichment ratio for each integrin, defined as the ratio between its fluorescence intensity inside FAs to the intensity inside a ROI that encompasses these FAs, was used to quantify integrin recruitment (see Fig. S3 for details on the analysis). Both α5β1 and αvβ3 integrins clustered in FAs on α5β1-selective substrates as evidenced by the similar enhancement ratios, which were >1 (corresponding to no enrichment). In contrast, on αvβ3-selective substrates there was very low enrichment of α5β1 compared to αvβ3 integrins in FAs. Data from three independent experiments and at least 10 cells per experiment are presented (n is the number of ROIs analyzed). The middle line in box plots indicates the median, the box indicates the interquartile range, the whiskers the 5th and 95th percentiles. Data were compared using a non-parametric (Mann–Whitney) t-test. n.s., not significant; ****P<0.0001.

Fig. 3.

αvβ3 integrin is recruited to α5β1 integrin-based focal adhesions but not vice versa. (A) Inverted confocal microscopy images of ECs seeded on α5β1 or αvβ3 integrin-selective substrates for 3 h in complete medium, fixed and immunostained against the α5 integrin subunit (clone mAb11) and αvβ3 integrins (clone LM609). Clustering of both α5β1 and αvβ3 integrins was observed on α5β1 integrin-selective, but not on αvβ3 integrin-selective substrates. Line scans were performed along the green and orange lines shown in the detail panels; line scan intensity profiles highlight the presence or absence of integrin co-localization. Scale bars: 10 μm; 5 μm, detail panels. (B) The enrichment ratio for each integrin, defined as the ratio between its fluorescence intensity inside FAs to the intensity inside a ROI that encompasses these FAs, was used to quantify integrin recruitment (see Fig. S3 for details on the analysis). Both α5β1 and αvβ3 integrins clustered in FAs on α5β1-selective substrates as evidenced by the similar enhancement ratios, which were >1 (corresponding to no enrichment). In contrast, on αvβ3-selective substrates there was very low enrichment of α5β1 compared to αvβ3 integrins in FAs. Data from three independent experiments and at least 10 cells per experiment are presented (n is the number of ROIs analyzed). The middle line in box plots indicates the median, the box indicates the interquartile range, the whiskers the 5th and 95th percentiles. Data were compared using a non-parametric (Mann–Whitney) t-test. n.s., not significant; ****P<0.0001.

In order to validate the microscopy observations and quantify recruitment, we defined the enrichment ratio as the ratio of antibody staining intensity inside adhesion clusters to the intensity in a surrounding region of interest (for details on enrichment ratio calculation see Fig. S3). Adhesion clusters were determined from the antibody staining of the integrin corresponding to the presented ligand (e.g. adhesion clusters on α5β1-selective substrates were outlined from α5 integrin staining). An enrichment ratio value of 1.0 corresponds to similar staining inside and outside adhesion clusters, and therefore absence of integrin recruitment, while higher values correspond to integrin enrichment within clusters. The enrichment ratio of integrins to their corresponding ligands was high, between 1.5 and 2.0, as expected (Fig. 3B, solid green and orange boxes). The enrichment ratio for αvβ3 integrin on α5β1-selective substrates was similar to that for α5β1 integrins, confirming the observed recruitment (Fig. 3B, hatched green box). By contrast, the enrichment ratio for α5β1 integrins on αvβ3 integrin-selective substrates was significantly lower than that for αvβ3 integrins, and close to the value of 1.0 (Fig. 3B, hatched orange box).

The integrin-specific recruitment was additionally detected on a second, non-transformed cell type, namely on primary human dermal fibroblasts (pHDFs) (Fig. S2G). Moreover, analogous results showing αvβ3 integrin recruitment to FAs of ECs assembled on α5β1-selective substrates, but not vice versa, was observed on PEG-based, non-patterned, integrin-selective substrates (Fig. S2H,I), thus excluding a potential artifact of nano-patterned ligand presentation. Overall, our quantitative data demonstrate the recruitment of αvβ3 integrins to FAs, which are assembled following initial substrate engagement through α5β1 integrins.

Recruitment of αvβ3 integrin to α5β1 integrin-based FAs is independent of substrate mechanical properties above a stiffness threshold

Previous work implicating acto-myosin contractility in the regulation of αv and β1 integrin crosstalk (Schiller et al., 2013) led us to the hypothesis that RhoA-mediated intracellular signaling is responsible for the observed recruitment of αvβ3 integrin to α5β1 integrin-based FAs. Thus, we examined how modulation of actomyosin contractility affected the aforementioned integrin recruitment, by culturing ECs on PEG-based, nano-patterned hydrogels (Fig. S4A) (Aydin et al., 2010). Hydrogel Young's moduli were controlled in the physiologically relevant range of 1–40 kPa, similar to those reported for arterial vessels (Klein et al., 2009), by changing the concentration of the precursor PEG di-acrylate (Fig. S4B). As expected, ECs responded to substrate elasticity by altering their spread area, FA assembly and actin stress fiber organization (Fig. S4C–E). Importantly, ECs responded to substrate stiffness on both integrin-selective ligands, with a similar increase of cell area (Fig. S4D), and a more pronounced increase in FA area on αvβ3-selective substrates with increasing stiffness (Fig. S4E). In addition, an enhancement in nuclear localization of the transcriptional coactivator with a PDZ-binding motif (TAZ) on the stiffer hydrogels was noted, further validating the mechanosensing ability of ECs on the nano-patterned hydrogels (Fig. S4F,G).

Having validated the suitability of integrin-selective, nano-patterned hydrogels for EC mechanosensitivity studies, we examined the distribution of α5β1 and αvβ3 integrins as a function of substrate elasticity. On α5β1-selective 10 kPa and 40 kPa hydrogels, α5β1 and αvβ3 integrins clustered in FAs (Fig. 4A), similar to the observations on the rigid (50–100 GPa) nano-patterned glass substrates (Fig. 3). The softer 1 kPa hydrogels did not support FA formation, and α5β1 and αvβ3 integrins co-localized at the cell periphery, independently of the presented ligand (Fig. 4A). Accordingly, direct inhibition of myosin II using blebbistatin resulted in disassembly of FAs on rigid, nano-patterned glass substrates and a similar co-localization of both integrins at the periphery of ECs, independent of the presented integrin-selective ligand, and despite their more spread morphology (Fig. S5). The observed integrin distribution on soft hydrogels and blebbistatin-treated cells is reminiscent of reports on activated integrin localization at the cell edge during active cell protrusion (Galbraith et al., 2007; Kiosses et al., 2001). Quantification of integrin recruitment to FAs using the enrichment ratio parameter confirmed the microscopy observations, and did not reveal changes in the extent of recruitment of αvβ3-to-α5β1 integrin-based adhesions for the 10 kPa and 40 kPa substrates (Fig. 4B).

Fig. 4.

Recruitment of αvβ3 integrin to α5β1 integrin-based focal adhesions as a function of substrate mechanical properties. (A) Confocal microscopy images of ECs on nano-patterned hydrogels functionalized with indicated ligands and immunostained against the α5 integrin subunit (clone Mab11) and αvβ3 integrin (clone LM609), 3 h after seeding. ECs on 10 kPa and 40 kPa hydrogels exhibited αvβ3 integrin recruitment to α5β1 integrin-based focal adhesions, similarly to nano-patterned glass substrates. By contrast, ECs spreading on 1 kPa hydrogels did not form FAs and exhibited co-localization of both integrins at the cell periphery, independent of presented integrin-selective ligand. Line scans were performed along the green and orange dotted lines shown in the detail panels; line scan profiles corroborate the co-localization on soft hydrogels and α5β1-selective stiff hydrogels. Scale bars: 10 μm; 5 μm, detail panels. (B) Enrichment ratios calculated from two independent experiments and indicated number of cells (n; 1 ROI per cell). The middle line in box plots indicates the median, the box indicates the interquartile range, the whiskers the 5th and 95th percentiles. Data were compared with a non-parametric (Mann–Whitney) t-test. n.s., not significant; ****P<0.0001.

Fig. 4.

Recruitment of αvβ3 integrin to α5β1 integrin-based focal adhesions as a function of substrate mechanical properties. (A) Confocal microscopy images of ECs on nano-patterned hydrogels functionalized with indicated ligands and immunostained against the α5 integrin subunit (clone Mab11) and αvβ3 integrin (clone LM609), 3 h after seeding. ECs on 10 kPa and 40 kPa hydrogels exhibited αvβ3 integrin recruitment to α5β1 integrin-based focal adhesions, similarly to nano-patterned glass substrates. By contrast, ECs spreading on 1 kPa hydrogels did not form FAs and exhibited co-localization of both integrins at the cell periphery, independent of presented integrin-selective ligand. Line scans were performed along the green and orange dotted lines shown in the detail panels; line scan profiles corroborate the co-localization on soft hydrogels and α5β1-selective stiff hydrogels. Scale bars: 10 μm; 5 μm, detail panels. (B) Enrichment ratios calculated from two independent experiments and indicated number of cells (n; 1 ROI per cell). The middle line in box plots indicates the median, the box indicates the interquartile range, the whiskers the 5th and 95th percentiles. Data were compared with a non-parametric (Mann–Whitney) t-test. n.s., not significant; ****P<0.0001.

Overall, the above results suggest that αvβ3 integrin recruitment to α5β1 integrin-based FAs occurs under varied mechanical properties of the substrate, as long as these support FA formation. In the absence of FAs due to reduced contractility, there is no integrin segregation, with both integrins seemingly probing the substrate at cell edges.

Blocking of αvβ3 integrins impairs FA formation and efficient cell spreading on α5β1 integrin-selective substrates

The positive staining of clustered β3 integrins with the β3 integrin-specific LIBS2 antibody (LIBS, ligand-induced binding site) (Du et al., 1993; Wierzbicka-Patynowski et al., 1999) in α5β1 integrin-based FAs suggested that αvβ3 integrins are in an extended (active) conformation and hence more likely to bind an extracellular ligand. We investigated the possibility that such a hypothetical ligand accumulates at FAs and precedes αvβ3 integrin recruitment. Incubation of ECs with soluble αvβ3 antagonists in order to block αvβ3 integrin engagement did not alter adhesion efficiency on α5β1 integrin-selective substrates (Fig. 5A,B; Fig. S2B), but resulted in dramatic changes of cell morphology: cells exhibited an irregular, non-cohesive shape with multiple, long protrusions (Fig. 5A). Similar results were obtained using a blocking antibody against αv integrins (Fig. S6A). Image analysis revealed a significant reduction in EC spread area (Fig. 5C) and convexity (a measure of shape cohesion) upon αvβ3 blocking (Fig. 5D). The loss of cohesion upon αvβ3 integrin blocking on α5β1 integrin-selective substrates was reminiscent of the phenotype observed when myosin II activity is lost or inhibited (Rossier et al., 2010). Consistent with the reported localization of αvβ3 integrin at areas of high traction forces (Schiller et al., 2013) and its suggested role in mediating rigidity sensing (Roca-Cusachs et al., 2009), blocking of αvβ3 integrin could render the cells unable to sustain cell-generated tractions solely through α5β1 integrins and lead to loss of actin bridges along the cell edge. In line with such a scenario, αvβ3 integrin blocking impaired assembly of bona fide FAs, suggesting that αvβ3 integrin accumulation at initial adhesion clusters of α5β1 integrin is necessary for their maturation (Fig. 5E). Similar results were obtained for fibroblasts (Fig. S6B).

Fig. 5.

Blocking of αvβ3 integrin hinders αvβ3 integrin recruitment to α5β1 integrin-based FAs. (A) Confocal microscopy images of phalloidin-stained ECs seeded on α5β1 integrin-selective substrates for 3 h in complete medium. When αvβ3 integrin was blocked using soluble peptidomimetics (10 μΜ), ECs showed a change in cell morphology compared with the control conditions. Scale bars: 50 μm. (B) The percentage of cells adhering to α5β1-selective substrates was not significantly affected by blocking αvβ3 integrins. Mean and individual values from three independent experiments are presented (7–12 regions per substrate were analyzed). Data were compared using a Welch's t-test. (C,D) Projected cell area (C) and convexity values (D) of adherent ECs decreased upon αvβ3 integrin blocking. The irregular shape with multiple protrusions observed on ECs upon αvβ3 integrin blocking was quantified using the convexity parameter, which compares the actual perimeter of each cell with the perimeter of a convex hull fitted on the cell (see schematic in D). Mean±s.e.m. values from >2 independent experiments are presented. Data were compared using a Welch's t-test. (E) Confocal microscopy images of ECs seeded on α5β1-selective substrates, or FN-coated substrates for 3 h in complete medium, with αvβ3 integrins blocked. ECs were fixed, immunostained against activated β1 and β3 integrins and stained with phalloidin to reveal the actin cytoskeleton. Blocking of αvβ3 integrins inhibited FA and stress fiber formation on α5β1 integrin-selective substrates, but not on FN-coated substrates. Scale bars: 20 μm; 5 μm, detail panels.

Fig. 5.

Blocking of αvβ3 integrin hinders αvβ3 integrin recruitment to α5β1 integrin-based FAs. (A) Confocal microscopy images of phalloidin-stained ECs seeded on α5β1 integrin-selective substrates for 3 h in complete medium. When αvβ3 integrin was blocked using soluble peptidomimetics (10 μΜ), ECs showed a change in cell morphology compared with the control conditions. Scale bars: 50 μm. (B) The percentage of cells adhering to α5β1-selective substrates was not significantly affected by blocking αvβ3 integrins. Mean and individual values from three independent experiments are presented (7–12 regions per substrate were analyzed). Data were compared using a Welch's t-test. (C,D) Projected cell area (C) and convexity values (D) of adherent ECs decreased upon αvβ3 integrin blocking. The irregular shape with multiple protrusions observed on ECs upon αvβ3 integrin blocking was quantified using the convexity parameter, which compares the actual perimeter of each cell with the perimeter of a convex hull fitted on the cell (see schematic in D). Mean±s.e.m. values from >2 independent experiments are presented. Data were compared using a Welch's t-test. (E) Confocal microscopy images of ECs seeded on α5β1-selective substrates, or FN-coated substrates for 3 h in complete medium, with αvβ3 integrins blocked. ECs were fixed, immunostained against activated β1 and β3 integrins and stained with phalloidin to reveal the actin cytoskeleton. Blocking of αvβ3 integrins inhibited FA and stress fiber formation on α5β1 integrin-selective substrates, but not on FN-coated substrates. Scale bars: 20 μm; 5 μm, detail panels.

By contrast, blocking of αvβ3 integrins had no effect on morphology of ECs spreading on FN-coated glass, even though αvβ3 integrin blocking prevented its clustering at FAs (Fig. 5E). We propose that the synergy site adjacent to the RGD sequence present in the central cell-binding domain of FN, which is missing on the reductionist, nano-patterned α5β1 integrin-selective substrates, is at least partially responsible for the distinct cell behavior between these substrates. Indeed, α5β1 integrin–FN bond strength is decreased when the synergy site is mutated, leading to a dramatic reduction of cell adhesion strength (Friedland et al., 2009; Sun et al., 2005). Moreover, a recent report showed that the synergy site can compensate for loss of αvβ3 integrins, and vice versa (Benito-Jardón et al., 2017), supporting our finding that selective engagement of α5β1 integrin to the RGD peptidomimetics does not promote cell spreading when αvβ3 integrins are blocked.

αvβ3 integrin is recruited rapidly to α5β1 integrin-based FAs

If αvβ3 integrins are recruited to FAs on α5β1 integrin-selective substrates by binding an independently accumulated extracellular ligand, then recruitment should occur subsequent to, and not concurrent with, initial α5β1 integrin-induced formation of adhesion clusters. We therefore examined the kinetics of αvβ3 integrin recruitment to α5β1 integrin-based adhesion clusters during early EC spreading. αvβ3 integrin co-clustered with β1 integrins as early as 15 min following cell seeding to α5β1 integrin-selective substrates; after 30–60 min, mature FAs clearly exhibited integrin co-localization (Fig. 6A). The enrichment ratio for each integrin, calculated in this case based on staining of the early FA protein paxillin, showed a monotonic increase for αvβ3 integrins during the first hour after seeding, but instead reached a plateau for β1 integrins within 30 min. Moreover, the fraction of adhesion clusters that did not exhibit integrin enrichment (enrichment ratio ≤1) was higher for αvβ3 integrin at early time points (Fig. 6B).

Fig. 6.

αvβ3 integrin is rapidly recruited to α5β1 integrin-based FAs. (A) Confocal microscopy images of immunostained ECs seeded on α5β1 integrin-selective substrates in complete medium, washed and fixed at indicated time points. ECs were stained against paxillin, and β1 (clone 9EG7) and β3 (clone Ab62; LIBS2) integrins. Co-localization of both integrins was evident even at the earliest points (15 min) of adhesion cluster assembly and became more evident as FA matured over time (30 min and 60 min). (B) Enrichment ratio values were calculated for adhesion clusters, which in this case were identified using paxillin staining. Each data point corresponds to a FA and the percentage of adhesions exhibiting enrichment ratio <1.0 is noted on the graph. The enrichment ratio for β1 integrins increased faster compared to β3 integrins, and then decreased slightly after 60 min, most likely due to integrin translocation towards fibrillar adhesion. The enrichment ratio for β3 integrins increased monotonically over time. The fraction of adhesions with enrichment ratio <1.0 was higher for β3 integrins, indicating their sequential recruitment. Data were compared using one way ANOVA with Tukey's test. ns, not significant; *P<0.05; **P<0.01; ***P<0.001. (C) Confocal microscopy images of fixed and immunostained ECs seeded on α5β1 integrin-selective substrates in complete medium, before and at different time points after washout of blebbistatin (25 μM). Upon blebbistatin washout and FA assembly, β1 and β3 integrins that were localized at the cell edge begin to co-cluster, with some clusters containing β1 but not β3 integrins. (D) The enrichment ratio values for β1 and β3 integrins increase with time after blebbistatin washout. The higher fraction of adhesion clusters that do not exhibit αvβ3 integrin enrichment compared to α5β1 integrin further suggests that αvβ3 integrin are recruited subsequently to α5β1 integrin in the assembling FAs. Non-parametric t-tests were used to compare data: **P<0.01; ****P<0.0001. Scale bars: 20 μm; 5 μm, detail panels. Arrowheads indicate assembling adhesion clusters.

Fig. 6.

αvβ3 integrin is rapidly recruited to α5β1 integrin-based FAs. (A) Confocal microscopy images of immunostained ECs seeded on α5β1 integrin-selective substrates in complete medium, washed and fixed at indicated time points. ECs were stained against paxillin, and β1 (clone 9EG7) and β3 (clone Ab62; LIBS2) integrins. Co-localization of both integrins was evident even at the earliest points (15 min) of adhesion cluster assembly and became more evident as FA matured over time (30 min and 60 min). (B) Enrichment ratio values were calculated for adhesion clusters, which in this case were identified using paxillin staining. Each data point corresponds to a FA and the percentage of adhesions exhibiting enrichment ratio <1.0 is noted on the graph. The enrichment ratio for β1 integrins increased faster compared to β3 integrins, and then decreased slightly after 60 min, most likely due to integrin translocation towards fibrillar adhesion. The enrichment ratio for β3 integrins increased monotonically over time. The fraction of adhesions with enrichment ratio <1.0 was higher for β3 integrins, indicating their sequential recruitment. Data were compared using one way ANOVA with Tukey's test. ns, not significant; *P<0.05; **P<0.01; ***P<0.001. (C) Confocal microscopy images of fixed and immunostained ECs seeded on α5β1 integrin-selective substrates in complete medium, before and at different time points after washout of blebbistatin (25 μM). Upon blebbistatin washout and FA assembly, β1 and β3 integrins that were localized at the cell edge begin to co-cluster, with some clusters containing β1 but not β3 integrins. (D) The enrichment ratio values for β1 and β3 integrins increase with time after blebbistatin washout. The higher fraction of adhesion clusters that do not exhibit αvβ3 integrin enrichment compared to α5β1 integrin further suggests that αvβ3 integrin are recruited subsequently to α5β1 integrin in the assembling FAs. Non-parametric t-tests were used to compare data: **P<0.01; ****P<0.0001. Scale bars: 20 μm; 5 μm, detail panels. Arrowheads indicate assembling adhesion clusters.

Monitoring adhesion cluster formation during cell spreading did not allow the assessment of earlier time points (<15 min after seeding) since only a very small fraction of ECs firmly adhered on substrates and resisted washing. Therefore, we opted for a blebbistatin wash-out experiment: blebbistatin-treated ECs are well spread but lack FAs, which rapidly assemble upon drug removal (Shutova et al., 2012). At the earliest stages of assembly (5 min after wash-out), some adhesion clusters exhibited α5β1 integrin staining with absent, or faint αvβ3 staining; over time, αvβ3 integrins accumulated to maturing FAs (Fig. 6C). These observations are reflected in the calculated enrichment ratio values, which show a significant increase for both integrins between 5 min and 10 min after blebbistatin removal (Fig. 6D). Of note, the difference in absolute enrichment ratio values between the cell spreading and blebbistatin wash-out experiments most likely results from differences in integrin localization and availability on the cell membrane during the FA assembly process. Overall, our data indicate that αvβ3 integrins are recruited rapidly following the assembly of optically-resolved adhesion clusters by α5β1 integrins.

Recruitment of αvβ3 integrin to α5β1 integrin-based FAs requires the presence of serum but is not mediated by FN

Our results so far are consistent with a hypothetical model where initial α5β1 integrin adhesion clusters recruit an unidentified protein, which then provides binding sites for αvβ3 integrin recruitment, and leads to FA maturation and efficient cell spreading on α5β1 integrin-selective substrates. We next aimed to identify the hypothetical αvβ3 integrin binding ligand. During fibroblast spreading on FN, adhesion clusters are initially rich in α5β1 integrins and gradually recruit αvβ3 integrins (Zamir et al., 2000), which then allows centripetal α5β1 integrin translocation and fibrillar adhesion formation (Pankov et al., 2000). Since the α5β1 integrin-selective substrates do not theoretically provide ligands for αvβ3 integrin, we reasoned that cellular FN secreted by ECs at adhesion sites could potentially serve this role.

Cellular FN, containing the alternatively spliced EDA domain (EDA FN), was not present in FAs at early time points (Fig. 7A; Fig. S7A). At later time points, EDA FN could be detected in the formation of fibrillar adhesions, and was sometimes present in the proximal part of mature FAs (Fig. 7A). Accordingly, when FN synthesis was blocked using the total protein synthesis inhibitor cycloheximide, ECs were still able to spread and formed enlarged, peripheral FAs where αvβ3 and α5β1 integrins co-localized (Fig. 7B; Fig. S7A). Interestingly, cycloheximide treatment inhibited fibrillar adhesions, suggesting that cellular FN production is required for their formation (Fig. 7B; Fig. S7A). Analogous results were obtained for fibroblasts, with cycloheximide treatment blocking the translocation of α5β1 integrin from FAs to fibrillar adhesions, and resulting in more pronounced integrin co-localization compared to control conditions (Fig. S7B). Taken together, our data indicate that the accumulation of αvβ3 integrins at newly assembled FAs on α5β1 integrin-selective substrates is not dependent on binding cell-secreted FN.

Fig. 7.

Serum, but not FN or VN, is required for αvβ3 integrin recruitment to α5β1 integrin-based FAs. (A) Confocal microscopy images of ECs cultured on α5β1 integrin-selective substrates for the indicated time durations in complete medium and immunostained against indicated proteins. Early FAs (30 min) contained β1 integrins, but not cellular FN, which was mostly localized in a perinuclear region. At later time points (2 h), cellular FN localized in fibrillar adhesions and the proximal part of some FAs, but again was not present in peripheral FAs. (B) Confocal microscopy images of fixed, immunostained ECs on α5β1 integrin-selective substrates 2 h after seeding under different culture conditions: in the presence of 50 μg/ml cycloheximide, in the presence of FN-depleted serum or in the absence of serum. Cycloheximide treatment did not impair αvβ3 integrin recruitment to α5β1 integrin-based FAs, but inhibited fibrillar adhesion assembly. In absence of serum, ECs did not spread efficiently, but formed a few thin adhesions that did not stain positive for αvβ3 integrins. FN depletion did not have an effect on EC spreading or αvβ3 integrin recruitment to α5β1 integrin-based FAs. Enrichment ratio values did not reveal significant differences in integrin clustering for the different conditions permissive to FA assembly (cycloheximide treatment or absence of FN in serum). Each data point corresponds to the mean enrichment ratio value for one cell (5–20 adhesion clusters/cell). Data from two independent experiments are presented and compared using Kruskal–Wallis test. n.s., not significant. (C) Confocal microscopy images of fixed, immunostained ECs on α5β1 integrin-selective substrates cultured for indicated times in complete medium supplemented with pre-labeled soluble FN. Fluorescent FN (FFN) was captured by ECs and assembled into fibrils over time, but did not accumulate to newly formed peripheral FAs. Arrowheads indicate paxillin-positive adhesion clusters. (D) Confocal microscopy images of fixed ECs on α5β1 integrin-selective substrates cultured for 2 h in complete medium and immunostained against paxillin and vitronectin (VN). VN staining was absent from mature, peripheral FAs. Arrows indicate paxillin-positive adhesion clusters. (E) Confocal microscopy images of fixed, paxillin-stained ECs on α5β1 integrin-selective substrates cultured for 2 h in complete medium supplemented with pre-labeled soluble VN. VN was present in a perinuclear region of ECs, but not at mature FAs at the cell edge. Scale bars: 10 μm; 5 μm, detail panels.

Fig. 7.

Serum, but not FN or VN, is required for αvβ3 integrin recruitment to α5β1 integrin-based FAs. (A) Confocal microscopy images of ECs cultured on α5β1 integrin-selective substrates for the indicated time durations in complete medium and immunostained against indicated proteins. Early FAs (30 min) contained β1 integrins, but not cellular FN, which was mostly localized in a perinuclear region. At later time points (2 h), cellular FN localized in fibrillar adhesions and the proximal part of some FAs, but again was not present in peripheral FAs. (B) Confocal microscopy images of fixed, immunostained ECs on α5β1 integrin-selective substrates 2 h after seeding under different culture conditions: in the presence of 50 μg/ml cycloheximide, in the presence of FN-depleted serum or in the absence of serum. Cycloheximide treatment did not impair αvβ3 integrin recruitment to α5β1 integrin-based FAs, but inhibited fibrillar adhesion assembly. In absence of serum, ECs did not spread efficiently, but formed a few thin adhesions that did not stain positive for αvβ3 integrins. FN depletion did not have an effect on EC spreading or αvβ3 integrin recruitment to α5β1 integrin-based FAs. Enrichment ratio values did not reveal significant differences in integrin clustering for the different conditions permissive to FA assembly (cycloheximide treatment or absence of FN in serum). Each data point corresponds to the mean enrichment ratio value for one cell (5–20 adhesion clusters/cell). Data from two independent experiments are presented and compared using Kruskal–Wallis test. n.s., not significant. (C) Confocal microscopy images of fixed, immunostained ECs on α5β1 integrin-selective substrates cultured for indicated times in complete medium supplemented with pre-labeled soluble FN. Fluorescent FN (FFN) was captured by ECs and assembled into fibrils over time, but did not accumulate to newly formed peripheral FAs. Arrowheads indicate paxillin-positive adhesion clusters. (D) Confocal microscopy images of fixed ECs on α5β1 integrin-selective substrates cultured for 2 h in complete medium and immunostained against paxillin and vitronectin (VN). VN staining was absent from mature, peripheral FAs. Arrows indicate paxillin-positive adhesion clusters. (E) Confocal microscopy images of fixed, paxillin-stained ECs on α5β1 integrin-selective substrates cultured for 2 h in complete medium supplemented with pre-labeled soluble VN. VN was present in a perinuclear region of ECs, but not at mature FAs at the cell edge. Scale bars: 10 μm; 5 μm, detail panels.

Our experiments with ECs described thus far were performed in complete medium, which contains 2% serum among other components. Therefore, we next examined the possibility that αvβ3 integrin is recruited to α5β1 integrin-based FAs after recruitment of soluble FN that is pulled from the culture medium to FAs, and serves as the ligand for αvβ3 integrin. Previous studies have demonstrated that α5β1 integrins, but not αvβ3 integrins, even when activated, can bind soluble FN and incorporate it in a fibrillar matrix (Huveneers et al., 2008), and that α5β1 integrins are responsible for FN transport at the leading edge of migrating cells (Sung et al., 2011). However, we found that ECs cultured in medium containing FN-depleted serum were still able to spread on α5β1 integrin-selective substrates, assemble FAs and recruit αvβ3 integrins (Fig. 7B). Moreover, soluble fluorescent FN (FFN) added to the culture medium during incubation of ECs on α5β1 integrin-selective substrates was absent from peripheral FAs and accumulated in fibrils assembled under the cells over time, suggesting that it is not the ligand that recruits αvβ3 integrin (Fig. 7C; Fig. S7D). Accordingly, immunolabeling against plasma FN did not stain FN on newly formed peripheral FAs at the ventral side of cells (Fig. S7C). Therefore, our data exclude serum FN as the soluble factor responsible for recruiting αvβ3 integrins to α5β1 integrin-based FAs.

We next performed experiments in the absence of media supplements and serum in order to test whether some soluble factor is required for αvβ3 integrin recruitment to α5β1 integrin-based FAs. Under these conditions, ECs were unable to efficiently spread on α5β1 integrin-selective substrates and exhibited a phenotype similar to that of ECs with αvβ3 integrin blocked (Fig. 7B). Nevertheless, a small fraction of ECs was still able to assemble sparse, thin and elongated FAs, which were positive for β1, but not β3 integrins (Fig. 7B; Fig. S8A). FA assembly in this subset of ECs presumably occurred due to lower cell contractility under serum-free conditions, which imposes reduced loads on the engaged α5β1 integrins. Accordingly, blocking of αvβ3 integrins under serum-free conditions did not abolish FAs, even though ECs were still unable to spread adherently over the substrate (Fig. S8A). Addition of FN to the medium did not elicit observable changes in FA or cell morphology, further supporting our conclusion that soluble FN is not responsible for αvβ3 integrin recruitment to α5β1 integrin-based FAs (Fig. S8A). Since complete medium contains supplements and growth factors besides serum, we tested cell spreading behavior of ECs cultured in media containing combinations of the different commercial components. ECs were able to spread and assemble mature FAs only when serum was present (Fig. S8B).

Another serum factor, present at high concentrations (300–600 μg/ml) in serum, which could serve as the ligand for αvβ3 integrins is vitronectin (VN). While α5β1 integrin cannot directly interact with VN, and hence recruit it to FAs, it is possible that it does so indirectly through complex formation. For example, the urokinase plasminogen activator receptor (uPAR, also known as PLAUR) can both associate with α5β1 integrin and bind VN (Ferraris et al., 2014; Wei et al., 2005, 1996). We tested VN localization on ECs spreading on α5β1 integrin-selective substrates in supplemented media using immunofluorescence (Fig. 7D) or fluorescently labeled VN (Fig. 7E) and observed no evidence of localization at FAs. Moreover, addition of soluble VN under serum-free conditions was not able to rescue FA maturation (Fig. S7E). Thus, our data indicate that VN is not responsible for αvβ3 integrin recruitment to α5β1 integrin-based FAs.

FAK inhibition attenuates αvβ3 integrin recruitment

Serum could also exert its effects through inside-out activation of αvβ3 integrins. A recent report described the rapid reinforcement of adhesions initiated by α5β1 integrin binding to FN, through rapid αvβ3 integrin recruitment and focal adhesion kinase (FAK)-mediated signaling (Strohmeyer et al., 2017). Given that FAK activation preferentially occurs through β1 compared to β3 integrins (Costa et al., 2013; Shibue and Weinberg, 2009), we hypothesized that αvβ3 integrin recruitment to α5β1 integrin-based FAs might be triggered by FAK activation. Pharmacological inhibition of FAK activation using the specific inhibitor PF-573228 (Slack-Davis et al., 2007) did not impair FA assembly or β1 integrin enrichment in FAs, but attenuated the enrichment of αvβ3 integrins as evidenced by optical microscopy, image analysis and quantification of enrichment ratio values (Fig. 8). Thus, in spite of lower αvβ3 integrin recruitment, FAs matured to a similar extent in FAK-inhibited ECs as in control cells, suggesting that even low amounts of αvβ3 integrin are sufficient for FA maturation. Notwithstanding the inherent limitations of inhibitor specificity, these data indicate a role for FAK-mediated signaling during αvβ3 integrin recruitment to adhesion clusters.

Fig. 8.

FAK inhibition attenuates β3 integrin recruitment to α5β1 integrin-based FAs. (A) Confocal microscopy images of ECs in complete medium treated with 10 μM PF-573228, or equivalent amount of DMSO, on α5β1 integrin-selective substrates 3 h after seeding, and stained against paxillin and indicated integrins. Scale bars: 20 μm; 5 μm, detail panels. (B) FAK inhibition did not affect FA area of ECs as shown by the histogram and box plots (inset). n=689 for DMSO; n=782 for PF-573228. (C) Enrichment ratio values for β1 integrins were not affected by FAK inhibition, but decreased for β3 integrins, indicating that αvβ3 integrin recruitment to α5β1- integrin based FAs is at least partly mediated by FAK. Each data point corresponds to the mean enrichment ratio value for one cell (12–50 adhesion clusters/cell). Data from two independent experiments are presented and compared using non-parametric t-tests. n.s., not significant; ****P<0.0001.

Fig. 8.

FAK inhibition attenuates β3 integrin recruitment to α5β1 integrin-based FAs. (A) Confocal microscopy images of ECs in complete medium treated with 10 μM PF-573228, or equivalent amount of DMSO, on α5β1 integrin-selective substrates 3 h after seeding, and stained against paxillin and indicated integrins. Scale bars: 20 μm; 5 μm, detail panels. (B) FAK inhibition did not affect FA area of ECs as shown by the histogram and box plots (inset). n=689 for DMSO; n=782 for PF-573228. (C) Enrichment ratio values for β1 integrins were not affected by FAK inhibition, but decreased for β3 integrins, indicating that αvβ3 integrin recruitment to α5β1- integrin based FAs is at least partly mediated by FAK. Each data point corresponds to the mean enrichment ratio value for one cell (12–50 adhesion clusters/cell). Data from two independent experiments are presented and compared using non-parametric t-tests. n.s., not significant; ****P<0.0001.

Our study highlights the utility of a well-defined, nano-patterned platform and of highly selective peptidomimetics to decouple individual integrin contributions and reveal an intriguing integrin crosstalk process at the early stages of adhesion cluster formation, namely the recruitment of αvβ3 integrins to α5β1 integrin-based FAs. Initial α5β1 integrin binding triggers the formation of adhesion clusters, which in the presence of serum rapidly, yet successively, recruit αvβ3 integrins, leading to FA maturation and efficient cell spreading. A similar process occurs with FN, which is initially bound to cells via α5β1 integrins; however, FN is able to bind both αvβ3 and α5β1 integrins in contrast to the substrates presented in our study. In the case for FN binding, an ‘integrin switch’ mechanism has been proposed, wherein αvβ3 integrins replace α5β1 integrins for binding to FN (Cao et al., 2017). It is therefore tempting to speculate that the processes responsible for recruiting αvβ3 integrins upon α5β1 integrin engagement are ‘hardwired’ in cell physiology and serve the purpose of stabilizing FAs, while enabling the translocation of α5β1 and consequent assembly of FN fibrils by the cells.

As opposed to adhesion on FN, cells initially adhering exclusively through α5β1 integrin on α5β1 integrin-selective substrates require the presence of serum to recruit αvβ3 integrins. When αvβ3 integrins were blocked with soluble ligands, cells were unable to spread cohesively, highlighting the functional role of this integrin in FA maturation on these nano-patterned substrates. We suggest two possible, not mutually exclusive, scenarios as the most likely to explain our results.

First, a yet-to-be identified factor in serum could accumulate at adhesion clusters as they mature, and then serve as a ligand for αvβ3 integrins. This hypothesis is consistent with our results from the αvβ3 integrin-blocking experiment (Fig. 5) and studies demonstrating that αvβ3 integrins promoted firm binding and retention in FAs, mediated by their extracellular part, while α5β1 integrins were more mobile (Rossier et al., 2012; Tsunoyama et al., 2018). In this work, we have excluded some obvious candidates for this hypothetical soluble serum factor, such as FN and VN; further work is required to test whether accumulation of a soluble αvβ3 integrin ligand from serum, directly or through complex formation, drives αvβ3 integrin recruitment to maturing FAs.

A second possibility we put forward is that soluble serum factors promote αvβ3 integrin activation through intracellular signaling mechanisms, following initial engagement of α5β1 integrin. Support for this hypothesis comes from a previous study implicating inside-out activation for αvβ3 integrin recruitment to FN (Strohmeyer et al., 2017), which is consistent with our preliminary findings that FAK inhibition attenuates αvβ3 integrin recruitment to FAs (Fig. 8). Importantly, a recent study demonstrated that a constitutively active αvβ3 integrin mutant that cannot bind ligands is nevertheless robustly recruited to adhesion clusters (Changede et al., 2019). In our study, it is unlikely that αvβ3 integrin remains unligated at FAs, since competition with a soluble ligand impairs adhesion formation (Fig. 5), but its accumulation to integrin clusters might be triggered by its inside-out activation. At these adhesion clusters, besides the possibility of binding a ligand recruited from serum as discussed above, it is conceivable that α5β1 integrin-mediated, inside-out activation of αvβ3 integrins and its high local concentration due to recruitment combine to eventually enable αvβ3 integrin binding to the α5β1 integrin-selective ligand. Indeed, the affinity of other integrin family members (α5β1, α4β1) towards the ligand increases hundreds- to thousands-fold upon integrin conformational changes occurring during their activation (Li and Springer, 2018; Li et al., 2017). Of note, however, is that αvβ3 integrin in absence of α5β1 integrin-mediated binding does not support cell adhesion on α5β1 integrin-selective substrates (Fig. S2B) (Rechenmacher et al., 2012). Future studies utilizing β3 integrin mutants that cannot bind RGD ligands (e.g. D119Y) or that constitutively activate the integrin (e.g. N305T) (Cluzel et al., 2005), or their combination, as subsequently described in double mutants (Changede et al., 2019), should shed more light on the mechanism of β3 integrin recruitment.

Notably, the type of crosstalk between α5β1 and αvβ3 integrins we have described is not bi-directional and any proposed mechanism must be consistent with the fact that αvβ3 integrin-based FAs do not recruit α5β1 integrin FAs, and are able to sustain FA assembly on their own. Intriguingly, a recent study (Kalappurakkal et al., 2019) showed that formation of plasma membrane nanodomains occurred exclusively on FN-coated substrates, via β1 but not αv integrins, supporting our findings, and raising the possibility that αvβ3 integrin recruitment is associated with changes in the local lipid microenvironment. Such accumulation of αvβ3 integrins could occur through diffusion at the plasma membrane (Rossier et al., 2012) or directed exocytosis at FAs (Huet-Calderwood et al., 2017).

Examined under the intense interest for selective integrin targeting to limit or induce angiogenesis depending on the context (e.g. tumor angiogenesis versus tissue engineering scaffold angiogenesis) (Mas-Moruno et al., 2010), our results point to the need to consider such crosstalk in intervention design. The initial steps of cell anchoring to a substrate, whether it be their ECM or an engineered biomaterial or scaffold, can dictate long-term cell behavior, even in the presence of the ensuing active remodeling of the microenvironment. Viewed in this respect, our data indicate that blocking αvβ3 integrins could have a destabilizing effect on EC adhesion and argue for the need to move towards a more global, mechanistic understanding of how specific integrins regulate EC adhesion.

Reagents and antibodies

A list of commercially available chemicals and antibodies used in this study can be found in Tables S2 and S3. A cyclic RGDfK peptide (c-RGDfK) functionalized with a thiol group, and integrin-selective, peptidomimetic ligands against α5β1 and ανβ3 integrins were synthesized as previously described (Kapp et al., 2017; Rechenmacher et al., 2012) and their structures are presented in Table S1. Fluorescent FN was prepared using plasma FN (Life Technologies, 33010018) and an Alexa Fluor 488 (Thermo Fisher Scientific, A10235) or Alexa Fluor 568 labeling kit (Thermo Fisher Scientific, A10238). Fluorescent vitronectin was prepared using human serum vitronectin (Life Technologies, PHE0011) and an Alexa Fluor 568 labeling kit (Thermo Fisher Scientific, A10238). Poly(ethylene glycol) diacrylate with molecular weight of 10,000 was synthesized as previously described (Aydin et al., 2010). FN-depleted serum was prepared as previously described (Missirlis et al., 2017). The absence of lower molecular weight FN fragments in FFN solution or FN-depleted serum was confirmed using western blot analysis (Fig. S7F–I). Polystyrene-block-poly(2-vinylpiridine) (PS-b-P2VP) block copolymers of different molecular weights were used as per the manufacturer's instructions (Polymer Source, Canada).

Nano-patterned substrates

Nano-patterned gold particles on glass coverslips were prepared using block copolymer micellar lithography (BCML) as previously described (Glass et al., 2003). In brief, PS-b-P2VP were dissolved in toluene and assembled into uniform inverse micelles. HAuCl4 was added to the micellar suspension and partitioned in the micelle cores after magnetic stirring for >1 day. A monolayer of micelles was then deposited on piranha solution-treated [3:1 (concentrated H2SO4):(30% H2O2)] glass coverslips either by dip-coating or spin-coating of the micellar suspension. Reduction of the gold and removal of the polymer was carried out by means of hydrogen plasma treatment (0.4 mbar H2, 600 W, 45 min) with a P210 microwave plasma system (PVA TePla Americana, Corona, USA). The area between the nano-patterned gold particles was passivated using triethoxysilane PEG [molecular weight (MW) of 2000] as previously described (Blümmel et al., 2007). Substrates were kept in a nitrogen atmosphere and used within 1 day following passivation.

For the preparation of nano-patterned hydrogels we adapted a previously published procedure from our group (Aydin et al., 2010). Nano-patterned glass coverslips were prepared as described above, except that (1) a milder plasma treatment was used [0.4 mbar, W7 gas (93% Ar, 7% H2), 200 W, 45 min] in order to allow for transfer to the hydrogel surface, and (2) the passivation step was omitted. Nano-patterned glass substrates were treated with N,N′-bis (acryloyl)cystamine solution (2 mM in ethanol) for 1 h and washed three times in ethanol to provide anchoring to the hydrogel and ensure the transfer of the pattern. The functionalized nano-patterns served as the top surface of a ‘sandwich’ setup for hydrogel formation. APTES [(3-aminopropyl)triethoxysilane] functionalized glass coverslips served as the bottom surface to ensure coupling to the hydrogel. The middle layer contained the hydrogel precursor consisting of 100 μl of PEG-DA in ultrapure water, 0.35 μl APS (10% w/v) and 0.05 µl TEMED per mg PEG-DA. The distance between the two coverslips was fixed at 1 mm using custom-made Teflon holders and the cross-linking reaction was left to proceed for 30 min. The entire ‘sandwich’ assembly was then placed in ultrapure and sterile water at 4°C for 48 h to allow for swelling of the hydrogels and detachment of the upper coverslip.

Substrate functionalization

The gold nanoparticles immobilized on the nano-patterned substrates were functionalized using thiol-gold chemistry. Thiol-modified c-RGDfK or integrin-selective antagonists presenting a thiol group were incubated with substrates at a concentration of 25–100 µM in MilliQ water for 1–2 h at room temperature. Substrates were thoroughly rinsed with MilliQ water and then with PBS to remove unbound ligands. Functionalized substrates were used for cell experiments immediately after washing. Glass coverslips coated with 10 μg/ml FN overnight at 4°C, washed and blocked with 1% BSA served as controls.

For non-patterned substrates, glass coverslips were initially coated with silane–PEG–maleimide. 20×20 mm coverslips were activated with O2 plasma (100 W, 0.4 mbar) for 5 min and then were immediately immersed in anhydrous toluene under inert (N2) atmosphere, in which a small amount of silane–PEG–maleimide was dissolved and a few drops of triethylamine added. The solution was heated at 80°C and left overnight. Coverslips were then washed with ethyl acetate and methanol, and dried with a gentle N2 flow. Thiol-containing, integrin-selective antagonists at a concentration of 25 μM in PBS were incubated with surfaces for 1 h, washed with PBS three times and used for cell experiments within 2 h.

Scanning electron microscopy

Nano-patterned structure dimensions were analyzed by scanning electron micrographs recorded on a field emission electron microscope (Zeiss Leo 1530, Zeiss SMT, Oberhocken, Germany). The inter-particle distances and the order of the hexagonal patterns were determined by analyzing the micrographs with a custom-made ImageJ plugin (available upon request). The successful transfer of the gold nano-pattern from glass to the respective hydrogel was validated by cryo-scanning electron microscopy (SEM) imaging using a Zeiss Ultra 55 field emission electron microscope. Cryo-SEM was performed under low temperature conditions (Top=−130±5°C) and low acceleration voltages of 1.0–1.5 kV due to the low conductivity of samples.

Atomic force microscopy

Nano-patterned hydrogels were mechanically characterized by indentation measurements using a Nano-Wizard III atomic force microscope (JPK Instruments AG, Germany). Cantilevers with a spherical, borosilicate glass probe (sQube) of 5 µm in diameter were used. Cantilever spring constants were determined using the thermal noise calibration method and ranged between 0.45 N/m and 0.60 N/m. Young's moduli were derived by fitting force–distance curves with a Hertz model using the free software AtomicJ (version 1.7.2) (Hermanowicz et al., 2014) and assuming a Poisson ratio of 0.5. Indentation measurements at three different locations per sample, with two samples per experiment and three independent experiments were analyzed.

Cell culture

Primary human umbilical vein endothelial cells (HUVEC), pooled from different donors, were purchased from Promocell (C-12203) and cultured according to the instructions provided as sub-confluent monolayers at 37°C and 5% CO2. HUVEC were cultured in supplemented endothelial cell growth medium (Promocell, C-22010), which consists of endothelial cell basal medium (Promocell, C-22110), supplemented with 2% fetal calf serum, 0.4% endothelial cell growth supplement, 0.1 ng/ml epidermal growth factor, 1 ng/ml basic fibroblast growth factor, 90 μg/ml heparin and 1 μg/ml hydrocortisone, and which we refer to as complete medium. Cells were detached using the Detach kit (Promocell, C-41200). HUVEC were used until passage 8.

Primary human dermal fibroblasts (pHDF) were purchased from ATCC (PCS-201-010) and cultured as sub-confluent monolayers at 37°C and 5% CO2, in Dulbecco's modified Eagle's medium (DMEM; Life Technologies, 10938), supplemented with 10% fetal bovine serum (FBS) and 1% penicillin/streptomycin. pHDF were used until passage 12. HUVEC and pHDF cultures were routinely checked for the absence of mycoplasma.

Flow cytometry

Surface integrin expression of HUVEC was analyzed by flow cytometry. HUVEC (5×105 cells) were suspended using trypsin and left to recover in serum-free medium for 10 min at 37°C and 5% CO2, prior to incubation with monoclonal antibodies (100 μl at a concentration of 10 μg/ml) against α5β1 (clone JBS5) or ανβ3 (LM609) for 30 min at 4°C. Cells were washed twice with serum-free medium and incubated for another 30 min with Alexa-Fluor 674-conjugated anti-mouse secondary antibody (100 μl; 5 μg/ml). Cells incubated only with the secondary antibody served as controls. After washing twice with serum-free medium, cells were analyzed on a LSRFortessa flow cytometer (BD Biosciences).

Cell adhesion assays

The efficiency and specificity of integrin-selective substrates for cell adhesion was assessed using integrin blocking experiments. HUVEC (6.7×103 cells) were incubated in absence or in presence of integrin antagonists for 15 min on ice, prior to seeding on α5β1- or ανβ3-selective substrates in 6-well plates. After 3 h, ECs were rinsed three times with PBS to remove non-adherent cells and fixed using PFA. The number of cells was manually counted at five random spots over the substrate and cell density was expressed as percentage adhesion relative to the seeding density.

The efficiency of HUVEC adhesion on FN-coated glass was measured assessed 30 min after seeding. Briefly, suspended cells were kept under ice for 10 min, in presence or absence of integrin-selective antagonists, before being seeded in 96-well plates coated with 1 μg/ml FN. After 30 min, wells were washed twice with ice-cold PBS, the solution was aspirated and plates were placed at −80°C overnight. Relative cell numbers were quantified using the Cyquant cell proliferation assay kit (Life Technologies).

Biochemical perturbations

Cells were seeded on integrin-selective substrates for 30 min and then incubated with 25 μM blebbistatin (in DMSO) for another 1 h in order to inhibit myosin II. For FAK inhibition experiments, cells were incubated 10 min prior to and during cell spreading with 10 μM PF-573228. DMSO-treated cells served as controls in both cases. For the blebbistatin wash-out experiment, cells were washed once with PBS, fresh medium added and cells were fixed after predefined time points. For the αvβ3 integrin blocking experiment, cells were incubated with the αvβ3 ligand for 15 min under ice before cell seeding. For serum-free experiments, non-supplemented endothelium cell basal medium (Promocell, C-22210) was used. To block endogenous protein synthesis, cells were incubated with 50 μg/ml cycloheximide 1 h before seeding and also during spreading on substrates.

Immunofluorescence microscopy

HUVEC or pHDF cells (1×104 cells/cm2) cultured on substrates under indicated conditions and for predetermined time periods were washed with PBS and fixed with 4% paraformaldehyde (PFA) for 15 min at room temperature. Membranes were permeabilized with Triton X-100 (0.1%) followed by blocking with BSA (3% in PBS). Primary antibodies (in 1% BSA) were incubated for 1–2 h at room temperature or overnight at 4°C, and Alexa Fluor-labeled secondary antibodies were incubated for 1 h at room temperature. DAPI and phalloidin were used to stain nuclei and filamentous actin (F-actin), respectively. Epifluorescence microscopy images were acquired using a Delta Vision (DV) system (Applied Precision Inc.) on an Olympus IX inverted microscope, equipped with a cooled CCD camera and a 60×/1.4 NA oil-immersion objective (Olympus) or a 20×/0.3 NA air objective (Olympus). Confocal microscopy images were acquired on a Zeiss LSM 880 laser scanning confocal microscope using a 63×/1.4 NA oil-immersion objective (Zeiss) or a 20×/0.8 NA objective (Zeiss).

Image analysis

Cell projected area, aspect ratios and convexity values (convex hull perimeter divided by cell perimeter) were determined through image analysis of phalloidin-stained cells using the ImageJ software. Focal adhesion morphology analysis was performed using a custom-made ImageJ plugin (Missirlis et al., 2016). Briefly, immunofluorescence images were first background-corrected using a rolling ball filter (diameter of 25 pixels) and then smoothed using a Gauss kernel with a 5-pixel radius and standard deviation of 1 pixel. Adhesions were identified as bright pixels after applying automatic threshold using Otsu's method. An area threshold of 0.4 μm2 was set to exclude small adhesions and noise. Bright spots were localized and used as binary masks for calculating sum intensities from the original images.

Statistical analysis

Statistical analyses were performed using Prism (GraphPad Inc.). Experimental data were analyzed using either unpaired t-tests or one-way ANOVA with Tukey's post-hoc analysis, unless otherwise noted. The middle line in box plots indicates the median, the box indicates the interquartile range, the whiskers the 5th and 95th percentiles and the cross the mean. Column plots represent mean±s.e.m. *P<0.05, **P<0.01, ***P<0.001, ****P<0.0001; n.s., not significant.

We thank Joachim P. Spatz at the MPI for Medical Research, Heidelberg, for useful discussions, for hosting this project and providing free access to the facilities in his lab.

Author contributions

Conceptualization: C.D., D.M.; Methodology: C.D., D.M.; Formal analysis: C.D., D.M.; Investigation: C.D., D.M.; Resources: S.N., F.R., H.K., D.M.; Writing - original draft: C.D., D.M.; Writing - review & editing: C.D., H.K., D.M.; Visualization: C.D., D.M.; Supervision: D.M.; Project administration: C.D., D.M.; Funding acquisition: C.D.

Funding

C.D. acknowledges financial support from a Horizon 2020 Framework Programme Marie Curie IIF fellowship (660205).

Aota
,
S.
,
Nomizu
,
M.
and
Yamada
,
K. M.
(
1994
).
The short amino acid sequence Pro-His-Ser-Arg-Asn in human fibronectin enhances cell-adhesive function
.
J. Biol. Chem.
269
,
24756
-
24761
.
Arnold
,
M.
,
Cavalcanti-Adam
,
E. A.
,
Glass
,
R.
,
Blümmel
,
J.
,
Eck
,
W.
,
Kantlehner
,
M.
,
Kessler
,
H.
and
Spatz
,
J. P.
(
2004
).
Activation of integrin function by nanopatterned adhesive interfaces
.
Chemphyschem
5
,
383
-
388
.
Aydin
,
D.
,
Louban
,
I.
,
Perschmann
,
N.
,
Blümmel
,
J.
,
Lohmüller
,
T.
,
Cavalcanti-Adam
,
E. A.
,
Haas
,
T. L.
,
Walczak
,
H.
,
Kessler
,
H.
,
Fiammengo
,
R.
, et al. 
(
2010
).
Polymeric substrates with tunable elasticity and nanoscopically controlled biomolecule presentation
.
Langmuir
26
,
15472
-
15480
.
Balcioglu
,
H. E.
,
van Hoorn
,
H.
,
Donato
,
D. M.
,
Schmidt
,
T.
and
Danen
,
E. H. J.
(
2015
).
The integrin expression profile modulates orientation and dynamics of force transmission at cell-matrix adhesions
.
J. Cell Sci.
128
,
1316
-
1326
.
Benito-Jardón
,
M.
,
Klapproth
,
S.
,
Gimeno-LLuch
,
I.
,
Petzold
,
T.
,
Bharadwaj
,
M.
,
Müller
,
D. J.
,
Zuchtriegel
,
G.
,
Reichel
,
C. A.
and
Costell
,
M.
(
2017
).
The fibronectin synergy site re-enforces cell adhesion and mediates a crosstalk between integrin classes
.
eLife
6
,
562
.
Bharadwaj
,
M.
,
Strohmeyer
,
N.
,
Colo
,
G. P.
,
Helenius
,
J.
,
Beerenwinkel
,
N.
,
Schiller
,
H. B.
,
Fässler
,
R.
and
Müller
,
D. J.
(
2017
).
αV-class integrins exert dual roles on α5β1 integrins to strengthen adhesion to fibronectin
.
Nat. Commun.
8
,
14348
.
Blümmel
,
J.
,
Perschmann
,
N.
,
Aydin
,
D.
,
Drinjakovic
,
J.
,
Surrey
,
T.
,
López-García
,
M.
,
Kessler
,
H.
and
Spatz
,
J. P.
(
2007
).
Protein repellent properties of covalently attached PEG coatings on nanostructured SiO2-based interfaces
.
Biomaterials
28
,
4739
-
4747
.
Cao
,
L.
,
Nicosia
,
J.
,
Larouche
,
J.
,
Zhang
,
Y.
,
Bachman
,
H.
,
Brown
,
A. C.
,
Holmgren
,
L.
and
Barker
,
T. H.
(
2017
).
Detection of an integrin-binding mechanoswitch within fibronectin during tissue formation and fibrosis
.
ACS Nano
11
,
7110
-
7117
.
Changede
,
R.
,
Cai
,
H.
,
Wind
,
S. J.
and
Sheetz
,
M. P.
(
2019
).
Integrin nanoclusters can bridge thin matrix fibres to form cell-matrix adhesions
.
Nat. Mater.
18
,
1366
-
1375
.
Cluzel
,
C.
,
Saltel
,
F.
,
Lussi
,
J.
,
Paulhe
,
F.
,
Imhof
,
B. A.
and
Wehrle-Haller
,
B.
(
2005
).
The mechanisms and dynamics of αvβ3 integrin clustering in living cells
.
J. Cell Biol.
171
,
383
-
392
.
Costa
,
P.
,
Scales
,
T. M. E.
,
Ivaska
,
J.
and
Parsons
,
M.
(
2013
).
Integrin-specific control of focal adhesion kinase and RhoA regulates membrane protrusion and invasion
.
PLoS ONE
8
,
e74659
.
Danen
,
E. H. J.
,
Sonneveld
,
P.
,
Brakebusch
,
C.
,
Fässler
,
R.
and
Sonnenberg
,
A.
(
2002
).
The fibronectin-binding integrins α5β1 and αvβ3 differentially modulate RhoA-GTP loading, organization of cell matrix adhesions, and fibronectin fibrillogenesis
.
J. Cell Biol.
159
,
1071
-
1086
.
Dejana
,
E.
,
Colella
,
S.
,
Conforti
,
G.
,
Abbadini
,
M.
,
Gaboli
,
M.
and
Marchisio
,
P. C.
(
1988
).
Fibronectin and vitronectin regulate the organization of their respective Arg-Gly-Asp adhesion receptors in cultured human endothelial cells
.
J. Cell Biol.
107
,
1215
-
1223
.
Du
,
X.
,
Gu
,
M.
,
Weisel
,
J. W.
,
Nagaswami
,
C.
,
Bennett
,
J. S.
,
Bowditch
,
R.
and
Ginsberg
,
M. H.
(
1993
).
Long range propagation of conformational changes in integrin αIIbβ3
.
J. Biol. Chem.
268
,
23087
-
23092
.
Eming
,
S. A.
and
Hubbell
,
J. A.
(
2011
).
Extracellular matrix in angiogenesis: dynamic structures with translational potential
.
Exp. Dermatol.
20
,
605
-
613
.
Feaver
,
R. E.
,
Gelfand
,
B. D.
,
Wang
,
C.
,
Schwartz
,
M. A.
and
Blackman
,
B. R.
(
2010
).
Atheroprone hemodynamics regulate fibronectin deposition to create positive feedback that sustains endothelial inflammation
.
Circ. Res.
106
,
1703
-
1711
.
Ferraris
,
G. M. S.
,
Schulte
,
C.
,
Buttiglione
,
V.
,
De Lorenzi
,
V.
,
Piontini
,
A.
,
Galluzzi
,
M.
,
Podestà
,
A.
,
Madsen
,
C. D.
and
Sidenius
,
N.
(
2014
).
The interaction between uPAR and vitronectin triggers ligand-independent adhesion signalling by integrins
.
EMBO J.
33
,
2458
-
2472
.
Fiore
,
V. F.
,
Strane
,
P. W.
,
Bryksin
,
A. V.
,
White
,
E. S.
,
Hagood
,
J. S.
and
Barker
,
T. H.
(
2015
).
Conformational coupling of integrin and Thy-1 regulates Fyn priming and fibroblast mechanotransduction
.
J. Cell Biol.
211
,
173
-
190
.
Friedland
,
J. C.
,
Lee
,
M. H.
and
Boettiger
,
D.
(
2009
).
Mechanically activated integrin switch controls α5β1 function
.
Science
323
,
642
-
644
.
Galbraith
,
C. G.
,
Yamada
,
K. M.
and
Galbraith
,
J. A.
(
2007
).
Polymerizing actin fibers position integrins primed to probe for adhesion sites
.
Science
315
,
992
-
995
.
García
,
A. J.
,
Schwarzbauer
,
J. E.
and
Boettiger
,
D.
(
2002
).
Distinct activation states of α5β1 integrin show differential binding to RGD and synergy domains of fibronectin
.
Biochemistry
41
,
9063
-
9069
.
Geiger
,
B.
,
Spatz
,
J. P.
and
Bershadsky
,
A. D.
(
2009
).
Environmental sensing through focal adhesions
.
Nat. Rev. Mol. Cell Biol.
10
,
21
-
33
.
George
,
E. L.
,
Georges-Labouesse
,
E. N.
,
Patel-King
,
R. S.
,
Rayburn
,
H.
and
Hynes
,
R. O.
(
1993
).
Defects in mesoderm, neural tube and vascular development in mouse embryos lacking fibronectin
.
Development
119
,
1079
-
1091
.
Glass
,
R.
,
Möller
,
M.
and
Spatz
,
J. P.
(
2003
).
Block copolymer micelle nanolithography
.
Nanotechnology
14
,
1153
-
1160
.
Guasch
,
J.
,
Conings
,
B.
,
Neubauer
,
S.
,
Rechenmacher
,
F.
,
Ende
,
K.
,
Rolli
,
C. G.
,
Kappel
,
C.
,
Schaufler
,
V.
,
Micoulet
,
A.
,
Kessler
,
H.
, et al. 
(
2015
).
Segregation versus colocalization: orthogonally functionalized binary micropatterned substrates regulate the molecular distribution in focal adhesions
.
Adv. Mater. Weinheim
27
,
3737
-
3747
.
Hahn
,
C.
and
Schwartz
,
M. A.
(
2009
).
Mechanotransduction in vascular physiology and atherogenesis
.
Nat. Rev. Mol. Cell Biol.
10
,
53
-
62
.
Haubner
,
R.
,
Gratias
,
R.
,
Diefenbach
,
B.
,
Goodman
,
S. L.
,
Jonczyk
,
A.
and
Kessler
,
H.
(
1996
).
Structural and functional aspects of RGD-containing cyclic pentapeptides as highly potent and selective integrin αvβ3 antagonists
.
J. Am. Chem. Soc.
118
,
7461
-
7472
.
Hermanowicz
,
P.
,
Sarna
,
M.
,
Burda
,
K.
and
Gabryś
,
H.
(
2014
).
AtomicJ: an open source software for analysis of force curves
.
Rev. Sci. Instrum.
85
,
063703
-
9
.
Hielscher
,
A.
,
Ellis
,
K.
,
Qiu
,
C.
,
Porterfield
,
J.
and
Gerecht
,
S.
(
2016
).
Fibronectin deposition participates in extracellular matrix assembly and vascular morphogenesis
.
PLoS ONE
11
,
e0147600
-
e27
.
Huet-Calderwood
,
C.
,
Rivera-Molina
,
F.
,
Iwamoto
,
D. V.
,
Kromann
,
E. B.
,
Toomre
,
D.
and
Calderwood
,
D. A.
(
2017
).
Novel ecto-tagged integrins reveal their trafficking in live cells
.
Nat. Commun.
8
,
570
.
Huveneers
,
S.
,
Truong
,
H.
,
Fässler
,
R.
,
Sonnenberg
,
A.
and
Danen
,
E. H. J.
(
2008
).
Binding of soluble fibronectin to integrin α5β1 – link to focal adhesion redistribution and contractile shape
.
J. Cell Sci.
121
,
2452
-
2462
.
Kalappurakkal
,
J. M.
,
Anilkumar
,
A. A.
,
Patra
,
C.
,
van Zanten
,
T. S.
,
Sheetz
,
M. P.
and
Mayor
,
S.
(
2019
).
Integrin mechano-chemical signaling generates plasma membrane nanodomains that promote cell spreading
.
Cell
177
,
1738
-
1756.e23
.
Kapp
,
T. G.
,
Rechenmacher
,
F.
,
Neubauer
,
S.
,
Maltsev
,
O. V.
,
Cavalcanti-Adam
,
E. A.
,
Zarka
,
R.
,
Reuning
,
U.
,
Notni
,
J.
,
Wester
,
H.-J.
,
Mas-Moruno
,
C.
, et al. 
(
2017
).
A comprehensive evaluation of the activity and selectivity profile of ligands for RGD-binding integrins
.
Sci. Rep.
7
,
39805
.
Kiosses
,
W. B.
,
Shattil
,
S. J.
,
Pampori
,
N.
and
Schwartz
,
M. A.
(
2001
).
Rac recruits high-affinity integrin αvβ3 to lamellipodia in endothelial cell migration
.
Nat. Cell Biol.
3
,
316
-
320
.
Klein
,
E. A.
,
Yin
,
L.
,
Kothapalli
,
D.
,
Castagnino
,
P.
,
Byfield
,
F. J.
,
Xu
,
T.
,
Levental
,
I.
,
Hawthorne
,
E.
,
Janmey
,
P. A.
and
Assoian
,
R. K.
(
2009
).
Cell-cycle control by physiological matrix elasticity and in vivo tissue stiffening
.
Curr. Biol.
19
,
1511
-
1518
.
Li
,
J.
and
Springer
,
T. A.
(
2018
).
Energy landscape differences among integrins establish the framework for understanding activation
.
J. Cell Biol.
217
,
397
-
412
.
Li
,
J.
,
Su
,
Y.
,
Xia
,
W.
,
Qin
,
Y.
,
Humphries
,
M. J.
,
Vestweber
,
D.
,
Cabañas
,
C.
,
Lu
,
C.
and
Springer
,
T. A.
(
2017
).
Conformational equilibria and intrinsic affinities define integrin activation
.
EMBO J.
36
,
629
-
645
.
Lin
,
G. L.
,
Cohen
,
D. M.
,
Desai
,
R. A.
,
Breckenridge
,
M. T.
,
Gao
,
L.
,
Humphries
,
M. J.
and
Chen
,
C. S.
(
2013
).
Activation of beta 1 but not beta 3 integrin increases cell traction forces
.
FEBS Lett.
587
,
763
-
769
.
Martino
,
M. M.
,
Mochizuki
,
M.
,
Rothenfluh
,
D. A.
,
Rempel
,
S. A.
,
Hubbell
,
J. A.
and
Barker
,
T. H.
(
2009
).
Controlling integrin specificity and stem cell differentiation in 2D and 3D environments through regulation of fibronectin domain stability
.
Biomaterials
30
,
1089
-
1097
.
Mas-Moruno
,
C.
,
Rechenmacher
,
F.
and
Kessler
,
H.
(
2010
).
Cilengitide: the first anti-angiogenic small molecule drug candidate. design, synthesis and clinical evaluation
.
Anticancer Agents Med. Chem.
10
,
753
-
768
.
Mas-Moruno
,
C.
,
Fraioli
,
R.
,
Rechenmacher
,
F.
,
Neubauer
,
S.
,
Kapp
,
T. G.
and
Kessler
,
H.
(
2016
).
αvβ3- or α5β1-integrin-selective peptidomimetics for surface coating
.
Angew. Chem. Int. Ed. Engl.
55
,
7048
-
7067
.
Missirlis
,
D.
,
Haraszti
,
T.
,
Scheele
,
C. V. C.
,
Wiegand
,
T.
,
Diaz
,
C.
,
Neubauer
,
S.
,
Rechenmacher
,
F.
,
Kessler
,
H.
and
Spatz
,
J. P.
(
2016
).
Substrate engagement of integrins α5β1 and αvβ3 is necessary, but not sufficient, for high directional persistence in migration on fibronectin
.
Sci. Rep.
6
,
23258
.
Missirlis
,
D.
,
Haraszti
,
T.
,
Kessler
,
H.
and
Spatz
,
J. P.
(
2017
).
Fibronectin promotes directional persistence in fibroblast migration through interactions with both its cell-binding and heparin-binding domains
.
Sci. Rep.
7
,
3711
.
Moreno-Layseca
,
P.
,
Icha
,
J.
,
Hamidi
,
H.
and
Ivaska
,
J.
(
2019
).
Integrin trafficking in cells and tissues
.
Nat. Cell Biol.
21
,
122
-
132
.
Morgan
,
M. R.
,
Byron
,
A.
,
Humphries
,
M. J.
and
Bass
,
M. D.
(
2009
).
Giving off mixed signals–distinct functions of α5β1 and αvβ3 integrins in regulating cell behaviour
.
IUBMB Life
61
,
731
-
738
.
Mould
,
A. P.
,
Askari
,
J. A.
,
Aota
,
S.-I.
,
Yamada
,
K. M.
,
Irie
,
A.
,
Takada
,
Y.
,
Mardon
,
H. J.
and
Humphries
,
M. J.
(
1997
).
Defining the topology of integrin α5β1-fibronectin interactions using inhibitory anti-α5 and anti-β1 monoclonal antibodies. Evidence that the synergy sequence of fibronectin is recognized by the amino-terminal repeats of the α5 subunit
.
J. Biol. Chem.
272
,
17283
-
17292
.
Muth
,
C. A.
,
Steinl
,
C.
,
Klein
,
G.
and
Lee-Thedieck
,
C.
(
2013
).
Regulation of hematopoietic stem cell behavior by the nanostructured presentation of extracellular matrix components
.
PLoS ONE
8
,
e54778
.
Pankov
,
R.
,
Cukierman
,
E.
,
Katz
,
B.-Z.
,
Matsumoto
,
K.
,
Lin
,
D. C.
,
Lin
,
S.
,
Hahn
,
C.
and
Yamada
,
K. M.
(
2000
).
Integrin dynamics and matrix assembly: tensin-dependent translocation of α5β1 integrins promotes early fibronectin fibrillogenesis
.
J. Cell Biol.
148
,
1075
-
1090
.
Paul
,
N. R.
,
Jacquemet
,
G.
and
Caswell
,
P. T.
(
2015
).
Endocytic trafficking of integrins in cell migration
.
Curr. Biol.
25
,
R1092
-
R1105
.
Petrie
,
T. A.
,
Capadona
,
J. R.
,
Reyes
,
C. D.
and
García
,
A. J.
(
2006
).
Integrin specificity and enhanced cellular activities associated with surfaces presenting a recombinant fibronectin fragment compared to RGD supports
.
Biomaterials
27
,
5459
-
5470
.
Plow
,
E. F.
,
Meller
,
J.
and
Byzova
,
T. V.
(
2014
).
Integrin function in vascular biology: a view from 2013
.
Curr. Opin. Hematol.
21
,
241
-
247
.
Rahmouni
,
S.
,
Lindner
,
A.
,
Rechenmacher
,
F.
,
Neubauer
,
S.
,
Sobahi
,
T. R. A.
,
Kessler
,
H.
,
Cavalcanti-Adam
,
E. A.
and
Spatz
,
J. P.
(
2013
).
Hydrogel micropillars with integrin selective peptidomimetic functionalized nanopatterned tops: a new tool for the measurement of cell traction forces transmitted through αvβ3- or α5β1-integrins
.
Adv. Mater. Weinheim
25
,
5869
-
5874
.
Rechenmacher
,
F.
,
Neubauer
,
S.
,
Polleux
,
J.
,
Mas-Moruno
,
C.
,
De Simone
,
M.
,
Cavalcanti-Adam
,
E. A.
,
Spatz
,
J. P.
,
Fässler
,
R.
and
Kessler
,
H.
(
2012
).
Functionalizing αvβ3- or α5β1-selective integrin antagonists for surface coating: a method to discriminate integrin subtypes in vitro
.
Angew. Chem. Int. Ed.
52
,
1572
-
1575
.
Roca-Cusachs
,
P.
,
Gauthier
,
N. C.
,
del Rio
,
A.
and
Sheetz
,
M. P.
(
2009
).
Clustering of α5β1 integrins determines adhesion strength whereas αvβ3 and talin enable mechanotransduction
.
Proc. Natl. Acad. Sci. USA
106
,
16245
-
16250
.
Rohwedder
,
I.
,
Montanez
,
E.
,
Beckmann
,
K.
,
Bengtsson
,
E.
,
Dunér
,
P.
,
Nilsson
,
J.
,
Soehnlein
,
O.
and
Fässler
,
R.
(
2012
).
Plasma fibronectin deficiency impedes atherosclerosis progression and fibrous cap formation
.
EMBO Mol. Med.
4
,
564
-
576
.
Rossier
,
O. M.
,
Gauthier
,
N.
,
Biais
,
N.
,
Vonnegut
,
W.
,
Fardin
,
M.-A.
,
Avigan
,
P.
,
Heller
,
E. R.
,
Mathur
,
A.
,
Ghassemi
,
S.
,
Koeckert
,
M. S.
, et al. 
(
2010
).
Force generated by actomyosin contraction builds bridges between adhesive contacts
.
EMBO J.
29
,
1055
-
1068
.
Rossier
,
O.
,
Octeau
,
V.
,
Sibarita
,
J.-B.
,
Leduc
,
C.
,
Tessier
,
B.
,
Nair
,
D.
,
Gatterdam
,
V.
,
Destaing
,
O.
,
Albigès-Rizo
,
C.
,
Tampé
,
R.
, et al. 
(
2012
).
Integrins β1 and β3 exhibit distinct dynamic nanoscale organizations inside focal adhesions
.
Nat. Cell Biol.
14
,
1057
-
1067
.
Schaufler
,
V.
,
Czichos-Medda
,
H.
,
Hirschfeld-Warnecken
,
V.
,
Neubauer
,
S.
,
Rechenmacher
,
F.
,
Medda
,
R.
,
Kessler
,
H.
,
Geiger
,
B.
,
Spatz
,
J. P.
and
Cavalcanti-Adam
,
E. A.
(
2016
).
Selective binding and lateral clustering of α5β1 and αvβ3 integrins: unraveling the spatial requirements for cell spreading and focal adhesion assembly
.
Cell Adh. Migr.
10
,
505
-
515
.
Schiller
,
H. B.
,
Hermann
,
M.-R.
,
Polleux
,
J.
,
Vignaud
,
T.
,
Zanivan
,
S.
,
Friedel
,
C. C.
,
Sun
,
Z.
,
Raducanu
,
A.
,
Gottschalk
,
K.-E.
,
Théry
,
M.
, et al. 
(
2013
).
β1- and αv-class integrins cooperate to regulate myosin II during rigidity sensing of fibronectin-based microenvironments
.
Nat. Cell Biol.
15
,
625
-
636
.
Schwartz
,
M. A.
and
Denninghoff
,
K.
(
1994
).
Alpha v integrins mediate the rise in intracellular calcium in endothelial cells on fibronectin even though they play a minor role in adhesion
.
J. Biol. Chem.
269
,
11133
-
11137
.
Shibue
,
T.
and
Weinberg
,
R. A.
(
2009
).
Integrin β1-focal adhesion kinase signaling directs the proliferation of metastatic cancer cells disseminated in the lungs
.
Proc. Natl. Acad. Sci. USA
106
,
10290
-
10295
.
Shutova
,
M.
,
Yang
,
C.
,
Vasiliev
,
J. M.
and
Svitkina
,
T.
(
2012
).
Functions of nonmuscle myosin II in assembly of the cellular contractile system
.
PLoS ONE
7
,
e40814
.
Slack-Davis
,
J. K.
,
Martin
,
K. H.
,
Tilghman
,
R. W.
,
Iwanicki
,
M.
,
Ung
,
E. J.
,
Autry
,
C.
,
Luzzio
,
M. J.
,
Cooper
,
B.
,
Kath
,
J. C.
,
Roberts
,
W. G.
, et al. 
(
2007
).
Cellular characterization of a novel focal adhesion kinase inhibitor
.
J. Biol. Chem.
282
,
14845
-
14852
.
Strohmeyer
,
N.
,
Bharadwaj
,
M.
,
Costell
,
M.
,
Fässler
,
R.
and
Müller
,
D. J.
(
2017
).
Fibronectin-bound α5β1 integrins sense load and signal to reinforce adhesion in less than a second
.
Nat. Mater.
16
,
1262
-
1270
.
Sun
,
Z.
,
Martinez-Lemus
,
L. A.
,
Trache
,
A.
,
Trzeciakowski
,
J. P.
,
Davis
,
G. E.
,
Pohl
,
U.
and
Meininger
,
G. A.
(
2005
).
Mechanical properties of the interaction between fibronectin and α5β1-integrin on vascular smooth muscle cells studied using atomic force microscopy
.
Am. J. Physiol. Heart C
289
,
H2526
-
H2535
.
Sung
,
B. H.
,
Zhu
,
X.
,
Kaverina
,
I.
and
Weaver
,
A. M.
(
2011
).
Cortactin controls cell motility and lamellipodial dynamics by regulating ECM secretion
.
Curr. Biol.
21
,
1460
-
1469
.
Tsunoyama
,
T. A.
,
Watanabe
,
Y.
,
Goto
,
J.
,
Naito
,
K.
,
Kasai
,
R. S.
,
Suzuki
,
K. G. N.
,
Fujiwara
,
T. K.
and
Kusumi
,
A.
(
2018
).
Super-long single-molecule tracking reveals dynamic-anchorage-induced integrin function
.
Nat. Chem. Biol.
14
,
497
-
506
.
van der Flier
,
A.
,
Badu-Nkansah
,
K.
,
Whittaker
,
C. A.
,
Crowley
,
D.
,
Bronson
,
R. T.
,
Lacy-Hulbert
,
A.
and
Hynes
,
R. O.
(
2010
).
Endothelial α5 and αv integrins cooperate in remodeling of the vasculature during development
.
Development
137
,
2439
-
2449
.
van Geemen
,
D.
,
Smeets
,
M. W. J.
,
van Stalborch
,
A.-M. D.
,
Woerdeman
,
L. A. E.
,
Daemen
,
M. J. A. P.
,
Hordijk
,
P. L.
and
Huveneers
,
S.
(
2014
).
F-actin-anchored focal adhesions distinguish endothelial phenotypes of human arteries and veins
.
Arterioscler. Thromb. Vasc. Biol.
34
,
2059
-
2067
.
Wang
,
K.
,
Seo
,
B. R.
,
Fischbach
,
C.
and
Gourdon
,
D.
(
2016
).
Fibronectin mechanobiology regulates tumorigenesis
.
Cell Mol. Bioeng.
9
,
1
-
11
.
Wei
,
Y.
,
Lukashev
,
M.
,
Simon
,
D. I.
,
Bodary
,
S. C.
,
Rosenberg
,
S.
,
Doyle
,
M. V.
and
Chapman
,
H. A.
(
1996
).
Regulation of integrin function by the urokinase receptor
.
Science
273
,
1551
-
1555
.
Wei
,
Y.
,
Czekay
,
R.-P.
,
Robillard
,
L.
,
Kugler
,
M. C.
,
Zhang
,
F.
,
Kim
,
K. K.
,
Xiong
,
J.-P.
,
Humphries
,
M. J.
and
Chapman
,
H. A.
(
2005
).
Regulation of α5β1 integrin conformation and function by urokinase receptor binding
.
J. Cell Biol.
168
,
501
-
511
.
Wierzbicka-Patynowski
,
I.
,
Niewiarowski
,
S.
,
Marcinkiewicz
,
C.
,
Calvete
,
J. J.
,
Marcinkiewicz
,
M. M.
and
McLane
,
M. A.
(
1999
).
Structural requirements of echistatin for the recognition of αvβ3 and α5β1 integrins
.
J. Biol. Chem.
274
,
37809
-
37814
.
Yang
,
J. T.
and
Hynes
,
R. O.
(
1996
).
Fibronectin receptor functions in embryonic cells deficient in alpha 5 beta 1 integrin can be replaced by alpha V integrins
.
Mol. Biol. Cell
7
,
1737
-
1748
.
Yang
,
J. T.
,
Rayburn
,
H.
and
Hynes
,
R. O.
(
1993
).
Embryonic mesodermal defects in α5 integrin-deficient mice
.
Development
119
,
1093
-
1105
.
Zamir
,
E.
,
Katz
,
M.
,
Posen
,
Y.
,
Erez
,
N.
,
Yamada
,
K. M.
,
Katz
,
B.-Z.
,
Lin
,
S.
,
Lin
,
D. C.
,
Bershadsky
,
A.
,
Kam
,
Z.
, et al. 
(
2000
).
Dynamics and segregation of cell-matrix adhesions in cultured fibroblasts
.
Nat. Cell Biol.
2
,
191
-
196
.
Zhong
,
C.
,
Chrzanowska-Wodnicka
,
M.
,
Brown
,
J.
,
Shaub
,
A.
,
Belkin
,
A. M.
and
Burridge
,
K.
(
1998
).
Rho-mediated contractility exposes a cryptic site in fibronectin and induces fibronectin matrix assembly
.
J. Cell Biol.
141
,
539
-
551
.
Zhou
,
X.
,
Rowe
,
R. G.
,
Hiraoka
,
N.
,
George
,
J. P.
,
Wirtz
,
D.
,
Mosher
,
D. F.
,
Virtanen
,
I.
,
Chernousov
,
M. A.
and
Weiss
,
S. J.
(
2008
).
Fibronectin fibrillogenesis regulates three-dimensional neovessel formation
.
Genes Dev.
22
,
1231
-
1243
.
Zovein
,
A. C.
,
Luque
,
A.
,
Turlo
,
K. A.
,
Hofmann
,
J. J.
,
Yee
,
K. M.
,
Becker
,
M. S.
,
Fässler
,
R.
,
Mellman
,
I.
,
Lane
,
T. F.
and
Iruela-Arispe
,
M. L.
(
2010
).
Beta1 integrin establishes endothelial cell polarity and arteriolar lumen formation via a Par3-dependent mechanism
.
Dev. Cell
18
,
39
-
51
.

Competing interests

The authors declare no competing or financial interests.

Supplementary information