Abstract
Cell migration is a fundamental process for life and is highly dependent on the dynamical and mechanical properties of the cytoskeleton. Intensive physical and biochemical crosstalk among actin, microtubules, and intermediate filaments ensures their coordination to facilitate and enable migration. In this review, we discuss the different mechanical aspects that govern cell migration and provide, for each mechanical aspect, a novel perspective by juxtaposing two complementary approaches to the biophysical study of cytoskeletal crosstalk: live-cell studies (often referred to as top-down studies) and cell-free studies (often referred to as bottom-up studies). We summarize the main findings from both experimental approaches, and we provide our perspective on bridging the two perspectives to address the open questions of how cytoskeletal crosstalk governs cell migration and makes cells move.
I. INTRODUCTION
Cell migration is a process that is fundamental for life. It is a major contributor to tissue morphogenesis in developing embryos1 and drives angiogenesis,2 bone formation,3 tissue repair,4 and immune surveillance.5 On the flip side, however, cell migration is also responsible for pathological cell migration during chronic inflammation6 and cancer metastasis.7 Cell migration depends on the mechanical and dynamical properties of the cytoskeleton, a network of dynamic biopolymers that self-assemble from small protein building blocks. There are three main cytoskeletal biopolymers: actin filaments, microtubules, and intermediate filaments [Fig. 1(a)]. They have markedly different structural, mechanical, and dynamical properties. In addition, septins are identified as the fourth cytoskeletal component.8
Actin filaments are double helices with a diameter of nm, made of two strands of globular monomers.9 The filaments are semiflexible since their thermal persistence length (where κ is the bending rigidity and kBT is the thermal energy) is m, of the same order as the filament contour length.10 Actin filaments have an intrinsic structural polarity with a “barbed end” and a “pointed end.” Polymerization-linked ATP hydrolysis causes treadmilling, where the filaments grow at the barbed end and disassemble from the pointed end.11 Filaments reconstituted from purified actin turn over slowly (one subunit every 3–4 s), but actin turnover in the cell is catalyzed by actin-binding proteins. Typical actin network turnover times are of order seconds in the leading edge of motile cells12 to minutes in the actin cortex.13 Together with myosin motor proteins, actin filaments form networks and bundles that generate contractile forces.14 Microtubules form hollow tubes of 13 protofilaments that are much wider (∼25 nm)15 and hence substantially stiffer (lp ∼ 1 mm)16 than actin filaments. Like actin filaments, microtubules have an intrinsic structural polarity with distinct plus and minus ends. GTP hydrolysis results in dynamic instability, characterized by alternating phases of microtubule growth and shrinkage.17 In the cell, this process is tightly regulated by accessory proteins that bind at the microtubule tip or lattice. Intermediate filaments are homo-/heteropolymers made of rod-shaped proteins that are encoded by more than 70 genes in humans.18 Intermediate filament proteins are expressed in a cell-type-specific manner. Mesenchymal cells, for instance, express vimentin, whereas epithelial cells express keratins. The intermediate filament proteins share a common secondary structure consisting of an alpha-helical rod domain flanked by intrinsically disordered head and tail domains. Intermediate filaments are somewhat thicker (∼10 nm)19 than actin filaments, but they are, nevertheless, much more flexible (lp ∼ 0.5–2 μm, depending on intermediate filament composition and ionic strength20–24) because of their hierarchical rope-like structure. Intermediate filaments are much more stable than actin filaments and microtubules, with slow subunit exchange along their length and annealing and fragmentation on hour time scales in reconstituted systems25,26 and in cells.27,28
The physical properties of the cytoskeletal filaments are directly connected to their functions in cell migration. Actin, with its ability to generate protrusive and contractile forces, provides the main driving forces for polymerization-driven mesenchymal migration and bleb-based amoeboid migration.29,30 Meanwhile microtubules play a key role in establishing front-rear polarity and promoting persistent migration, aided by their large persistence length that is much longer than the size of the cell.31 Finally, intermediate filaments, with their mechanical resilience, protect the migrating cell and its nucleus from mechanical damage, which is especially important when cells squeeze through confined environments.32,33
There is growing evidence that cell migration requires a dynamic interplay between the three cytoskeletal filament systems that depends on mechanical and signaling crosstalk. In mesenchymal migration, coupling of actin to microtubules and intermediate filaments is, for instance, essential to polarize the actin cytoskeleton and control force generation.34 In Fig. 2, we highlight examples of the different cytoskeletal crosstalk modalities (Table I). Here, we review recent insights in the role of cytoskeletal crosstalk in cell migration, with a focus on mechanical aspects. For more detailed cell biological insights, we refer the reader to several excellent reviews.34–38 We take a mainly experimental perspective and refer the reader to other reviews for more theoretically oriented perspectives.39,40 Throughout this review, we confront two opposite experimental approaches to studying the biophysics of cytoskeletal crosstalk: live cell (top-down) studies [Fig. 1(b)] vs cell-free (bottom-up) studies of simplified model systems reconstituted from component parts [Fig. 1(c)]. Live-cell studies have the benefit of physiological relevance, but mechanistic dissection is challenging because of the cell's compositional complexity. Each cytoskeletal system exhibits enormous compositional diversity with different isoforms and posttranslational modifications.41 Moreover, cytoskeletal coupling is mechanosensitive as a consequence of mechanosensory signaling loops and transcriptional regulation.42 Cell-free studies provide a powerful approach to complement live-cell studies because they allow for highly controlled experiments from the level of single protein, to filaments, to networks.
TABLE I.
Modality | Cytoskeletal element #1 | Cytoskeletal element #2 | Mediator(s) | Function of the crosstalk |
---|---|---|---|---|
Direct interaction | Microtubules | Vimentin | Excluded volume | Promoted strain-stiffening135,136 |
Actin | Vimentin | Excluded volume | Mechanical synergy132 | |
Crosslinker | Actin | Vimentin | Plectin | Force transmission to nucleus146 |
Microtubule | Actin | ACF7/MACF | Guided filament growth140,189 | |
Molecular signaling | Actin | Microtubules | GEF-H1/RhoA | Detachment of leader cells228 |
Actin | Vimentin | RhoGTPases | Inhibition of stress fiber assembly155 | |
Steric interaction | Actin | Microtubules | Electrostatic interactions | Promoting filament alignment140 |
Cytoskeletal crosstalk contributes to every aspect of cell migration (Fig. 2). We structure the review according to these aspects, from cell deformability, to front-back polarity, contractility, adhesion control in collective cell migration, and finally plasticity, which refers to the ability of cells to adapt their mode of migration to their environment.43 We end with a perspective on how connections can be made between cell-free and live-cell studies to address the many open questions on the role of cytoskeletal crosstalk in cell migration.
II. CELL DEFORMABILITY
A. Live-cell studies
1. Mechanical challenges in cell migration
Migrating cells must deform their nucleus and cytoskeleton, especially when they move through interstitial tissues that impose significant confinement. Depending on tissue type, cells encounter extracellular matrix (ECM) and interstices between tissues with sizes ranging between 2 and 30 μm, comparable to their own body and sometimes even nuclear size.44 Metastasizing cancer cells have to overcome even more severe physical barriers as they intravasate across the endothelium into blood vessels or across epithelial tissues into lymphatic vessels. Cell deformability is, therefore, an important determinant of cell migration.45 For many cancer cells, for instance, lower stiffness correlates with higher motility.46
Migrating cells experience a complex combination of tensile, compressive, and shear deformations of varying amplitudes and rates. The effect of these mechanical parameters on cell deformability has been extensively characterized using quantitative biophysical techniques. To study the viscoelastic properties of cells without the impact of cell adhesion, cells can be detached from their substrate and measured in suspension by micropipette aspiration,47 optical stretching,48 parallel-plates rheometers,49 or high-throughput microfluidic methods.50 These measurements are mostly relevant for amoeboid migration where cells exhibit only weak adhesion to their environment. For mesenchymal migration, it is more relevant to study mechanics on adherent cells. Whole-cell measurements of adherent cells can be done by monolayer rheology or stretching51–53 or by single-cell atomic force microscopy.54,55 Localized measurements to resolve the mechanics of specific subcellular regions can be done by magnetic twisting cytometry,56 atomic force microscopy,57 optical or magnetic tweezers,58,59 or particle tracking microrheology.60 Some of these methods allow for in situ measurements in migrating cells.61,62
The nucleus is the stiffest and largest organelle with a stiffness ranging from 0.1 to 10 kPa, dependent on cell-type.63 The main contributors to the rigidity of the nucleus are heterochromatin and the nuclear lamina (also termed the nucleoskeleton), which contains lamin intermediate filaments.64–66 Intact nuclei in situ have a higher stiffness than isolated nuclei because the nucleus is coupled to the cytoskeleton through the LINC complex (composed of SUN-domain proteins and KASH-domain proteins, which physically connect the cytoskeleton to the nucleoskeleton).67 Since the nucleus is not only the stiffest but also the largest cellular organelle, it poses a major bottleneck for confined migration.68 When cells are embedded in collagen networks or microfabricated microchannels, their migration velocity linearly decreases with decreasing pore size until migration is physically blocked when the pore size reaches 10% of the nuclear cross section.69 Under highly confined conditions, cells can only move if they are able to remove blocking ECM fibers with proteolytic enzymes.69,70
2. Regulation of cell deformability through cytoskeletal crosstalk
The actin cytoskeleton is often considered the main determinant of cell mechanics. Drug-induced depolymerization of actin filaments indeed significantly softens cells both under non-adherent and adherent conditions.71,72 An important contribution of the actin cytoskeleton to cell stiffness comes from contractile forces generated by actin-myosin stress fibers73 and by the actin cortex.74 Intermediate filaments form dense networks that are mainly perinuclear, so they contribute little to cortical stiffness but strongly affect the cytoplasmic shear modulus75 and the resistance of cells to compression.55 For leukocytes and tumor cells performing 3D migration, the intermediate filament cytoskeleton is a major determinant of cell deformability. Intermediate filament protein deletion or network disruption causes significant cell softening while at the same time enhancing cell migration.33,76–78 Microtubules generally do not contribute much to cell stiffness, with drugs that interfere with microtubule polymerization having minor effects on cell mechanics.79 Recently, though, microtubules were shown to exhibit interesting mechano-responsive properties. Cytoskeletal compression induced by cyclic cell stretching or by confined migration was shown to stabilize deformed microtubules by triggering recruitment of the microtubule-binding protein CLASP2.80 When cells are transferred from rigid 2D substrates to softer 3D hydrogels, the mechanical contribution of microtubules becomes more important because actin stress fibers become less prominent. In cells migrating through collagen gels, microtubules for instance play a crucial role in mechanical support of cellular protrusions.81
It remains an open question how the interactions between the three cytoskeletal biopolymers influence the mechanics of the composite cytoskeleton. Theoretical models predict that composite networks composed of interpenetrating networks of rigid and flexible polymers are substantially stiffer than expected from the sum of the moduli of the separate networks.82 Rigid fiber networks by themselves are expected to be soft at low deformation because they deform in a non-affine manner, where the elasticity is governed by fiber bending.83 The presence of a background network of flexible polymer suppresses these non-affine bending deformations.82 Unfortunately, this prediction is difficult to directly test in live-cell experiments because it is very challenging to specifically remove one cytoskeletal network without also affecting the others. Microtubule depolymerization is, for instance, well-known to activate acto-myosin contraction by the release of the microtubule-associated guanine nucleotide exchange factor GEF-H1.84 At large strains, there is some evidence of mechanical synergy between the cytoskeletal networks. Epithelial cell layers are able to undergo extreme stretching under constant tension (“active superelasticity”) by strain-softening of the actin cortex followed by re-stiffening thanks to the keratin intermediate filament network.85 Physical cross-linking between actin and keratin is essential for the maintenance of epithelial stability.86 The ability of flexible polymers to suppress bending deformations of rigid polymers has the interesting consequence that the rigid polymers are reinforced against compressive loads.87 Under compressive loading, rigid polymers exhibit an Euler buckling instability at a critical compression force , where L is the polymer length. For microtubules, the critical compression force is only of order 1 pN.88 In the cell, however, microtubules can bear 100-fold larger compression forces because the surrounding actin and intermediate filament cytoskeleton constrains microtubule buckling.89,90 This is consistent with the so-called tensegrity model, which states that cellular shape stability is achieved via a balance between actin filaments and intermediate filaments loaded under tension, and microtubules and thick actin bundles under compression.91
B. Cell-free studies
1. Mechanical properties of individual cytoskeletal components
Live-cell mechanical measurements can be difficult to interpret in quantitative terms because they are sensitive to the amplitude, type and rate of deformation, geometry of the mechanical probe, the probed location in the cell, and the cell's extracellular environment.92 Cell-free studies provide a useful complement because they permit quantitative measurements of the mechanical properties of isolated cytoskeletal components, both at the single filament and at the network level.
At the single filament level, cytoskeletal biopolymers have been bent, stretched, compressed, and twisted using optical and magnetic tweezers,93–96 atomic force microscopy,97,98 and microfluidic devices.99 Actin filaments and microtubules have a high bending and stretching rigidity, but they break at rather low tensile strains (∼150% strains).100,101 Moreover, actin filaments become more fragile under torsion100 and microtubules soften upon repeated bending.99 This fragility is likely related to the fact that actin filaments and microtubules are made of globular subunits. In contrast, intermediate filaments are made of fibrous subunits held together by extensive lateral interactions. Intermediate filaments easily stretch and bend due to their open structure, and they can withstand tensile strains of more than 200% before rupture.97 Similar to a car's safety belt, intermediate filaments are soft under small and slow deformations but stiff under large and fast deformations.93 Recent evidence suggests that different intermediate filament proteins respond differently to tensile loads. When subjected to stretch-relax cycles, keratin filaments elongate with every cycle but keep the same stiffness, whereas vimentin filaments soften with every cycle but always return to the same initial length.102 It appears that vimentin stretches by monomer unfolding,103 whereas keratin filaments stretch by viscous sliding of subunits.102 It will be interesting to see what further diversity may be generated by co-polymerization of different intermediate filament proteins and by post-translational modifications.
On the network level, mechanical properties of cytoskeletal filaments are most conveniently probed by either bulk rheology or microrheology. In bulk rheology, cytoskeletal networks are sheared between the two parallel plates of a rheometer, providing a readout of the macroscopic viscoelastic response.104 Microrheology instead probes the localized viscoelastic response of a material by tracking the motion of embedded probe particles, either in response to thermal fluctuations (passive microrheology105) or to a force applied by optical or magnetic tweezers (active microrheology106) The mechanical response of cytoskeletal networks is determined by an interplay of the stiffness of the filaments and their interactions.
2. Effects of cross-linking on the mechanical properties of cytoskeletal networks
Actin filaments and microtubules form entangled networks that easily fluidize under shear due to filament disentanglement.107–109 Filament cross-linking via crosslinker proteins prevents this fluidization and causes the networks to strain-stiffen. This strain-stiffening response is only moderate for microtubules because of their high rigidity and because shearing causes force-induced unbinding of crosslinks.106,107 Actin networks exhibit more pronounced strain-stiffening because their elasticity is affected by the entropic elastic response of the filaments to tensile loading.110 Tensile loading reduces the conformational entropy of actin filaments, pulling out bending fluctuations, causing entropic strain-stiffening.111 Increased cross-link densities shift the onset of strain-stiffening to smaller shear strains because less excess length is stored in bending fluctuations when the crosslinks are more closely spaced.110 Some crosslinker proteins (most notably filamin) are so large that their compliance directly contributes to the network response. Crosslinker extensibility increases the rupture strain by postponing the point where the actin filaments experience tensile loading.112,113 Bundling of actin filaments, which is common at high concentrations of crosslinker proteins, suppresses entropic elasticity. Bundled actin networks still strain-stiffen,110 but by an enthalpic mechanism that involves a transition from soft bending modes at low strains to rigid stretching modes at high strain.114,115 Under compression, actin and microtubule networks soften due to filament buckling.55 For branched actin networks, compressive softening has been shown to be reversible, likely because the buckled filaments are prevented from collapsing by their connections with the network.116
The mechanical properties of intermediate filament networks differ in various respects from those of actin and microtubule networks. First, intermediate filaments form strain-stiffening networks even in the absence of any crosslinker proteins, as demonstrated for vimentin, neurofilaments, desmin, and keratin.117,118 The filaments spontaneously form crosslinks mediated by electrostatic interactions between their disordered C-terminal tails. Upon tail truncation, the networks no longer strain-stiffen.119–121 The effective cross-link density depends on the concentration of divalent cations such Mg2+, Ca2+, or Zn2+118,119,122,123 and is sensitive to the buffer ionic strength and pH.124 For keratins, there are additional hydrophobic interactions between the central rod domains that stiffen the networks.22 Second, intermediate filament networks have much larger rupture strains than actin and microtubule networks as a consequence of the larger single-filament extensibility. This is reflected in the dependence of the elastic modulus K on the applied shear stress σ. While actin networks only exhibit an entropic strain-stiffening regime where K increases as , intermediate filament networks exhibit an additional enthalpic regime where K increases more weakly, reflecting strain-induced filament alignment.117 After yielding, intermediate filament networks can even recover their initial shear modulus, likely by the re-establishment of tail-tail crosslinks.120,125
3. Mechanical characterizations of composite cytoskeletal networks
Recently, there has been increasing attention for the mechanical properties of cytoskeletal composites. Reconstitution of composite networks requires careful tuning of the buffer conditions since the different cytoskeletal polymers are traditionally reconstituted in their own optimized buffer conditions. Intermediate filaments are especially sensitive to solution pH and ionic concentrations, forming filaments of different widths and protein mass-per-length ratios depending on the buffer.126 Until now nearly all studies of composite networks have focused on two-component composites of cytoskeletal filaments co-polymerized in the absence of crosslinkers. At small strains, co-entangled composites (specifically combinations of actin/vimentin,127 actin/keratin,128,129 actin/microtubules,130 and vimentin/microtubules131) have generally been shown to exhibit a simple additive viscoelastic response. However, there is evidence for direct interactions of vimentin filaments with actin filaments132 as well as microtubules,133 which could potentially influence the network rheology. These interactions could potentially lead to cell-type specific cytoskeletal crosstalk, since they are mediated by the C-terminal tail of intermediate filaments that shows large length and sequence variations between different intermediate filament proteins. It was furthermore shown that vimentin can impose steric constraints that hamper actin network formation and thus cause network weakening.134 At large strains, there is evidence of synergistic enhancement of the mechanical properties in certain cytoskeletal composites. For actin/keratin composites, the strong strain-stiffening response of the keratin network was found to dominate the high-strain response of the composites.129 For actin/microtubule composites, microtubules were shown to promote strain-stiffening of the actin networks, even at low density.135,136 This effect was explained by the ability of rigid microtubules to suppress nonaffine bending fluctuations of actin filaments. It will be interesting to explore how these synergies are modified in the presence of crosslinkers. Recent work showed that when actin filaments and microtubules are crosslinked to each other by biotin-streptavidin, the composite is more elastic than when both filaments are independently crosslinked.137
To the best of our knowledge, there has so far been only one study of three-component networks combining actin, vimentin, and microtubules.138 It was shown by microrheology that the linear elastic modulus of the composite is dominated by actin, with little contribution from either microtubules or vimentin. Yet vimentin was shown to significantly extend the elastic regime to longer timescales. The authors proposed that the vimentin network that fills in the pore spaces of the actin network139 slows stress relaxation by constraining actin reptation. More work is needed to systematically study cytoskeletal composites and to explore the impact of cross-linking with cytolinker proteins such as plectin. Due to their high molecular weight, these proteins are difficult to purify. To circumvent this problem, one can engineer proteins that contain only the cytoskeletal binding domains separated by a spacer.140 Using this approach, we recently found that cross-linking with a plectin-mimetic crosslinker causes synergistic stiffening of actin-vimentin composites.141
III. CELL CONTRACTILITY
A. Live-cell studies
1. Acto-myosin contraction is the driving force behind cell migration
The actin cytoskeleton is the main engine behind cell migration.142 Depending on the extracellular environment, cells can switch between different mechanisms that use actin-based forces in different ways.143 Fibroblasts and other adherent cells perform mesenchymal migration, which relies on integrin-based adhesion to the extracellular matrix (ECM). The process occurs via a four-step cycle. First, actin polymerization pushes against the membrane at the leading edge, producing lamellipodia in cells migrating on flat rigid surfaces or pseudopodia in cells migrating in 3D extracellular matrices. Next, the cell generates integrin-based adhesions with the substrate that connect to the contractile machinery of acto-myosin stress fibers. Through a combination of pulling from the front and squeezing from the rear, the cell body moves forward. Finally, old adhesions are detached from the substrate or dissolved at the trailing edge.144 The contractile forces involved in cell migration have been measured through the traction forces exerted on the substrate. This is usually done by adhering cells to a hydrogel substrate with known mechanical properties, such as polyacrylamide. By measuring the displacements of fluorescent tracer particles incorporated in the gel with fluorescence microscopy, one can computationally infer the traction forces using continuum mechanics models.145 Adherent cells that experience strong confinement utilize a nuclear piston mechanism where actin-myosin contraction in front of the nucleus pulls the nucleus forward. Since the nucleus divides the cell in forward and rearward compartments, it acts as a piston that pressurizes the forward compartment and drives forward a cylindrical lobopodial protrusion.146 Weakly adherent cells such as leukocytes and physically confined fibroblasts and cancer cells perform amoeboid migration, characterized by spherical membrane blebs at the leading edge (reviewed in Ref. 147). Blebs are created by myosin-driven contraction of the actin cortex underneath the cell membrane, which builds up hydrostatic pressure in the cytoplasm. Local rupture of the actin cortex or its attachment to the membrane causes local membrane delamination, pushing forward a membrane bleb. Over time the actin cortex regrows under the bleb membrane and myosin contraction drives bleb retraction. Confinement can also induce other migration modes that require little substrate adhesion. Cells can move via friction generated by actin flows within the cortex generated by myosin contraction and actin turnover,148 and some tumor cells can still migrate by using active transport of water from the front to the back of the cell to propel themselves forward (osmotic engine model149).
2. The role of microtubules and intermediate filaments in acto-myosin mediated cell contractility
While not being components of the contractile machinery, both microtubules and intermediate filaments are important for regulating cell contraction. Microtubules negatively regulate the assembly and contractility of actin stress fibers by sequestering GEF-H1, an activator of the small GTPase Rho, in an autoinhibited state.150 Microtubule depolymerization by nocodazole releases active GEF-H1, leading to a global increase in contractility as measured by traction force microscopy.151 During both mesenchymal and amoeboid migration, microtubule depolymerization and consequent GEF-H1 is tightly regulated so that actin contractility can be precisely timed and localized in a mechanosensitive manner.152–154 In addition to biochemical regulation, it is likely that mechanical synergy is also involved in microtubule-based control of actin contractility, since microtubules are able to absorb some of the forces from the contractile actin cytoskeleton.91
Intermediate filaments likewise regulate actin-based cell contraction by a combination of mechanical synergy and biochemical signaling. In cells migrating on flat surfaces, vimentin has been reported to inhibit stress fiber assembly and contractility through down-regulating GEF-H1 and RhoA.155 Nevertheless, traction force measurements have shown that vimentin-null cells are less contractile than their wild-type counterparts.139 Taken together with the observation that vimentin filaments orient traction stresses along the front-rear axis, this suggests a mechanical synergy where vimentin helps build up and transmit larger contractile forces.156 Recently, it was shown by structured illumination microscopy and electron microscopy that vimentin filaments are closely associated with actin stress fibers, forming meshworks that wrap around stress fibers or co-align with them.42,139 Physical coupling between the two systems is dependent on the cytolinker protein plectin.157 Interestingly, it was recently shown that plectin binds vimentin in response to acto-myosin pulling forces.158 The mechanism for this mechanosensitivity is unknown but could involve catch bonding.159 Plectin-mediated coupling of actin and vimentin was recently shown to be essential for cells migrating via the nuclear piston mechanism.158 The vimentin network helps transmit acto-myosin pulling forces to the nucleus, thus enhancing the pressure in the front of the nucleus. It is not yet known whether intermediate filaments also influence cell migration modes driven by contractile activity of the actin cortex, but recent observations that vimentin and F-actin are associated within the cell cortex suggest this is likely.139
B. Cell-free reconstitution studies
There is an extensive body of work using cell-free reconstitution to elucidate the mechanisms by which myosin II motor proteins contract actin networks (reviewed in Ref. 14). The contraction mechanism has been found to depend on the actin network connectivity, which is controlled by filament length and by cross-linking. Well-connected networks of long filaments contract because myosins generate compressive stress that causes the actin filaments to buckle and break.160 In contrast, when the filaments are short, myosins contract the network by polarity sorting, transporting and clustering actin filament plus ends to form polar actin asters.161 In both cases, the length scale of contraction is set by the network connectivity. Global network contraction requires the actin network to be crosslinked above a critical percolation threshold.162 However, excessive cross-linking will prevent contraction by making the network too rigid.163 As described above, several cell migration mechanisms rely on myosin-driven contraction of the actin cortex. Recently several groups have been able to reconstitute biomimetic actin cortices by co-encapsulating actin and myosin inside cell-sized lipid vesicles. For weak actin-membrane attachment, the network detaches from the membrane upon contraction.164 In the case of stronger attachment, myosin contraction can cause membrane blebbing.165 Cortical flows that are important for driving amoeboid migration require not only myosin activity, but also network remodeling through actin depolymerization.166 Under particular conditions, crosslinked actin-myosin cortical networks in emulsion droplets have been observed to exhibit cortical flows,167 likely because myosin can promote actin turnover.168 Cell extracts, which contain additional proteins to promote actin turnover, also exhibit cortical flows when encapsulated in emulsion droplets.169–171 When these droplets are confined, the myosin-driven cortical flows can propel the droplets forward due to friction with the channel walls, mimicking amoeboid migration of nonadhesive cells.172
So far only few studies have looked at the effect of intermediate filaments or microtubules on contraction of actin-myosin networks. The addition of a vimentin network that interpenetrates an actin network has been shown to promote myosin-driven contraction by increasing the network connectivity.138 Similarly, also the addition of microtubules has been shown to promote uniform macroscopic myosin-driven contraction.173
IV. FRONT-REAR POLARIZATION
A. Live-cell studies
1. Crosstalk between actin and microtubules governs cell polarization in migration
Directed cell migration requires the breaking of cell symmetry to generate a cell front and a cell rear along an axis aligned with the direction of locomotion. Until now, the role of cytoskeletal crosstalk in front-rear polarity has mostly been studied in the context of 2D mesenchymal cell migration.174 It is long known that the microtubule cytoskeleton is essential for maintaining a polarized distribution of actin-based forces with actin polymerization in the front and myosin II-based contraction forces in the cell body and rear.175 Microtubules align along the axis of cell movement with their plus ends oriented toward the leading edge. They appear to stimulate actin-driven cell protrusion by multiple mechanisms. They activate Rac1 and inhibit Rho, therefore promoting actin polymerization and preventing myosin-II-driven contractility at the leading edge. Moreover, actin filaments have been observed to grow directly from microtubule tips toward the leading edge in growth cones of neurons, with the help of protein complexes involving APC and CLIP-170.176,177 There is an interesting actin/microtubule reciprocity, though, since the microtubules require guidance along actin stress fibers to reach the leading edge. This guidance requires actin-microtubule cross-linking, for instance, by ACF7, Growth Arrest-Specific Proteins (Gas2L1), CLIP-associating proteins (CLASPs), or drebrins (reviewed in Ref. 35). These proteins target growing microtubule plus ends by binding to EB (end-binding) proteins, and all of them except drebrin also possess a microtubule-lattice-binding domain. When these crosslinkers are depleted from cells, microtubules cease to grow along actin stress fibers and the microtubule array loses its front-rear polarity.178,179 Persistent cell migration is strongly hampered as a consequence, not only because actin-based protrusions are misregulated, but also because microtubules fail to reach cortical microtubule stabilizing complexes (CMSCs) that surround focal adhesions.180 Microtubule dynamics have been shown to regulate amoeboid cell migration by locally promoting the retraction of protrusions. In migrating dendritic cells, microtubule depolymerization within protrusions distant from the microtubule organizing center triggers actomyosin contractility, which is controlled by RhoA and its corresponding exchange factor Lfc.154 Tethering and stabilization of microtubule plus ends by CMSC binding is required for microtubule-dependent focal adhesion turnover, which is essential for migration (reviewed in Ref. 38). It is not yet clear how these crosstalk mechanisms are modified when cells perform 3D mesenchymal migration, but likely the core mechanisms are shared. One important new factor in 3D migration is that microtubules have a more important mechanical role and are needed to support pseudopodia.81 A second important new factor is that the rigidity of the nucleus hampers migration through small pores. It was recently shown that microtubules anchored to the nucleus play an important role in active transport of MT1-MMP, which degrades the extracellular matrix, to the cell surface where it drives extracellular matrix proteolysis in front of the nucleus.181
2. Intermediate filaments contribute to cell polarization via crosstalk with actin and microtubules
Although intermediate filaments lack intrinsic polarity, they do contribute to directed mesenchymal migration.36 When the vimentin network is disassembled using peptides or when vimentin expression is knocked down, cells lose their polarity and lamellipodia appear all around the cell.182 Vimentin forms closely associated parallel arrays with microtubules in migrating cells.156,183 Experiments conducted using vimentin-deficient mouse embryonic fibroblasts attached to polarized and non-polarized protein micropatterns demonstrated that the lack of vimentin alters microtubule organization, disrupting cell polarity.184 The two cytoskeletal networks organize in an interdependent manner. The vimentin distribution is polarized by a collaboration between active motor-driven transport along microtubules and actin-driven retrograde flow.185 Conversely, since the vimentin network is about 10-fold more long-lived than the microtubule network, it can serve as a template for guiding microtubule growth along previous microtubule tracks.156 This provides a feedback mechanism to sustain front-rear polarity. Moreover, the alignment of the vimentin network with the polarity axis mechanically integrates actin-based forces and orients them to promote directional migration.186 This mechanical integration is probably aided by vimentin-microtubule crosslinker proteins such as plectin and APC.183 In addition to this mechanical role, there is growing evidence for signaling functions of intermediate filaments in cell migration (reviewed in Ref. 34). At the cell periphery, there is for instance Rac-mediated crosstalk between vimentin and actin, where Rac causes vimentin disassembly by controlling the phosphorylation of vimentin at Ser-38, a p21-activated kinase phosphorylation site, promoting actin-driven membrane protrusion.182 Intermediate filaments also regulate focal adhesion clustering and turnover by binding integrins and via biochemical signaling.187
B. Cell-free reconstitution studies
1. Polarization crosstalk studies involving intermediate filaments
Several studies have explored how interactions between two different cytoskeletal filament types may contribute to the front-rear polarity of migrating cells. These studies mostly used surface assays where one or both cytoskeletal filaments were surface-anchored to facilitate imaging and control the geometry of interaction. These assays allow to probe the crosstalk involved in filament polymerization, an essential component of cell polarization. Just a few of these studies investigated the interplay of intermediate filaments with actin or microtubules. When surface-anchored microtubules are grown in the presence of an entangled vimentin network, they were found to be stabilized against depolymerization by direct interactions with vimentin filaments.133 Vimentin attachment reduced the catastrophe frequency and induced rescue of depolymerizing microtubules. However, in the absence of crosslinker proteins, these interactions were found to be short and infrequent. It is likely that vimentin-microtubule crosslinkers such as APC and plectin create more drastic effects on vimentin and microtubule polymerization. Interestingly, the vimentin-binding region of APC by itself was shown to promote vimentin polymerization,183 which may perhaps promote vimentin polymerization along microtubules. It was recently shown that actin and vimentin filaments do not interact in the absence of crosslinkers, but when an engineered plectin-mimicking crosslinker was added, actin filaments polymerized along surface-anchored vimentin filaments.141
2. Mutual regulation of polarization by actin and microtubules
A larger set of studies investigated the interplay of microtubules with different actin network structures designed to mimic structures found at the front of crawling cells. Branched or densely entangled actin network that mimic the dense actin array in the lamellipodium were shown to act as a steric barier for microtubule growth.188–190 However, when microtubules were crosslinked to actin by Tau protein, they were able to generate sufficient polymerization force to penetrate dense actin barriers.190 In contrast, when actin was arranged in stiff bundles that mimic actin stress fibers and bundles in filopodia, steric interactions were instead found to promote alignment and growth of microtubules along the actin bundles.188,190 Actin-microtule cross-linking proteins such as ACF7, Gas2L1, or CLASP2 were shown to promote actin-guided microtubule growth by allowing growing microtubules to be captured by and zippered along the actin bundles.140,188,190–193 Conversely, microtubules can also influence actin polymerization. Microtubule-lattice binding crosslinkers can induce guided polymerization of actin filaments along microtubules.192,194 Microtubule-tip binding crosslinkers can induce active transport of actin filaments by the growing microtubule tip.140,195 Computer simulations and theoretical modeling showed that this transport is driven by the affinity of the cross-linker for the chemically distinct microtubule tip region.195 These interactions may potentially enable growing microtubules to relocate newly nucleated actin filaments to the leading edge of the cell and thus boost migration. Altogether, these studies suggest that coupled polarization of the three cytoskeletal filament systems can at least partly be understood on the basis of a mechanical interplay.
V. COLLECTIVE MIGRATION AND INTERCELLULAR ADHESIONS
A. Live-cell studies
1. Intercellular adhesions establish mechanical coupling among cells during collective migration
Many cell types have the ability to synchronize their movement and perform collective migration. Collective migration is important for organogenesis and wound healing but also contributes to cancer metastasis. Depending on cell type and tissue context, different modes of collective migration can emerge. Epithelial cells tend to move as sheets adhered to the extracellular matrix,196 while cancer cells often migrate as three-dimensional strands or clusters through tissues.197 Remarkably, multi-cellular migrating structures behave similarly to liquid crystalline materials and undergo solid-to-liquid transitions in response to confinement. These jamming/unjamming transitions are linked to cell and nucleus shapes198,199 and are determined by molecular interactions that regulate cell–matrix and cell–cell adhesions.200–202 Traction force measurements for epithelial and endothelial monolayers have shown that cells within the monolayer tend to migrate in the direction in which the normal stress is greatest and the shear stress least.203 This mechanism of collective cell guidance called plithotaxis critically relies on mechanical coupling between the cells by cell–cell adhesions. Plithotaxis is regulated by the tumor suppressor protein merlin, which coordinates polarized Rac1 activation and lamellipodium formation at the multicellular scale.204 We speculate that, since Rac1 is an important shared regulator of all three cytoskeletal systems, there could be crosstalk with intermediate filaments and microtubules in plithotaxis. At the same time, intercellular adhesions help collectively migrating cells to establish supracellular polarization with leader cells at the front and follower cells behind.205 The leader cells explore the tissue environment using focal adhesions, find the path, and—if necessary—carve out a path by degrading the ECM. Cancer cells dynamically rearrange leader and follower positions during collective invasion to cope with the large energy usage of the leader cells.206
Epithelial and endothelial cells interact through mechanosensitive adherens junctions based on classical cadherins and VE-cadherins, respectively, which connect to the actin cytoskeleton via α-catenin and vinculin.205,207 Endothelial cells are additionally connected by complexus adhaerentes junctions that connect to vimentin via VE-cadherin.208 Epithelial cells are additionally connected by desmosomes based on desmosomal cadherins that connect to the keratin intermediate filament cytoskeleton via the adaptor proteins desmoplakin, plakophilin, and plakoglobin.207,209,210 The desmosomal junctions are expendable under homeostatic conditions but strictly required for preserving cell–cell adhesion under mechanical stress.211 As reviewed elsewhere,34,212 there is growing evidence that cytoskeletal crosstalk is important for the organization, dynamics and mechanoresponsiveness of intercellular adhesions during multicellular migration, although the exact extent is not clear.
2. Cytoskeletal crosstalk mediates intercellular adhesion dynamics
Keratin intermediate filaments and actin jointly influence desmosomes and adherens junctions because they are crosslinked via plectin. In epithelia, plectin organizes keratin into a rim-and-spoke configuration where contractile forces generated by acto-myosin are balanced by compressive elements provided by the keratin network, thus balancing internal tension and stabilizing cell–cell contacts.86 Deletion of plectin therefore causes perturbations of both desmosomes and adherens junctions. Although endothelial cells do not have desmosomes, also here plectin-mediated cross-linking between F-actin and vimentin intermediate filaments regulates adherens junction strength and tissue integrity.213 Migrating epithelial cells need to dynamically rearrange their adhesive contacts. Desmosome remodeling is dependent on both actin and keratin. Assembly of desmosomes at the leading edge and subsequent transport to the lateral sides is mediated through extensive actin remodeling, while more matured desmosomes are guided via keratins to the cell center to eventually disassemble.214 Intermediate filaments have an important mechanical role in distributing actin-myosin based forces, similar to their role in single-cell migration. Collective movement of both embryonic cells and astrocytes with proper leader-follower dynamics was shown to be highly dependent on this mechanical synergy of actin and intermediate filaments.215,216 3D tumor cultures of ex vivo breast cancer demonstrated heterotypic keratin expression between leader and follower cells,217 which was speculated to be necessary to regulate individual cytoplasmic viscoelasticity and mechanical coupling through desmosome anchoring during collective invasion.218 Invasion assays of epithelial cancers with manipulated keratin expressions also indicate that keratin expression can regulate migration mode; keratin-14-positive cells are leaders of migrating strands in organoid carcinoma assays, while keratin-8 and keratin-18 depletion can shift migration from individual to collective.217,219 Interestingly, keratins can also organize asymmetrically in migrating cells and localize in lamellipodia to support polarization and invasive phenotype, which is mediated by actin filaments.220
There is ample evidence that microtubule-actin crosstalk affects adherens junctions via mechanisms similar to those observed for focal adhesions. Adherens junctions contain multiple proteins that bind microtubule plus ends, including APC, ACF7, and CLASP.221,222 This allows microtubules to promote myosin II activation and local concentration of cadherin molecules223 and facilitates trafficking of junctional components to the cell surface.224,225 At the same time, microtubules promote junctional actin assembly by promoting liquid-liquid-phase separation of the actin nucleator cordon bleu (Cobl).226 During collective migration in vivo, it was found that cell–cell contacts differed in their requirement for dynamic microtubules along the leader-follower axis.227 Cells of the leading domain remained cohesive in the absence of dynamic microtubules, whereas dynamic microtubules were essential for the conversion of leader cells to epithelial followers. Interestingly, it was recently shown that physical confinement of collectively migrating cancer cells can induce the dissociation of leader cells by actin-microtubule crosstalk.228 Confinement-induced microtubule destabilization releases and activates GEF-H1, which promotes RhoA activation and results in leader cell detachment.
B. Cell-free reconstitution studies
Interactions between cytoskeletal filaments and cell–cell adhesion complexes are relatively unexplored in reconstituted systems. Adherens junctions are based on transmembrane cadherin adhesion receptors. The extracellular domains of cadherins on adjacent cells form adhesions by homodimerization. The cytosolic domains of cadherins bind β-catenin, which in turn binds α-catenin.229 Biochemical studies showed that actin filaments do not bind directly to cadherins, but are tethered indirectly via α-catenin.230,231 In solution, α-catenin does not interact with actin filaments and the cadherin-β-catenin complex simultaneously.232 However, under tension α-catenin stably connects the cadherin-catenin complex to actin filaments via a directionally asymmetric catch bond.233,234 This mechanosensitivity implies directional regulation of cell–cell adhesion in response to tension, which may connect to cooperative phenomena such as (un)jamming and plithotaxis mentioned above. Biochemical studies have shown that α-catenin influences the organization of junctional actin both directly, by inhibiting barbed-end growth, and indirectly, by interactions with various actin-binding proteins including Arp2/3.235 With the complexity of cadherin–actin interactions starting to become uncovered, it will be interesting to study how crosstalk with microtubules may further regulate adherens junctions. Recently, a first study in this direction was able to reconstitute the effect of microtubules on junctional actin nucleation, as was described above in live-cell studies.226 When dynamic microtubules were incubated together with G-actin and Cobl, actin filaments were found to be nucleated via Cobl condensates from both the tips and lattice of microtubules.
Regarding desmosomes, so far there have been mainly biochemical and structural studies, but few cell-free biophysical studies. Biochemical studies of the desmosomal cadherins showed that they form cell–cell connections via heterodimerization of the extracellular domains of desmogleins and desmocollins.236 Single-molecule force spectroscopy by AFM showed that the binding force is 30–40 pN.237 The cytoplasmic domains of desmosomal cadherins bind plakoglobin and plakophilins, an interaction that has also been reconstituted.238 Plakoglobin and plakophilins in turn bind to desmoplakin,239 which finally binds keratins. Biochemical and structural studies showed that desmoplakin interacts via its C-terminus with the rod domain of keratin.240,241 To the best of our knowledge, there are no cell-free studies addressing cytoskeletal crosstalk with desmosomes.
VI. PLASTICITY OF CELL MIGRATION
A. Live-cell studies
1. Cellular plasticity is mediated through focal adhesions
Most terminally differentiated cells such as epithelial and stromal cells migrate only during morphogenesis. However, tissue injury can induce cell plasticity. Mature cells can reenter the cell cycle and change their phenotype guided by paligenosis programs.242 Unfortunately, plasticity can also contribute to disease. For instance, malignant cancer cells are often hyperplastic, contributing to their invasiveness. A well-studied example of cell plasticity is epithelial-mesenchymal transition (EMT), a reversible process in which epithelial cells lose polarity through cytoskeletal remodeling, individualize and gain motility. EMT is a critical process in embryonic development and wound healing, but it also plays a key role in fibrosis and cancer invasion. EMT and the reverse mesenchymal-epithelial transition (MET) are influenced not only by biochemical cues, but also by mechanical properties of the ECM.243,244
Cells sense the mechanical properties of the ECM through their acto-myosin cytoskeleton at focal adhesions, mediating mechanotransduction and activating downstream plasticity mechanisms in response to environmental changes. Focal adhesions are based on integrin adhesion receptors that interact with the ECM through their extracellular domains and with the actin cytoskeleton through their cytoplasmic tails.245 The actin-integrin connection is mediated through talin and kindlin (reviewed in Ref. 246). Single integrins form small and transient junctions, but mechanical stimulation reinforces integrin adhesions by causing maturation into large focal adhesions. Upon mechanical stimulation, talin and kindlin undergo conformational changes that expose cryptic binding sites for additional cytoskeletal and signaling proteins.245 Mechanical stimulation further reinforces focal adhesions by inducing actin polymerization.247 Variations in the biochemical composition and physical properties of the ECM can elicit different 3D cell migration modes characterized by different amounts of cell-ECM adhesion.143,248 Highly crosslinked and dense matrices elicit lobopodial migration, characterized by a high number of focal adhesions and high actomyosin contractility. Less dense, fibrous environments elicit mesenchymal migration with a characteristic front-to-rear gradient of focal adhesions. In low-confining areas that lack adhesion sites, cells depend on bleb formation to drive themselves forward, a mechanism that does not require focal adhesions.143
Not much is known about the role of cytoskeletal crosstalk in migration plasticity. Microtubules are likely involved through their feedback interactions with actin near focal adhesions. It was recently shown that higher substrate rigidity promotes microtubule acetylation through the recruitment of α-tubulin acetyltransferase (α TAT) to focal adhesions by talin.42 In turn, microtubule acetylation tunes the mechanosensitivity of focal adhesions by promoting the release of GEF-H1 from microtubules to activate RhoA and thereby promote actomyosin contractility.42 In breast cancer cells, actin-microtubule crosstalk near focal adhesions via the scaffolding protein IQGAP1 was demonstrated to promote invasion in wound healing and transwell assays.249 In fibrosarcoma cells, the microtubule-destabilizing protein stathmin was shown to influence migration mode switching.250 Increased stathmin activity, and as a result less stable microtubule networks, promoted amoeboid-like migration, while phosphorylation of stathmin led to a more elongated migratory phenotype. In addition to crosstalk near focal adhesions, microtubules can also influence migration mode switching through mechanical effects. In confined or compressed cells, microtubules are stabilized through CLASP2 localization to the lattice, providing a mechanosensitive pathways for cells to adapt to highly constricting environments.80
2. Control of migration modes by cytoskeletal crosstalk at focal adhesions
The intermediate filament protein vimentin is considered a key cellular plasticity regulator and marker of tumor cell malignancy, especially based on its general upregulation in EMT and in motile cancer cells.251 Carcinoma cells in addition express integrin α6β4, which recruits vimentin to focal adhesions through its binding to plectin, promoting a 3D invasive phenotype switch.252 Together with nuclear lamins, vimentin contributes to migration plasticity through regulation of nuclear deformation, for instance allowing for a transition toward amoeboid-like and faster migration in Hela cells and melanoma cells when encountering confinement.33,253 Moreover, cells migrating under high confinement use their nucleus as a piston to squeeze through small pores. This complicated pulling mechanism is regulated through crosstalk between the vimentin cage around the nucleus and actomyosin in front of the nucleus.146
B. Cell-free reconstitution studies
Understanding the molecular basis of migration plasticity is an enormous challenge since integrin-based matrix adhesions contain over 100 types of molecules that are potentially mechanosensitive.254 One of the first mechanotransduction events during adhesion maturation is stretching of talin, followed by vinculin binding and activation. This core process was elegantly reconstituted by overlaying a network of actin-myosin bundles mimicking stress fibers on a talin-micropatterned surface.255 It was shown that direct binding of the contractile actin-myosin network to talin was sufficient to stretch the protein and induce the association and activation of vinculin. Talin binding facilitates vinculin activation by allosterically weakening the head–tail interaction that keeps it in an auto-inhibited conformation.256 Exposure of the actin-binding tail induces a positive feedback that reinforces the connection with actin.255 Using the same assay, it was shown that activated vinculin can interact with Arp2/3 complex-mediated branched actin networks and modify their organization by cross-linking actin filaments into bundles.257 This is likely an important step toward focal adhesion maturation. Single-molecule studies showed that vinculin forms a directionally asymmetric catch bond with F-actin.258 In this way, vinculin can organize the polarity of the actin cytoskeleton and contribute to front-rear asymmetry in migrating cells. Recently, the interaction of integrins, talin and kindlin, another major focal adhesion regulator, was reconstituted on giant unilamellar vesicles.259 It was shown that phosphoinositide-rich membranes recruit talin and kindlin, which then cause the formation of large integrin clusters that can recruit actin-myosin. Another study showed that membrane-bound talin can also activate vinculin and the two proteins together can link actin to the membrane.260
Cell-free reconstitution studies suggest that the actin cytoskeleton itself also contains proteins that mediate mechanotransduction. An example is filamin A (FLNA), a large multi-domain scaffolding protein that cross-links actin filaments and binds numerous proteins via cryptic binding sites along its length. Using reconstituted actin networks crosslinked with FLNA, it was shown that mechanical strain on the FLNA crosslinks alters its binding affinity for its binding partners.261 Both externally imposed bulk shear and contraction by myosin-II increased binding of the cytoplasmic tail of β-integrin while it weakened binding of FilGAP, a GTPase that inactivates Rac. Mechanical strain on FLNA can thus stabilize extracellular matrix binding and at the same time influence actin dynamics through Rac activity.
VII. THE ROAD AHEAD
Cytoskeletal crosstalk is increasingly recognized as a major determinant of cell migration. Coupling among the actin, microtubule, and intermediate filament cytoskeleton influences cell migration by regulating cell deformability, contractility, front-rear polarity, and migration plasticity. Coupling of filaments through entanglements, cross-linking, and bundling regulates a variety of mechanisms that mediate cellular mechanics and cytoskeletal dynamics involved in these migration strategies (Fig. 3). To complicate matters, there is growing evidence that a fourth cytoskeletal protein family, the septins, also strongly impacts cell migration. Septins are well-equipped to mediate cytoskeletal crosstalk since they can bind to the cell membrane, actin and microtubules.262 Recent research demonstrated roles of septins in mesenchymal and amoeboid single-cell migration263,264 and in the regulation of endothelial and epithelial cell–cell adhesion.265,266 Elucidating the biophysical mechanisms by which cytoskeletal crosstalk regulates cell migration is challenging due to the enormous molecular complexity of the cell and feedback between mechanical forces and biochemical signaling. Cell-free reconstitution provides a valuable complement to live-cell studies because it simplifies the challenge of separating biochemical and physical contributions to cytoskeletal crosstalk. We note that caution should be used when translating findings regarding cytoskeletal crosstalk from bottom-up studies to cells. Cell-free model systems present a highly simplified environment in terms of molecular complexity, crowding, ionic conditions, spatial heterogeneities, mechanical conditions, etc. Cytoskeletal interactions or mechanical effects observed in a cell-free system do not necessarily occur (in the same way) in the cellular environment. The power of cell-free studies is to develop and test hypotheses regarding possible interactions and their biophysical impact under well-controlled conditions. Using cell-based studies, one can then test how these findings carry over to the cellular environment. Clearly, there still remains an enormous gap between the complexity of cells and the simplicity of reconstituted systems. How can this gap be bridged?
We propose different routes to bridge this gap (Fig. 4). One obvious direction to bridge this gap is to enhance the complexity of cell-free assays. Simple assays combining cytoskeletal filaments without any accessory proteins showed that steric interactions alone suffice to give rise to nonlinear stiffening and enhanced compressive strength. There is some evidence that intermediate filaments directly interact with actin filaments and microtubules via electrostatic interactions,132,133 but this could be an artifact of in vitro conditions. Single-molecule measurements of filament interactions within cells or cell lysates could shed light on this issue. An important next step to bridge the gap to the cell is to incorporate accessory proteins that mediate cytoskeletal coupling. Several studies have shown that cross-linking via cytolinkers is sufficient to give rise to cytoskeletal filament co-alignment and mechanical synergy (e.g., Refs. 140 and 141). More detailed investigations of the effects of cytolinkers, both in vitro and in cells, will be important to delineate their roles in cytoskeletal co-organization, mechanical synergy, and mechanotransduction. In the cell, cytoskeletal crosstalk is guided by geometrical constraints provided by the cell membrane. The membrane organizes the cytoskeleton through spatial confinement and by providing adhesion sites where cytoskeletal filaments are nucleated or anchored. Reconstitution experiments have begun to recapitulate these constraints by encapsulating cytoskeletal proteins inside cell-sized emulsion droplets or lipid vesicles, including actin/microtubule and actin/keratin composites.128,267 These model systems could form a basis for reconstituting synthetic cells capable of migration. Adhesion-independent migration is probably easiest to reconstitute. Flow-driven confined migration was recently reconstituted, although based on a cell extract, so the minimal set of ingredients is not yet known.172 It will be interesting to incorporate microtubules and/or intermediate filaments in this assay to control cell polarity and mechanics. Mesenchymal migration is likely more challenging to reconstitute because it requires coordinated actin polymerization, contraction, and cell-matrix adhesion. Motility driven by actin polymerization has been successfully reconstituted on the outer surface of lipid vesicles (reviewed in Ref. 268), but motility of vesicles with actin polymerization inside will require substrate adhesion. Surface micropatterning provides an interesting approach to impose polarized shapes to synthetic cells by forcing them to adapt to the pattern shape and size.269 In addition, one can use light-induced dimerization to induce spatial patterning and symmetry breaking of cytoskeletal networks. Light-inducible dimers (LIDs) come from photoactivatable systems naturally occurring in plants and allow for reversible photoactivation.270 Recently, it was, for instance, shown that microtubule-interacting proteins fused to optochemical dimerization domains can be used to drive symmetry breaking of microtubule networks inside emulsion droplets.271
The opposite direction to bridge the gap between live-cell and cell-free studies is to tame the complexity of living cells. Some of the same techniques that can provide more control over cell-free systems can also provide control over the behavior of living cells. Surface micropatterning, for instance, allows one to confine cells to adhesive islands with precisely controlled geometries, forcing the cells to adopt prescribed shape and corresponding cytoskeletal organizations. Imaging many cells adhered on the same type of pattern greatly facilitates quantification of cytoskeletal crosstalk.184 Moreover, micropatterning can be used to investigate how cytoskeletal interactions affect single-cell and collective cell migration dynamics.272 Light-inducible dimerization can be used to manipulate cytoskeletal interactions with high spatial and temporal control. It was, for instance, recently shown that F-actin can be crosslinked to microtubule plus ends by transfecting cells with an iLID-tagged EB-binding SxIP peptide and SspB-tagged actin-binding domains.273 This could be an interesting tool to systematically study the crosstalk of microtubules with actin and intermediate filaments that takes place near cell adhesions. Finally, molecular tension sensors provide a very interesting tool to selectively interrogate mechanical interactions between cytoskeletal networks. Tension sensors consist of two fluorescent proteins separated by a peptide with a calibrated mechanical compliance. Under strain, the fluorescent proteins are separated, decreasing fluorescence energy transfer (FRET) between them. By embedding a tension sensor in the actin crosslinker FLNA, it was recently shown that molecular tension can be measured within the actin cytoskeleton.274 It will be interesting to use a similar approach to measure tension within the intermediate filament cytoskeleton and test force transmission between the actin and intermediate filament cytoskeleton.
ACKNOWLEDGMENTS
We gratefully acknowledge funding from the VICI project How cytoskeletal teamwork makes cells strong (Project No. VI.C.182.004) and from the OCENW.GROOT.20t9.O22 project The Active Matter Physics of Collective Metastasis, both financed by the Dutch Research Council (NWO).
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
James P. Conboy, Irene Istúriz Petitjean, Anouk van der Net, and Gijsje H. Koenderink contributed equally to this work.
James P. Conboy: Conceptualization (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (supporting). Irene Istúriz Petitjean: Conceptualization (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (supporting). Anouk van der Net: Conceptualization (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (supporting). Gijsje H. Koenderink: Conceptualization (equal); Funding acquisition (lead); Project administration (lead); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal).
DATA AVAILABILITY
The data that support the findings of this study are available from the corresponding author upon reasonable request.
References
- 1. Scarpa E. and Mayor R., “ Collective cell migration in development,” J. Cell Biol. 212, 143–155 (2016). 10.1083/jcb.201508047 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2. Lamalice L., Boeuf F. L., and Huot J., “ Endothelial cell migration during angiogenesis,” Cir. Res. 100, 782–794 (2007). 10.1161/01.RES.0000259593.07661.1e [DOI] [PubMed] [Google Scholar]
- 3. Su P., Tian Y., Yang C., Ma X., Wang X., Pei J., and Qian A., “ Mesenchymal stem cell migration during bone formation and bone diseases therapy,” Int. J. Mol. Sci. 19, 2343 (2018). 10.3390/ijms19082343 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4. Fu X., Liu G., Halim A., Ju Y., Luo Q., and Song G., “ Mesenchymal stem cell migration and tissue repair,” Cells. 8, 784 (2019). 10.3390/cells8080784 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5. Delgado M. and Lennon-Duménil A. M., “ How cell migration helps immune sentinels,” Front. Cell Dev. Biol. 10, 932472 (2022). 10.3389/fcell.2022.932472 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6. Liu J., Zhang X., Cheng Y., and Cao X., “ Dendritic cell migration in inflammation and immunity,” Cell. Mol. Immunol. 18, 2461–2471 (2021). 10.1038/s41423-021-00726-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7. Mierke C. T., “ The matrix environmental and cell mechanical properties regulate cell migration and contribute to the invasive phenotype of cancer cells,” Rep. Prog. Phys. 82, 064602 (2019). 10.1088/1361-6633/ab1628 [DOI] [PubMed] [Google Scholar]
- 8. Mostowy S. and Cossart P., “ Septins: The fourth component of the cytoskeleton,” Nature Rev. Mol. Biol. 13, 183 (2012). 10.1038/nrm3284 [DOI] [PubMed] [Google Scholar]
- 9. Galkin V. E., Orlova A., Vos M. R., Schröder G. F., and Egelman E. H., “ Near-atomic resolution for one state of F-actin,” Structure 23, 173–182 (2015). 10.1016/j.str.2014.11.006 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10. Kang H., Bradley M. J., McCullough B. R., Pierre A., Grintsevich E. E., Reisler E., and De La Cruz E. M., “ Identification of cation-binding sites on actin that drive polymerization and modulate bending stiffness,” Proc. Natl. Acad. Sci. U. S. A. 109, 16923–16927 (2012). 10.1073/pnas.1211078109 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11. Carman P. J., Barrie K. R., Rebowski G., and Dominguez R., “ Structures of the free and capped ends of the actin filament,” Science 380, 1287–1292 (2023). 10.1126/science.adg6812 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12. Watanabe N. and Mitchison T. J., “ Single-molecule speckle analysis of actin filament turnover in lamellipodia,” Science 295, 1083–1086 (2002). 10.1126/science.1067470 [DOI] [PubMed] [Google Scholar]
- 13. Fritzsche M., Lewalle A., Duke T., Kruse K., and Charras G., “ Analysis of turnover dynamics of the submembranous actin cortex,” Mol. Biol. Cell 24, 757 (2013). 10.1091/mbc.e12-06-0485 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14. Murrell M., Oakes P. W., Lenz M., and Gardel M. L., “ Forcing cells into shape: The mechanics of actomyosin contractility,” Nat. Rev. Mol. Cell Biol. 16, 486–498 (2015). 10.1038/nrm4012 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15. Zhang R., LaFrance B., and Nogales E., “ Separating the effects of nucleotide and EB binding on microtubule structure,” Proc. Natl. Acad. Sci. U. S. A. 115, E6191–E6200 (2018). 10.1073/pnas.1802637115 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16. Taute K. M., Pampaloni F., Frey E., and Florin E.-L., “ Microtubule dynamics depart from the wormlike chain model,” Phys. Rev. Lett. 100, 028102 (2008). 10.1103/PhysRevLett.100.028102 [DOI] [PubMed] [Google Scholar]
- 17. Brouhard G. J., “ Dynamic instability 30 years later: Complexities in microtubule growth and catastrophe,” Mol. Biol. Cell 26, 1207–1210 (2015). 10.1091/mbc.E13-10-0594 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18. Rölleke U., Kumari P., Meyer R., and Köster S., “ The unique biomechanics of intermediate filaments—From single filaments to cells and tissues,” Curr. Opin. Cell Biol. 85, 102263 (2023). 10.1016/j.ceb.2023.102263 [DOI] [PubMed] [Google Scholar]
- 19. Kang H., Bradley M. J., Cao W., Zhou K., Grintsevich E. E., Michelot A., Sindelar C. V., Hochstrasser M., and De La Cruz E. M., “ Site-specific cation release drives actin filament severing by vertebrate cofilin,” Proc. Natl. Acad. Sci. U. S. A. 111, 17821–17826 (2014). 10.1073/pnas.1413397111 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20. Beck R., Deek J., Choi M. C., Ikawa T., Watanabe O., Frey E., Pincus P., and Safinya C. R., “ Unconventional salt trend from soft to stiff in single neurofilament biopolymers,” Langmuir 26, 18595–18599 (2010). 10.1021/la103655x [DOI] [PubMed] [Google Scholar]
- 21. Nöding B. and Köster S., “ Intermediate filaments in small configuration spaces,” Phys. Rev. Lett. 108, 088101 (2012). 10.1103/PhysRevLett.108.088101 [DOI] [PubMed] [Google Scholar]
- 22. Pawelzyk P., Mücke N., Herrmann H., and Willenbacher N., “ Attractive interactions among intermediate filaments determine network mechanics in vitro,” PLoS One. 9, e93194 (2014). 10.1371/journal.pone.0093194 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23. Schopferer M., Bär H., Hochstein B., Sharma S., Mücke N., Herrmann H., and Willenbacher N., “ Desmin and vimentin intermediate filament networks: Their viscoelastic properties investigated by mechanical rheometry,” J. Mol. Biol. 388, 133–143 (2009). 10.1016/j.jmb.2009.03.005 [DOI] [PubMed] [Google Scholar]
- 24. Mücke N., Kreplak L., Kirmse R., Wedig T., Herrmann H., Aebi U., and Langowski J., “ Assessing the flexibility of intermediate filaments by atomic force microscopy,” J. Mol. Biol. 335, 1241–1250 (2004). 10.1016/j.jmb.2003.11.038 [DOI] [PubMed] [Google Scholar]
- 25. Nöding B., Herrmann H., and Köster S., “ Direct observation of subunit exchange along mature vimentin intermediate filaments,” Biophys. J. 107, 2923–2931 (2014). 10.1016/j.bpj.2014.09.050 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26. Tran Q. D., Sorichetti V., Pehau-Arnaudet G., Lenz M., and Leduc C., “ Fragmentation and entanglement limit vimentin intermediate filament assembly,” Phys. Rev. X 13, 011014 (2023). 10.1103/PhysRevX.13.011014 [DOI] [Google Scholar]
- 27. Robert A., Rossow M. J., Hookway C., Adam S. A., and Gelfand V. I., “ Vimentin filament precursors exchange subunits in an ATP-dependent manner,” Proc. Natl. Acad. Sci. U. S. A. 112, E3505–E3514 (2015). 10.1073/pnas.1505303112 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28. Hookway C., Ding L., Davidson M. W., Rappoport J. Z., Danuser G., and Gelfand V. I., “ Microtubule-dependent transport and dynamics of vimentin intermediate filaments,” Mol. Biol. Cell 26, 1675–1686 (2015). 10.1091/mbc.E14-09-1398 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29. Rottner K. and Schaks M., “ Assembling actin filaments for protrusion,” Curr. Opin. Cell Biol. 56, 53–63 (2019). 10.1016/j.ceb.2018.09.004 [DOI] [PubMed] [Google Scholar]
- 30. Ullo M. F. and Logue J. S., “ ADF and cofilin-1 collaborate to promote cortical actin flow and the leader bleb-based migration of confined cells,” eLife 10, 67856 (2021). 10.7554/eLife.67856 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31. Akhmanova A. and Kapitein L. C., “ Mechanisms of microtubule organization in differentiated animal cells,” Nat. Rev. Mol. Cell Biol. 23, 541–558 (2022). 10.1038/s41580-022-00473-y [DOI] [PubMed] [Google Scholar]
- 32. Patteson A. E., Vahabikashi A., Pogoda K., Adam S. A., Mandal K., Kittisopikul M., Sivagurunathan S., Goldman A., Goldman R. D., and Janmey P. A., “ Vimentin protects cells against nuclear rupture and DNA damage during migration,” J. Cell Biol. 218, 4079–4092 (2019). 10.1083/jcb.201902046 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33. Lavenus S. B., Tudor S. M., Ullo M. F., Vosatka K. W., and Logue J. S., “ A flexible network of vimentin intermediate filaments promotes migration of amoeboid cancer cells through confined environments,” J. Biol. Chem. 295, 6700–6709 (2020). 10.1074/jbc.RA119.011537 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34. Seetharaman S. and Etienne-Manneville S., “ Cytoskeletal crosstalk in cell migration,” Trends Cell Biol. 30, 720–735 (2020). 10.1016/j.tcb.2020.06.004 [DOI] [PubMed] [Google Scholar]
- 35. Dogterom M. and Koenderink G. H., “ Actin–microtubule crosstalk in cell biology,” Nat. Rev. Mol. Cell Biol. 20, 38–54 (2019). 10.1038/s41580-018-0067-1 [DOI] [PubMed] [Google Scholar]
- 36. Chung B. M., Rotty J. D., and Coulombe P. A., “ Networking galore: Intermediate filaments and cell migration,” Curr. Opin. Cell Biol. 25, 600–612 (2013). 10.1016/j.ceb.2013.06.008 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37. Pimm M. L. and Henty-Ridilla J. L., “ New twists in actin-microtubule interactions,” Mol. Biol. Cell 32, 211–217 (2021). 10.1091/mbc.E19-09-0491 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38. Schmidt C. J. and Stehbens S. J., “ Microtubule control of migration: Coordination in confinement,” Curr. Opin. Cell Biol. 86, 102289 (2023). 10.1016/j.ceb.2023.102289 [DOI] [PubMed] [Google Scholar]
- 39. Buttenschön A. and Edelstein-Keshet L., “ Bridging from single to collective cell migration: A review of models and links to experiments,” PLoS Comput. Biol. 16, e1008411 (2020). 10.1371/journal.pcbi.1008411 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40. Banerjee S. and Marchetti M. C., “ Continuum models of collective cell migration,” Adv. Exp. Med. Biol. 1146, 45–66 (2019). 10.1007/978-3-030-17593-1_4 [DOI] [PubMed] [Google Scholar]
- 41. MacTaggart B. and Kashina A., “ Posttranslational modifications of the cytoskeleton,” Cytoskeleton 78, 142–173 (2021). 10.1002/cm.21679 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42. Seetharaman S., Vianay B., Roca V., Farrugia A. J., De Pascalis C., Boëda B., Dingli F., Loew D., Vassilopoulos S., Bershadsky A., Théry M., and Etienne-Manneville S., “ Microtubules tune mechanosensitive cell responses,” Nat. Mater. 21, 366–377 (2022). 10.1038/s41563-021-01108-x [DOI] [PubMed] [Google Scholar]
- 43. te Boekhorst V., Preziosi L., and Friedl P., “ Plasticity of cell migration in vivo and in silico,” Annu. Rev. Cell Dev. Biol. 32, 491–526 (2016). 10.1146/annurev-cellbio-111315-125201 [DOI] [PubMed] [Google Scholar]
- 44. Wolf K., Alexander S., Schacht V., Coussens L. M., von Andrian U. H., van Rheenen J., Deryugina E., and Friedl P., “ Collagen-based cell migration models in vitro and in vivo,” Semin. Cell Dev. Biol. 20, 931–941 (2009). 10.1016/j.semcdb.2009.08.005 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45. Guck J., Lautenschläger F., Paschke S., and Beil M., “ Critical review: Cellular mechanobiology and amoeboid migration,” Integr. Biol. 2, 575–583 (2010). 10.1039/c0ib00050g [DOI] [PubMed] [Google Scholar]
- 46. Liu Z., Lee S. J., Park S., Konstantopoulos K., Glunde K., Chen Y., and Barman I., “ Cancer cells display increased migration and deformability in pace with metastatic progression,” FASEB J. 34, 9307–9315 (2020). 10.1096/fj.202000101RR [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47. González-Bermúdez B., Guinea G. V., and Plaza G. R., “ Advances in micropipette aspiration: Applications in cell biomechanics, models, and extended studies,” Biophys. J. 116, 587–594 (2019). 10.1016/j.bpj.2019.01.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48. Guck J., Ananthakrishnan R., Mahmood H., Moon T. J., Cunningham C. C., and Käs J., “ The optical stretcher: A novel laser tool to micromanipulate cells,” Biophys. J. 81, 767–784 (2001). 10.1016/S0006-3495(01)75740-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49. Thoumine O. and Ott A., “ Time scale dependent viscoelastic and contractile regimes in fibroblasts probed by microplate manipulation,” J. Cell Sci. 110, 2109–2116 (1997). 10.1242/jcs.110.17.2109 [DOI] [PubMed] [Google Scholar]
- 50. Urbanska M., Muñoz H. E., Shaw Bagnall J., Otto O., Manalis S. R., Di Carlo D., and Guck J., “ A comparison of microfluidic methods for high-throughput cell deformability measurements,” Nat. Methods 17, 587–593 (2020). 10.1038/s41592-020-0818-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51. Fernandez P., Heymann L., Ott A., Aksel N., and Pullarkat P., “ Shear rheology of a cell monolayer,” New J. Phys. 9, 419 (2007). 10.1088/1367-2630/9/11/419 [DOI] [Google Scholar]
- 52. Sutton A. A., Molter C. W., Amini A., Idicula J., Furman M., Tirgar P., Tao Y., Ghagre A., Koushki N., Khavari A., and Ehrlicher A. J., “ Cell monolayer deformation microscopy reveals mechanical fragility of cell monolayers following EMT,” Biophys. J. 121, 629–643 (2022). 10.1016/j.bpj.2022.01.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53. Bashir K. M. I., Lee S., Jung D. H., Basu S. K., Cho M.-G., and Wierschem A., “ Narrow-gap rheometry: A novel method for measuring cell mechanics,” Cells 11, 2010 (2022). 10.3390/cells11132010 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54. Beicker K., O'Brien III E. T., Falvo M. R., and Superfine R., “ Vertical light sheet enhanced side-view imaging for AFM cell mechanics studies,” Sci. Rep. 8, 1504 (2018). 10.1038/s41598-018-19791-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55. Pogoda K., Byfield F., Deptuła P., Cieśluk M., Suprewicz Ł., Skłodowski K., Shivers J. L., van Oosten A., Cruz K., Tarasovetc E., Grishchuk E. L., Mackintosh F. C., Bucki R., Patteson A. E., and Janmey P. A., “ Unique role of vimentin networks in compression stiffening of cells and protection of nuclei from compressive stress,” Nano Lett. 22, 4725–4732 (2022). 10.1021/acs.nanolett.2c00736 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56. Fabry B., Maksym G. N., Butler J. P., Glogauer M., Navajas D., and Fredberg J. J., “ Scaling the microrheology of living cells,” Phys. Rev. Lett. 87, 148102 (2001). 10.1103/PhysRevLett.87.148102 [DOI] [PubMed] [Google Scholar]
- 57. Kerdegari S., Canepa P., Odino D., Oropesa-Nuñez R., Relini A., Cavalleri O., and Canale C., “ Insights in cell biomechanics through atomic force microscopy,” Materials 16, 2980 (2023). 10.3390/ma16082980 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58. Català-Castro F., Schäffer E., and Krieg M., “ Exploring cell and tissue mechanics with optical tweezers,” J. Cell Sci. 135, 259–355 (2022). 10.1242/jcs.259355 [DOI] [PubMed] [Google Scholar]
- 59. Bausch A. R., Ziemann F., Boulbitch A. A., Jacobson K., and Sackmann E., “ Local measurements of viscoelastic parameters of adherent cell surfaces by magnetic bead microrheometry,” Biophys. J. 75, 2038–2049 (1998). 10.1016/S0006-3495(98)77646-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60. Lau A. W. C., Hoffman B. D., Davies A., Crocker J. C., and Lubensky T. C., “ Microrheology, stress fluctuations, and active behavior of living cells,” Phys. Rev. Lett. 91, 198101 (2003). 10.1103/PhysRevLett.91.198101 [DOI] [PubMed] [Google Scholar]
- 61. Panorchan P., Lee J. S. H., Kole T. P., Tseng Y., and Wirtz D., “ Microrheology and ROCK signaling of human endothelial cells embedded in a 3D matrix,” Biophys. J. 91, 3499–3507 (2006). 10.1529/biophysj.106.084988 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62. Staunton J. R., Doss B. L., Lindsay S., and Ros R., “ Correlating confocal microscopy and atomic force indentation reveals metastatic cancer cells stiffen during invasion into collagen I matrices,” Sci. Rep. 6, 19686 (2016). 10.1038/srep19686 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63. Swift J., Ivanovska I. L., Buxboim A., Harada T., Dingal P. C. D. P., Pinter J., Pajerowski J. D., Spinler K. R., Shin J.-W., Tewari M., Rehfeldt F., Speicher D. W., and Discher D. E., “ Nuclear lamin-A scales with tissue stiffness and enhances matrix-directed differentiation,” Science 341, 1240104 (2013). 10.1126/science.1240104 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64. Lammerding J., Fong L. G., Ji J. Y., Reue K., Stewart C. L., Young S. G., and Lee R. T., “ Lamins A and C but not lamin B1 regulate nuclear mechanics,” J. Biol. Chem. 281, 25768–25780 (2006). 10.1074/jbc.M513511200 [DOI] [PubMed] [Google Scholar]
- 65. Stephens A. D., Banigan E. J., Adam S. A., Goldman R. D., and Marko J. F., “ Chromatin and lamin A determine two different mechanical response regimes of the cell nucleus,” Mol. Biol. Cell. 28, 1984–1996 (2017). 10.1091/mbc.e16-09-0653 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66. Bell E. S., Shah P., Zuela-Sopilniak N., Kim D., Varlet A.-A., Morival J. L. P., McGregor A. L., Isermann P., Davidson P. M., Elacqua J. J., Lakins J. N., Vahdat L., Weaver V. M., Smolka M. B., Span P. N., and Lammerding J., “ Low lamin A levels enhance confined cell migration and metastatic capacity in breast cancer,” Oncogene 41, 4211–4230 (2022). 10.1038/s41388-022-02420-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67. Liu H., Wen J., Xiao Y., Liu J., Hopyan S., Radisic M., Simmons C. A., and Sun Y., “ In situ mechanical characterization of the cell nucleus by atomic force microscopy,” ACS Nano 8, 3821–3828 (2014). 10.1021/nn500553z [DOI] [PubMed] [Google Scholar]
- 68. McGregor A. L., Hsia C. R., and Lammerding J., “ Squish and squeeze—The nucleus as a physical barrier during migration in confined environments,” Curr. Opin. Cell Biol. 40, 32–40 (2016). 10.1016/j.ceb.2016.01.011 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69. Wolf K., te Lindert M., Krause M., Alexander S., te Riet J., Willis A. L., Hoffman R. M., Figdor C. G., Weiss S. J., and Friedl P., “ Physical limits of cell migration: Control by ECM space and nuclear deformation and tuning by proteolysis and traction force,” J. Cell Biol. 201, 1069–1084 (2013). 10.1083/jcb.201210152 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70. Davidson P. M., Denais C., Bakshi M. C., and Lammerding J., “ Nuclear deformability constitutes a rate-limiting step during cell migration in 3-D environments,” Cell. Mol. Bioeng. 7, 293–306 (2014). 10.1007/s12195-014-0342-y [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71. Lautenschläger F., Paschke S., Schinkinger S., Bruel A., Beil M., and Guck J., “ The regulatory role of cell mechanics for migration of differentiating myeloid cells,” Proc. Natl. Acad. Sci. U. S. A. 106, 15696–15701 (2009). 10.1073/pnas.0811261106 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72. Rotsch C. and Radmacher M., “ Drug-induced changes of cytoskeletal structure and mechanics in fibroblasts: An atomic force microscopy study,” Biophys. J. 78, 520–535 (2000). 10.1016/S0006-3495(00)76614-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73. Martens J. C. and Radmacher M., “ Softening of the actin cytoskeleton by inhibition of myosin II,” Pflugers Arch. 456, 95–100 (2008). 10.1007/s00424-007-0419-8 [DOI] [PubMed] [Google Scholar]
- 74. Fischer-Friedrich E., Toyoda Y., Cattin C. J., Müller D. J., Hyman A. A., and Jülicher F., “ Rheology of the active cell cortex in mitosis,” Biophys. J. 111, 589–600 (2016). 10.1016/j.bpj.2016.06.008 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75. Guo M., Ehrlicher A. J., Mahammad S., Fabich H., Jensen M. H., Moore J. R., Fredberg J. J., Goldman R. D., and Weitz D. A., “ The role of vimentin intermediate filaments in cortical and cytoplasmic mechanics,” Biophys. J. 105, 1562–1568 (2013). 10.1016/j.bpj.2013.08.037 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76. Brown M. J., Hallam J. A., Colucci-Guyon E., and Shaw S., “ Rigidity of circulating lymphocytes is primarily conferred by vimentin intermediate filaments,” J. Immunol. 166, 6640–6646 (2001). 10.4049/jimmunol.166.11.6640 [DOI] [PubMed] [Google Scholar]
- 77. Seltmann K., Fritsch A. W., Käs J. A., and Magin T. M., “ Keratins significantly contribute to cell stiffness and impact invasive behavior,” Proc. Natl. Acad. Sci. U. S. A. 110, 18507–18512 (2013). 10.1073/pnas.1310493110 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78. Patteson A. E., Pogoda K., Byfield F. J., Mandal K., Ostrowska-Podhorodecka Z., Charrier E. E., Galie P. A., Deptuła P., Bucki R., McCulloch C. A., and Janmey P. A., “ Loss of vimentin enhances cell motility through small confining spaces,” Small 15, e1903180 (2019). 10.1002/smll.201903180 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79. Wang N., “ Mechanical interactions among cytoskeletal filaments,” Hypertension 32, 162–165 (1998). 10.1161/01.HYP.32.1.162 [DOI] [PubMed] [Google Scholar]
- 80. Li Y., Kučera O., Cuvelier D., Rutkowski D. M., Deygas M., Rai D., Pavlovič T., Vicente F. N., Piel M., Giannone G., Vavylonis D., Akhmanova A., Blanchoin L., and Théry M., “ Compressive forces stabilize microtubules in living cells,” Nat. Mater. 22, 913–924 (2023). 10.1038/s41563-023-01578-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81. Bouchet B. P., Noordstra I., van Amersfoort M., Katrukha E. A., Ammon Y.-C., Ter Hoeve N. D., Hodgson L., Dogterom M., Derksen P. W. B., and Akhmanova A., “ Mesenchymal cell invasion requires cooperative regulation of persistent microtubule growth by SLAIN2 and CLASP1,” Dev. Cell. 39, 708–723 (2016). 10.1016/j.devcel.2016.11.009 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82. van Doorn J. M., Lageschaar L., Sprakel J., and van der Gucht J., “ Criticality and mechanical enhancement in composite fiber networks,” Phys. Rev. E. 95, 042503 (2017). 10.1103/PhysRevE.95.042503 [DOI] [PubMed] [Google Scholar]
- 83. Chen S., Markovich T., and MacKintosh F. C., “ Nonaffine deformation of semiflexible polymer and fiber networks,” Phys. Rev. Lett. 130, 088101 (2023). 10.1103/PhysRevLett.130.088101 [DOI] [PubMed] [Google Scholar]
- 84. Chang Y.-C., Nalbant P., Birkenfeld J., Chang Z.-F., and Bokoch G. M., “ GEF-H1 couples nocodazole-induced microtubule disassembly to cell contractility via RhoA,” Mol. Biol. Cell 19, 2147–2153 (2008). 10.1091/mbc.e07-12-1269 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85. Latorre E., Kale S., Casares L., Gómez-González M., Uroz M., Valon L., Nair R. V., Garreta E., Montserrat N., Del Campo A., Ladoux B., Arroyo M., and Trepat X., “ Active superelasticity in three-dimensional epithelia of controlled shape,” Nature 563, 203–208 (2018). 10.1038/s41586-018-0671-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86. Prechova M., Adamova Z., Schweizer A.-L., Maninova M., Bauer A., Kah D., Meier-Menches S. M., Wiche G., Fabry B., and Gregor M., “ Plectin-mediated cytoskeletal crosstalk controls cell tension and cohesion in epithelial sheets,” J. Cell Biol. 221, 202105146 (2022). 10.1083/jcb.202105146 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87. Das M. and Mackintosh F. C., “ Poisson's ratio in composite elastic media with rigid rods,” Phys. Rev. Lett. 105, 138102 (2010). 10.1103/PhysRevLett.105.138102 [DOI] [PubMed] [Google Scholar]
- 88. Dogterom M. and Yurke B., “ Measurement of the force-velocity relation for growing microtubules,” Science 278, 856–860 (1997). 10.1126/science.278.5339.856 [DOI] [PubMed] [Google Scholar]
- 89. Brangwynne C. P., MacKintosh F. C., Kumar S., Geisse N. A., Talbot J., Mahadevan L., Parker K. K., Ingber D. E., and Weitz D. A., “ Microtubules can bear enhanced compressive loads in living cells because of lateral reinforcement,” J. Cell Biol. 173, 733–741 (2006). 10.1083/jcb.200601060 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90. Robison P., Caporizzo M. A., Ahmadzadeh H., Bogush A. I., Chen C. Y., Margulies K. B., Shenoy V. B., and Prosser B. L., “ Detyrosinated microtubules buckle and bear load in contracting cardiomyocytes,” Science 352, aaf0659 (2016). 10.1126/science.aaf0659 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91. Wang N., Naruse K., Stamenović D., Fredberg J. J., Mijailovich S. M., Tolić-Nørrelykke I. M., Polte T., Mannix R., and Ingber D. E., “ Mechanical behavior in living cells consistent with the tensegrity model,” Proc. Natl. Acad. Sci. U. S. A. 98, 7765–7770 (2001). 10.1073/pnas.141199598 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92. Wu P. H., Aroush D. R. B., Asnacios A., Chen W. C., Dokukin M. E., Doss B. L., Durand-Smet P., Ekpenyong A., Guck J., Guz N. V., Janmey P. A., Lee J. S., Moore N. M., Ott A., Poh Y. C., Ros R., Sander M., Sokolov I., Staunton J. R., Wang N., Whyte G., and Wirtz D., “ A comparison of methods to assess cell mechanical properties,” Nat. Methods 15, 491–498 (2018). 10.1038/s41592-018-0015-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93. Block J., Witt H., Candelli A., Danes J. C., Peterman E. J., Wuite G. J., Janshoff A., and Köster S., “ Viscoelastic properties of vimentin originate from nonequilibrium conformational changes,” Sci. Adv. 4, aat1161 (2018). 10.1126/sciadv.aat1161 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94. van Mameren J., Vermeulen K. C., Gittes F., and Schmidt C. F., “ Leveraging single protein polymers to measure flexural rigidity,” J. Phys. Chem. B 113, 3837–3844 (2009). 10.1021/jp808328a [DOI] [PubMed] [Google Scholar]
- 95. Yasuda R., Miyata H., and K. Kinosita, Jr. , “ Direct measurement of the torsional rigidity of single actin filaments,” J. Mol. Biol. 263, 227–236 (1996). 10.1006/jmbi.1996.0571 [DOI] [PubMed] [Google Scholar]
- 96. Bibeau J. P., Pandit N. G., Gray S., Shatery Nejad N., Sindelar C. V., Cao W., and De La Cruz E. M., “ Twist response of actin filaments,” Proc. Natl. Acad. Sci. U. S. A. 120, e2208536120 (2023). 10.1073/pnas.2208536120 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97. Kreplak L., Bär H., Leterrier J. F., Herrmann H., and Aebi U., “ Exploring the mechanical behavior of single intermediate filaments,” J. Mol. Biol. 354, 569–577 (2005). 10.1016/j.jmb.2005.09.092 [DOI] [PubMed] [Google Scholar]
- 98. Kis A., Kasas S., Babić B., Kulik A. J., Benoît W., Briggs G. A. D., Schönenberger C., Catsicas S., and Forró L., “ Nanomechanics of microtubules,” Phys. Rev. Lett. 89, 248101 (2002). 10.1103/PhysRevLett.89.248101 [DOI] [PubMed] [Google Scholar]
- 99. Schaedel L., John K., Gaillard J., Nachury M. V., Blanchoin L., and Théry M., “ Microtubules self-repair in response to mechanical stress,” Nat. Mater. 14, 1156–1163 (2015). 10.1038/nmat4396 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100. Tsuda Y., Yasutake H., Ishijima A., and Yanagida T., “ Torsional rigidity of single actin filaments and actin-actin bond breaking force under torsion measured directly by in vitro micromanipulation,” Proc. Natl. Acad. Sci. U. S. A. 93, 12937–12942 (1996). 10.1073/pnas.93.23.12937 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101. Endow S. A. and Marszalek P. E., “ An estimate to the first approximation of microtubule rupture force,” Eur. Biophys. J. 48, 569–577 (2019). 10.1007/s00249-019-01371-6 [DOI] [PubMed] [Google Scholar]
- 102. Lorenz C., Forsting J., Style R. W., Klumpp S., and Köster S., “ Keratin filament mechanics and energy dissipation are determined by metal-like plasticity,” Matter 6, 2019–2033 (2023). 10.1016/j.matt.2023.04.014 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103. Nunes Vicente F., Lelek M., Tinevez J.-Y., Tran Q. D., Pehau-Arnaudet G., Zimmer C., Etienne-Manneville S., Giannone G., and Leduc C., “ Molecular organization and mechanics of single vimentin filaments revealed by super-resolution imaging,” Sci. Adv. 8, eabm2696 (2022). 10.1126/sciadv.abm2696 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104. Gardel M. L., Kasza K. E., Brangwynne C. P., Liu J., and Weitz D. A., “ Chapter 19: Mechanical response of cytoskeletal networks,” Methods Cell Biol. 89, 487–519 (2008). 10.1016/S0091-679X(08)00619-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105. Weihs D., Mason T. G., and Teitell M. A., “ Bio-microrheology: A frontier in microrheology,” Biophys. J. 91, 4296–4305 (2006). 10.1529/biophysj.106.081109 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106. Yang Y., Bai M., Klug W. S., Levine A. J., and Valentine M. T., “ Microrheology of highly crosslinked microtubule networks is dominated by force-induced crosslinker unbinding,” Soft Matter 9, 383–393 (2013). 10.1039/C2SM26934A [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107. Lin Y. C., Koenderink G. H., MacKintosh F. C., and Weitz D. A., “ Viscoelastic properties of microtubule networks,” Macromolecules 40, 7714–7720 (2007). 10.1021/ma070862l [DOI] [Google Scholar]
- 108. Yang Y., Lin J., Kaytanli B., Saleh O. A., and Valentine M. T., “ Direct correlation between creep compliance and deformation in entangled and sparsely crosslinked microtubule networks,” Soft Matter 8, 1776–1784 (2012). 10.1039/C2SM06745E [DOI] [Google Scholar]
- 109. Kirchenbuechler I., Guu D., Kurniawan N. A., Koenderink G. H., and Lettinga M. P., “ Direct visualization of flow-induced conformational transitions of single actin filaments in entangled solutions,” Nat. Commun. 5, 5060 (2014). 10.1038/ncomms6060 [DOI] [PubMed] [Google Scholar]
- 110. Gardel M. L., Shin J. H., MacKintosh F. C., Mahadevan L., Matsudaira P., and Weitz D. A., “ Elastic behavior of cross-linked and bundled actin networks,” Science 304, 1301–1305 (2004). 10.1126/science.1095087 [DOI] [PubMed] [Google Scholar]
- 111. MacKintosh F. C., Käs J., and Janmey P. A., “ Elasticity of semiflexible biopolymer networks,” Phys. Rev. Lett. 75, 4425–4428 (1995). 10.1103/PhysRevLett.75.4425 [DOI] [PubMed] [Google Scholar]
- 112. Kasza K. E., Broedersz C. P., Koenderink G. H., Lin Y. C., Messner W., Millman E. A., Nakamura F., Stossel T. P., MacKintosh F. C., and Weitz D. A., “ Actin filament length tunes elasticity of flexibly cross-linked actin networks,” Biophys. J. 99, 1091–1100 (2010). 10.1016/j.bpj.2010.06.025 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113. Broedersz C. P., Kasza K. E., Jawerth L. M., Münster S., Weitz D. A., and MacKintosh F. C., “ Measurement of nonlinear rheology of cross-linked biopolymer gels,” Soft Matter 6, 4120–4127 (2010). 10.1039/c0sm00285b [DOI] [Google Scholar]
- 114. Wilhelm J. and Frey E., “ Elasticity of stiff polymer networks,” Phys. Rev. Lett. 91, 108103 (2003). 10.1103/PhysRevLett.91.108103 [DOI] [PubMed] [Google Scholar]
- 115. Head D. A., Levine A. J., and MacKintosh F. C., “ Deformation of cross-linked semiflexible polymer networks,” Phys. Rev. Lett. 91, 108102 (2003). 10.1103/PhysRevLett.91.108102 [DOI] [PubMed] [Google Scholar]
- 116. Chaudhuri O., Parekh S. H., and Fletcher D. A., “ Reversible stress softening of actin networks,” Nature 445, 295–298 (2007). 10.1038/nature05459 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117. Lin Y. C., Yao N. Y., Broedersz C. P., Herrmann H., MacKintosh F. C., and Weitz D. A., “ Origins of elasticity in intermediate filament networks,” Phys. Rev. Lett. 104, 058101 (2010). 10.1103/PhysRevLett.104.058101 [DOI] [PubMed] [Google Scholar]
- 118. Yao N. Y., Braedersz C. P., Lin Y. C., Kasza K. E., MacKintosh F. C., and Weitz D. A., “ Elasticity in ionically cross-linked neurofilament networks,” Biophys. J. 98, 2147–2153 (2010). 10.1016/j.bpj.2010.01.062 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119. Lin Y. C., Broedersz C. P., Rowat A. C., Wedig T., Herrmann H., MacKintosh F. C., and Weitz D. A., “ Divalent cations crosslink vimentin intermediate filament tail domains to regulate network mechanics,” J. Mol. Biol. 399, 637–644 (2010). 10.1016/j.jmb.2010.04.054 [DOI] [PubMed] [Google Scholar]
- 120. Aufderhorst-Roberts A. and Koenderink G. H., “ Stiffening and inelastic fluidization in vimentin intermediate filament networks,” Soft Matter 15, 7127–7136 (2019). 10.1039/C9SM00590K [DOI] [PubMed] [Google Scholar]
- 121. Bär H., Schopferer M., Sharma S., Hochstein B., Mücke N., Herrmann H., and Willenbacher N., “ Mutations in desmin's carboxy-terminal “tail” domain severely modify filament and network mechanics,” J. Mol. Biol. 397, 1188–1198 (2010). 10.1016/j.jmb.2010.02.024 [DOI] [PubMed] [Google Scholar]
- 122. Wu H., Shen Y., Wang D., Herrmann H., Goldman R. D., and Weitz D. A., “ Effect of divalent cations on the structure and mechanics of vimentin intermediate filaments,” Biophys. J. 119, 55–64 (2020). 10.1016/j.bpj.2020.05.016 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123. Fukuyama K., Murozuka T., Caldwell R., and Epstein W. L., “ Divalent cation stimulation of in vitro fibre assembly from epidermal keratin protein,” J. Cell Sci. 33, 255–263 (1978). 10.1242/jcs.33.1.255 [DOI] [PubMed] [Google Scholar]
- 124. Schepers A. V., Lorenz C., Nietmann P., Janshoff A., Klumpp S., and Köster S., “ Multiscale mechanics and temporal evolution of vimentin intermediate filament networks,” Proc. Natl. Acad. Sci. U. S. A. 118, 2102026118 (2021). 10.1073/pnas.2102026118 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125. Wagner O. I., Rammensee S., Korde N., Wen Q., Leterrier J. F., and Janmey P. A., “ Softness, strength and self-repair in intermediate filament networks,” Exp. Cell Res. 313, 2228 (2007). 10.1016/j.yexcr.2007.04.025 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126. Denz M., Marschall M., Herrmann H., and Köster S., “ Ion type and valency differentially drive vimentin tetramers into intermediate filaments or higher order assemblies,” Soft Matter 17, 870–878 (2021). 10.1039/D0SM01659D [DOI] [PubMed] [Google Scholar]
- 127. Golde T., Huster C., Glaser M., Händler T., Herrmann H., Käs J. A., and Schnauß J., “ Glassy dynamics in composite biopolymer networks,” Soft Matter 14, 7970–7978 (2018). 10.1039/C8SM01061G [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128. Deek J., Maan R., Loiseau E., and Bausch A. R., “ Reconstitution of composite actin and keratin networks in vesicles,” Soft Matter 14, 1897–1902 (2018). 10.1039/C7SM00819H [DOI] [PubMed] [Google Scholar]
- 129. Elbalasy I., Mollenkopf P., Tutmarc C., Herrmann H., and Schnauß J., “ Keratins determine network stress responsiveness in reconstituted actin-keratin filament systems,” Soft Matter 17, 3954–3962 (2021). 10.1039/D0SM02261F [DOI] [PubMed] [Google Scholar]
- 130. Pelletier V., Gal N., Fournier P., and Kilfoil M. L., “ Microrheology of microtubule solutions and actin-microtubule composite networks,” Phys. Rev. Lett. 102, 188303 (2009). 10.1103/PhysRevLett.102.188303 [DOI] [PubMed] [Google Scholar]
- 131. Piechocka I., “ Biopolymers: From structural hierarchy to nonlinear rheology,” Ph.D. thesis ( Vrije Universiteit Amsterdam, 2011). [Google Scholar]
- 132. Esue O., Carson A. A., Tseng Y., and Wirtz D., “ A direct interaction between actin and vimentin filaments mediated by the tail domain of vimentin,” J. Biol. Chem. 281, 30393–30399 (2006). 10.1074/jbc.M605452200 [DOI] [PubMed] [Google Scholar]
- 133. Schaedel L., Lorenz C., Schepers A. V., Klumpp S., and Köster S., “ Vimentin intermediate filaments stabilize dynamic microtubules by direct interactions,” Nat. Commun. 12, 3799 (2021). 10.1038/s41467-021-23523-z [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134. Jensen M. H., Morris E. J., Goldman R. D., and Weitz D. A., “ Emergent properties of composite semiflexible biopolymer networks,” BioArchitecture 4, 138–143 (2014). 10.4161/19490992.2014.989035 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135. Lin Y.-C., Koenderink G. H., MacKintosh F. C., and Weitz D. A., “ Control of non-linear elasticity in F-actin networks with microtubules,” Soft Matter 7, 902–906 (2011). 10.1039/C0SM00478B [DOI] [Google Scholar]
- 136. Ricketts S. N., Ross J. L., and Robertson-Anderson R. M., “ Co-entangled actin-microtubule composites exhibit tunable stiffness and power-law stress relaxation,” Biophys. J. 115, 1055–1067 (2018). 10.1016/j.bpj.2018.08.010 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137. Ricketts S. N., Francis M. L., Farhadi L., Rust M. J., Das M., Ross J. L., and Robertson-Anderson R. M., “ Varying crosslinking motifs drive the mesoscale mechanics of actin-microtubule composites,” Sci. Rep. 9, 12831 (2019). 10.1038/s41598-019-49236-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138. Shen Y., Wu H., Lu P. J., Wang D., Shayegan M., Li H., Shi W., Wang Z., Cai L. H., Xia J., Zhang M., Ding R., Herrmann H., Goldman R., Mackintosh F. C., Moncho-Jordá A., and Weitz D. A., “ Effects of vimentin intermediate filaments on the structure and dynamics of in vitro multicomponent interpenetrating cytoskeletal networks,” Phys. Rev. Lett. 127, 108101 (2021). 10.1103/PhysRevLett.127.108101 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139. Wu H., Shen Y., Sivagurunathan S., Weber M. S., Adam S. A., Shin J. H., Fredberg J. J., Medalia O., Goldman R., Weitz D. A., and Lippincott-Schwartz J., “ Vimentin intermediate filaments and filamentous actin form unexpected interpenetrating networks that redefine the cell cortex,” Proc. Natl. Acad. Sci. U. S. A. 119, 2115217119 (2022). 10.1073/pnas.2115217119 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140. Preciado López M., Huber F., Grigoriev I., Steinmetz M. O., Akhmanova A., Koenderink G. H., and Dogterom M., “ Actin-microtubule coordination at growing microtubule ends,” Nat. Commun. 5, 4778 (2014). 10.1038/ncomms5778 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141. Petitjean I. I., Tran Q. D., Goutou A., Kabir Z., Wiche G., Leduc C., and Koenderink G. H., “ Reconstitution of cytolinker-mediated crosstalk between actin and vimentin,” Eur. J. Cell Biol. 103, 151403 (2024). 10.1016/j.ejcb.2024.151403 [DOI] [PubMed] [Google Scholar]
- 142. SenGupta S., Parent C. A., and Bear J. E., “ The principles of directed cell migration,” Nat. Rev. Mol. Cell Biol. 22, 529–547 (2021). 10.1038/s41580-021-00366-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143. Petrie R. J. and Yamada K. M., “ Multiple mechanisms of 3D migration: The origins of plasticity,” Curr. Opin. Cell Biol. 42, 7–12 (2016). 10.1016/j.ceb.2016.03.025 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 144. Palecek S. P., Huttenlocher A., Horwitz A. F., and Lauffenburger D. A., “ Physical and biochemical regulation of integrin release during rear detachment of migrating cells,” J. Cell Sci. 111, 929–940 (1998). 10.1242/jcs.111.7.929 [DOI] [PubMed] [Google Scholar]
- 145. Style R. W., Boltyanskiy R., German G. K., Hyland C., Macminn C. W., Mertz A. F., Wilen L. A., Xu Y., and Dufresne E. R., “ Traction force microscopy in physics and biology,” Soft Matter 10, 4047–4055 (2014). 10.1039/c4sm00264d [DOI] [PubMed] [Google Scholar]
- 146. Petrie R. J., Koo H., and Yamada K. M., “ Generation of compartmentalized pressure by a nuclear piston governs cell motility in a 3D matrix,” Science 345, 1062–1065 (2014). 10.1126/science.1256965 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147. Paluch E. K. and Raz E., “ The role and regulation of blebs in cell migration,” Curr. Opin. Cell Biol. 25, 582–590 (2013). 10.1016/j.ceb.2013.05.005 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148. Liu Y.-J., Berre M. L., Lautenschlaeger F., Maiuri P., Callan-Jones A., Heuzé M., Takaki T., Voituriez R., and Piel M., “ Confinement and low adhesion induce fast amoeboid migration of slow mesenchymal cells,” Cell 160, 659–672 (2015). 10.1016/j.cell.2015.01.007 [DOI] [PubMed] [Google Scholar]
- 149. Stroka K. M., Jiang H., Chen S.-H., Tong Z., Wirtz D., Sun S. X., and Konstantopoulos K., “ Water permeation drives tumor cell migration in confined microenvironments,” Cell 157, 611–623 (2014). 10.1016/j.cell.2014.02.052 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150. Krendel M., Zenke F. T., and Bokoch G. M., “ Nucleotide exchange factor GEF-H1 mediates cross-talk between microtubules and the actin cytoskeleton,” Nat. Cell Biol. 4, 294–301 (2002). 10.1038/ncb773 [DOI] [PubMed] [Google Scholar]
- 151. Rape A., Guo W. H., and Wang Y. L., “ Microtubule depolymerization induces traction force increase through two distinct pathways,” J. Cell Sci. 124, 4233–4240 (2011). 10.1242/jcs.090563 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152. Azoitei M. L., Noh J., Marston D. J., Roudot P., Marshall C. B., Daugird T. A., Lisanza S. L., Sandí M.-J., Ikura M., Sondek J., Rottapel R., Hahn K. M., and Danuser G., “ Spatiotemporal dynamics of GEF-H1 activation controlled by microtubule- and Src-mediated pathways,” J. Cell Biol. 218, 3077–3097 (2019). 10.1083/jcb.201812073 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153. Heck J. N., Ponik S. M., Garcia-Mendoza M. G., Pehlke C. A., Inman D. R., Eliceiri K. W., and Keely P. J., “ Microtubules regulate GEF-H1 in response to extracellular matrix stiffness,” Mol. Biol. Cell 23, 2583–2592 (2012). 10.1091/mbc.e11-10-0876 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154. Kopf A., Renkawitz J., Hauschild R., Girkontaite I., Tedford K., Merrin J., Thorn-Seshold O., Trauner D., Häcker H., Fischer K. D., Kiermaier E., and Sixt M., “ Microtubules control cellular shape and coherence in amoeboid migrating cells,” J. Cell Biol. 219, e201907154 (2020). 10.1083/jcb.201907154 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155. Jiu Y., Peränen J., Schaible N., Cheng F., Eriksson J. E., Krishnan R., and Lappalainen P., “ Vimentin intermediate filaments control actin stress fiber assembly through GEF-H1 and RhoA,” J. Cell Sci. 130, 892–902 (2017). 10.1242/jcs.196881 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156. Gan Z., Ding L., Burckhardt C. J., Lowery J., Zaritsky A., Sitterley K., Mota A., Costigliola N., Starker C. G., Voytas D. F., Tytell J., Goldman R. D., and Danuser G., “ Vimentin intermediate filaments template microtubule networks to enhance persistence in cell polarity and directed migration,” Cell Syst. 3, 252–263.e8 (2016). 10.1016/j.cels.2016.08.007 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157. Jiu Y., Lehtimäki J., Tojkander S., Cheng F., Jäälinoja H., Liu X., Varjosalo M., Eriksson J. E., and Lappalainen P., “ Bidirectional interplay between vimentin intermediate filaments and contractile actin stress fibers,” Cell Rep. 11, 1511–1518 (2015). 10.1016/j.celrep.2015.05.008 [DOI] [PubMed] [Google Scholar]
- 158. Marks P. C., Hewitt B. R., Baird M. A., Wiche G., and Petrie R. J., “ Plectin linkages are mechanosensitive and required for the nuclear piston mechanism of three-dimensional cell migration,” Mol. Biol. Cell 33, ar104 (2022). 10.1091/mbc.E21-08-0414 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159. Daday C., Kolšek K., and Gräter F., “ The mechano-sensing role of the unique SH3 insertion in plakin domains revealed by molecular dynamics simulations,” Sci. Rep. 7, 11669 (2017). 10.1038/s41598-017-11017-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160. Murrell M. P. and Gardel M. L., “ F-actin buckling coordinates contractility and severing in a biomimetic actomyosin cortex,” Proc. Natl. Acad. Sci. U. S. A. 109, 20820–20825 (2012). 10.1073/pnas.1214753109 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 161. Wollrab V., Belmonte J. M., Baldauf L., Leptin M., Nédeléc F., and Koenderink G. H., “ Polarity sorting drives remodeling of actin-myosin networks,” J. Cell Sci. 132, jcs219717 (2018). 10.1242/jcs.219717 [DOI] [PubMed] [Google Scholar]
- 162. Alvarado J., Sheinman M., Sharma A., MacKintosh F. C., and Koenderink G. H., “ Force percolation of contractile active gels,” Soft Matter 13, 5624–5644 (2017). 10.1039/C7SM00834A [DOI] [PubMed] [Google Scholar]
- 163. Bendix P. M., Koenderink G. H., Cuvelier D., Dogic Z., Koeleman B. N., Brieher W. M., Field C. M., Mahadevan L., and Weitz D. A., “ A quantitative analysis of contractility in active cytoskeletal protein networks,” Biophys. J. 94, 3126–3136 (2008). 10.1529/biophysj.107.117960 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164. Carvalho K., Tsai F. C., Lees E., Voituriez R., Koenderink G. H., and Sykes C., “ Cell-sized liposomes reveal how actomyosin cortical tension drives shape change,” Proc. Natl. Acad. Sci. U. S. A. 110, 16456–16461 (2013). 10.1073/pnas.1221524110 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 165. Loiseau E., Schneider J. A. M., Keber F. C., Pelzl C., Massiera G., Salbreux G., and Bausch A. R., “ Shape remodeling and blebbing of active cytoskeletal vesicles,” Sci. Adv. 2, e1500465 (2016). 10.1126/sciadv.1500465 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166. McFadden W. M., McCall P. M., Gardel M. L., and Munro E. M., “ Filament turnover tunes both force generation and dissipation to control long-range flows in a model actomyosin cortex,” PLoS Comput. Biol. 13, e1005811 (2017). 10.1371/journal.pcbi.1005811 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 167. Vogel S. K., Wölfer C., Ramirez-Diaz D. A., Flassig R. J., Sundmacher K., and Schwille P., “ Symmetry breaking and emergence of directional flows in minimal actomyosin cortices,” Cells 9, 1432 (2020). 10.3390/cells9061432 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168. Sonal, Ganzinger K. A., Vogel S. K., Mücksch J., Blumhardt P., and Schwille P., “ Myosin-II activity generates a dynamic steady state with continuous actin turnover in a minimal actin cortex,” J. Cell Sci. 132, jcs219899 (2019). 10.1242/jcs.219899 [DOI] [PubMed] [Google Scholar]
- 169. Han Tan T., Malik-Garbi M., Abu-Shah E., Li J., Sharma A., MacKintosh F. C., Keren K., Schmidt C. F., and Fakhri N., “ Self-organized stress patterns drive state transitions in actin cortices,” Sci. Adv. 4, eaar2847 (2018). 10.1126/sciadv.aar2847 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 170. Pinot M., Steiner V., Dehapiot B., Yoo B.-K., Chesnel F., Blanchoin L., Kervrann C., and Gueroui Z., “ Confinement induces actin flow in a meiotic cytoplasm,” Proc. Natl. Acad. Sci. U. S. A. 109, 11705–11710 (2012). 10.1073/pnas.1121583109 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 171. Malik-Garbi M., Ierushalmi N., Jansen S., Abu-Shah E., Goode B. L., Mogilner A., and Keren K., “ Scaling behaviour in steady-state contracting actomyosin networks,” Nat. Phys. 15, 509–516 (2019). 10.1038/s41567-018-0413-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 172. Sakamoto R., Izri Z., Shimamoto Y., Miyazaki M., and Maeda Y. T., “ Geometric trade-off between contractile force and viscous drag determines the actomyosin-based motility of a cell-sized droplet,” Proc. Natl. Acad. Sci. U. S. A. 119, e2121147119 (2022). 10.1073/pnas.2121147119 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173. Lee G., Leech G., Rust M. J., Das M., Mcgorty R. J., Ross J. L., and Robertson-Anderson R. M., “ Myosin-driven actin-microtubule networks exhibit self-organized contractile dynamics,” Sci. Adv. 7, eabe4334 (2021). 10.1126/sciadv.abe4334 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 174. Cramer L. P., “ Forming the cell rear first: Breaking cell symmetry to trigger directed cell migration,” Nat. Cell Biol. 12, 628–632 (2010). 10.1038/ncb0710-628 [DOI] [PubMed] [Google Scholar]
- 175. Vasiliev J. M., Gelfand M., Domnina L. V., 1vanova O. Y., Komm S. G., and Olshevskaja L. V., “ Effect of colcemid on the locomotory behaviour of fibroblasts,” Development 24, 625–640 (1970). 10.1242/dev.24.3.625 [DOI] [PubMed] [Google Scholar]
- 176. Efimova N., Yang C., Chia J. X., Li N., Lengner C. J., Neufeld K. L., and Svitkina T. M., “ Branched actin networks are assembled on microtubules by adenomatous polyposis coli for targeted membrane protrusion,” J. Cell Biol. 219, e202003091 (2020). 10.1083/jcb.202003091 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177. Henty-Ridilla J. L., Rankova A., Eskin J. A., Kenny K., and Goode B. L., “ Accelerated actin filament polymerization from microtubule plus ends,” Science 352, 1004–1009 (2016). 10.1126/science.aaf1709 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 178. Kodama A., Karakesisoglou I., Wong E., Vaezi A., and Fuchs E., “ ACF7: An essential integrator of microtubule dynamics,” Cell 115, 343–354 (2003). 10.1016/S0092-8674(03)00813-4 [DOI] [PubMed] [Google Scholar]
- 179. Drabek K., van Ham M., Stepanova T., Draegestein K., van Horssen R., Sayas C. L., Akhmanova A., Hagen T. t, Smits R., Fodde R., Grosveld F., and Galjart N., “ Role of CLASP2 in microtubule stabilization and the regulation of persistent motility,” Curr. Biol. 16, 2259–2264 (2006). 10.1016/j.cub.2006.09.065 [DOI] [PubMed] [Google Scholar]
- 180. Bouchet B. P., Gough R. E., Ammon Y.-C., van de Willige D., Post H., Jacquemet G., Altelaar A. M., A. Heck, Jr. , Goult B. T., and Akhmanova A., “ Talin-KANK1 interaction controls the recruitment of cortical microtubule stabilizing complexes to focal adhesions,” eLife 5, e18124 (2016). 10.7554/eLife.18124 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 181. Infante E., Castagnino A., Ferrari R., Monteiro P., Agüera-González S., Paul-Gilloteaux P., Domingues M. J., Maiuri P., Raab M., Shanahan C. M., Baffet A., Piel M., Gomes E. R., and Chavrier P., “ LINC complex-Lis1 interplay controls MT1-MMP matrix digest-on-demand response for confined tumor cell migration,” Nat. Commun. 9, 2443 (2018). 10.1038/s41467-018-04865-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 182. Helfand B. T., Mendez M. G., Murthy S. N., Shumaker D. K., Grin B., Mahammad S., Aebi U., Wedig T., Wu Y. I., Hahn K. M., Inagaki M., Herrmann H., and Goldman R. D., “ Vimentin organization modulates the formation of lamellipodia,” Mol. Biol. Cell 22, 1274–1289 (2011). 10.1091/mbc.e10-08-0699 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 183. Sakamoto Y., Boëda B., and Etienne-Manneville S., “ APC binds intermediate filaments and is required for their reorganization during cell migration,” J. Cell Biol. 200, 249–258 (2013). 10.1083/jcb.201206010 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 184. Shabbir S. H., Cleland M. M., Goldman R. D., and Mrksich M., “ Geometric control of vimentin intermediate filaments,” Biomaterials 35, 1359–1366 (2014). 10.1016/j.biomaterials.2013.10.008 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 185. Leduc C. and Etienne-Manneville S., “ Regulation of microtubule-associated motors drives intermediate filament network polarization,” J. Cell Biol. 216, 1689–1703 (2017). 10.1083/jcb.201607045 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 186. Costigliola N., Ding L., Burckhardt C. J., Han S. J., Gutierrez E., Mota A., Groisman A., Mitchison T. J., and Danuser G., “ Vimentin fibers orient traction stress,” Proc. Natl. Acad. Sci. U. S. A. 114, 5195–5200 (2017). 10.1073/pnas.1614610114 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 187. Leube R. E., Moch M., and Windoffer R., “ Intermediate filaments and the regulation of focal adhesion,” Curr. Opin. Cell Biol. 32, 13–20 (2015). 10.1016/j.ceb.2014.09.011 [DOI] [PubMed] [Google Scholar]
- 188. Preciado López M., Huber F., Grigoriev I., Steinmetz M. O., Akhmanova A., Dogterom M., and Koenderink G. H., “ In vitro reconstitution of dynamic microtubules interacting with actin filament networks,” Methods Enzymol. 540, 301–320 (2014). 10.1016/B978-0-12-397924-7.00017-0 [DOI] [PubMed] [Google Scholar]
- 189. Inoue D., Obino D., Pineau J., Farina F., Gaillard J., Guerin C., Blanchoin L., Lennon-Duménil A.-M., and Théry M., “ Actin filaments regulate microtubule growth at the centrosome,” EMBO J. 38, e99630 (2019). 10.15252/embj.201899630 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 190. Gélin M., Schaeffer A., Gaillard J., Guérin C., Vianay B., Orhant-Prioux M., Braun M., Leterrier C., Blanchoin L., and Théry M., “ Microtubules under mechanical pressure can breach dense actin networks,” J. Cell Sci. 136, jcs261667 (2023). 10.1242/jcs.261667 [DOI] [PubMed] [Google Scholar]
- 191. van de Willige D., Hummel J. J., Alkemade C., Kahn O. I., Au F. K., Qi R. Z., Dogterom M., Koenderink G. H., Hoogenraad C. C., and Akhmanova A., “ Cytolinker Gas2L1 regulates axon morphology through microtubule-modulated actin stabilization,” EMBO Rep. 20, e47732 (2019). 10.15252/embr.201947732 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 192. Elie A., Prezel E., Guérin C., Denarier E., Ramirez-Rios S., Serre L., Andrieux A., Fourest-Lieuvin A., Blanchoin L., and Arnal I., “ Tau co-organizes dynamic microtubule and actin networks,” Sci. Rep. 5, 9964 (2015). 10.1038/srep09964 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193. Alkemade C., “ Reconstituting Microtubule-Actin Coordination by Cytolinkers,” Ph.D. thesis ( Delft University of Technology, 2021). [Google Scholar]
- 194. Rodgers N. C., Lawrence E. J., Sawant A. V., Efimova N., Gonzalez-Vasquez G., Hickman T. T., Kaverina I., and Zanic M., “ CLASP2 facilitates dynamic actin filament organization along the microtubule lattice,” Mol. Biol. Cell 34, br3 (2023). 10.1091/mbc.E22-05-0149 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 195. Alkemade C., Wierenga H., Volkov V. A., López M. P., Akhmanova A., ten Wolde P. R., Dogterom M., and Koenderink G. H., “ Cross-linkers at growing microtubule ends generate forces that drive actin transport,” Proc. Natl. Acad. Sci. U. S. A. 119, e2112799119 (2022). 10.1073/pnas.2112799119 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196. Vishwakarma M., Spatz J. P., and Das T., “ Mechanobiology of leader-follower dynamics in epithelial cell migration,” Curr. Opin. Cell Biol. 66, 97–103 (2020). 10.1016/j.ceb.2020.05.007 [DOI] [PubMed] [Google Scholar]
- 197. Haeger A., Wolf K., Zegers M. M., and Friedl P., “ Collective cell migration: Guidance principles and hierarchies,” Trends Cell Biol. 25, 556–566 (2015). 10.1016/j.tcb.2015.06.003 [DOI] [PubMed] [Google Scholar]
- 198. Grosser S., Lippoldt J., Oswald L., Merkel M., Sussman D. M., Renner F., Gottheil P., Morawetz E. W., Fuhs T., Xie X., Pawlizak S., Fritsch A. W., Wolf B., Horn L.-C., Briest S., Aktas B., Manning M. L., and Käs J. A., “ Cell and nucleus shape as an indicator of tissue fluidity in carcinoma,” Phys. Rev. X 11, 011033 (2021). 10.1103/PhysRevX.11.011033 [DOI] [Google Scholar]
- 199. Park J.-A., Hun Kim J., Bi D., Mitchel J. A., Taheri Qazvini N., Tantisira K., Park C. Y., McGill M., Kim S.-H., Gweon B., Notbohm J., R. Steward, Jr. , Burger S., Randell S. H., Kho A. T., Tambe D. T., Hardin C., Shore S. A., Israel E., Weitz D. A., Tschumperlin D. J., Henske E. P., Weiss S. T., Manning M. L., Butler J. P., Drazen J. M., and Fredberg J. J., “ Unjamming and cell shape in the asthmatic airway epithelium guided data interpretation on the biological relevance of cellular migration,” Nat. Mater. 14, 1040–1048 (2015). 10.1038/nmat4357 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 200. Ilina O., Gritsenko P. G., Syga S., Lippoldt J., La Porta C. A., Chepizhko O., Grosser S., Vullings M., Bakker G. J., Starruß J., Bult P., Zapperi S., Käs J. A., Deutsch A., and Friedl P., “ Cell–cell adhesion and 3D matrix confinement determine jamming transitions in breast cancer invasion,” Nat. Cell Biol. 22, 1103–1115 (2020). 10.1038/s41556-020-0552-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 201. Mongera A., Rowghanian P., Gustafson H. J., Shelton E., Kealhofer D. A., Carn E. K., Serwane F., Lucio A. A., Giammona J., and Campàs O., “ A fluid-to-solid jamming transition underlies vertebrate body axis elongation,” Nature 561, 401–405 (2018). 10.1038/s41586-018-0479-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 202. Palamidessi A., Malinverno C., Frittoli E., Corallino S., Barbieri E., Sigismund S., Beznoussenko G. V., Martini E., Garre M., Ferrara I., Tripodo C., Ascione F., Cavalcanti-Adam E. A., Li Q., Di Fiore P. P., Parazzoli D., Giavazzi F., Cerbino R., and Scita G., “ Unjamming overcomes kinetic and proliferation arrest in terminally differentiated cells and promotes collective motility of carcinoma,” Nat. Mater. 18, 1252–1263 (2019). 10.1038/s41563-019-0425-1 [DOI] [PubMed] [Google Scholar]
- 203. Tambe D. T., Hardin C. C., Angelini T. E., Rajendran K., Park C. Y., Serra-Picamal X., Zhou E. H., Zaman M. H., Butler J. P., Weitz D. A., Fredberg J. J., and Trepat X., “ Collective cell guidance by cooperative intercellular forces,” Nat. Mater. 10, 469–475 (2011). 10.1038/nmat3025 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 204. Das T., Safferling K., Rausch S., Grabe N., Boehm H., and Spatz J. P., “ A molecular mechanotransduction pathway regulates collective migration of epithelial cells,” Nat. Cell Biol. 17, 276–287 (2015). 10.1038/ncb3115 [DOI] [PubMed] [Google Scholar]
- 205. Campàs O., Noordstra I., and Yap A., “ Adherens junctions as molecular regulators of emergent tissue mechanics,” Nat. Rev. Mol. Cell Biol. 25, 252–269 (2023). 10.1038/s41580-023-00688-7 [DOI] [PubMed] [Google Scholar]
- 206. Zhang J., Goliwas K. F., Wang W., Taufalele P. V., Bordeleau F., and Reinhart-King C. A., “ Energetic regulation of coordinated leader–follower dynamics during collective invasion of breast cancer cells,” Proc. Natl. Acad. Sci. U. S. A. 116, 7867–7872 (2019). 10.1073/pnas.1809964116 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 207. Broussard J. A., Jaiganesh A., Zarkoob H., Conway D. E., Dunn A. R., Espinosa H. D., Janmey P. A., and Green K. J., “ Scaling up single-cell mechanics to multicellular tissues—The role of the intermediate filament-desmosome network,” J. Cell Sci. 133, jcs228031 (2020). 10.1242/jcs.228031 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 208. Schmelz M. and Franke W. W., “ Complexus adhaerentes, a new group of desmoplakin-containing junctions in endothelial cells: The syndesmos connecting retothelial cells of lymph nodes,” Eur. J. Cell Biol. 61, 274–289 (1993). [PubMed] [Google Scholar]
- 209. Hegazy M., Perl A. L., Svoboda S. A., and Green K. J., “ Desmosomal cadherins in health and disease,” Annu. Rev. Pathol. 17, 47–72 (2022). 10.1146/annurev-pathol-042320-092912 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 210. Lewis J. E., Wahl J. K., Sass K. M., Jensen P. J., Johnson K. R., and Wheelock M. J., “ Cross-talk between adherens junctions and desmosomes depends on plakoglobin,” J. Cell Biol. 136, 919–934 (1997). 10.1083/jcb.136.4.919 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 211. Sadhanasatish T., Augustin K., Windgasse L., Chrostek-Grashoff A., Rief M., and Grashoff C., “ A molecular optomechanics approach reveals functional relevance of force transduction across talin and desmoplakin,” Sci. Adv. 9, eadg3347 (2023). 10.1126/sciadv.adg3347 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 212. Etienne-Manneville S., “ Actin and microtubules in cell motility: Which one is in control?” Traffic 5, 470 (2004). 10.1111/j.1600-0854.2004.00196.x [DOI] [PubMed] [Google Scholar]
- 213. Osmanagic-Myers S., Rus S., Wolfram M., Brunner D., Goldmann W. H., Bonakdar N., Fischer I., Reipert S., Zuzuarregui A., Walko G., and Wiche G., “ Plectin reinforces vascular integrity by mediating crosstalk between the vimentin and the actin networks,” J. Cell Sci. 128, 4138–4150 (2015). 10.1242/jcs.172056 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 214. Roberts B. J., Pashaj A., Johnson K. R., and Wahl III J. K., “ Desmosome dynamics in migrating epithelial cells requires the actin cytoskeleton,” Exp. Cell Res. 317, 2814–2822 (2011). 10.1016/j.yexcr.2011.09.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 215. Sonavane P. R., Wang C., Dzamba B., Weber G. F., Periasamy A., and DeSimone D. W., “ Mechanical and signaling roles for keratin intermediate filaments in the assembly and morphogenesis of Xenopus mesendoderm tissue at gastrulation,” Development 144, 4363–4376 (2017). 10.1242/dev.155200 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 216. De Pascalis C., Pérez-González C., Seetharaman S., Boëda B., Vianay B., Burute M., Leduc C., Borghi N., Trepat X., and Etienne-Manneville S., “ Intermediate filaments control collective migration by restricting traction forces and sustaining cell-cell contacts,” J. Cell Biol. 217, 3031–3044 (2018). 10.1083/jcb.201801162 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 217. Cheung K. J., Gabrielson E., Werb Z., and Ewald A. J., “ Collective invasion in breast cancer requires a conserved basal epithelial program,” Cell 155, 1639–1651 (2013). 10.1016/j.cell.2013.11.029 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 218. Yoon S. and Leube R. E., “ Keratin intermediate filaments: Intermediaries of epithelial cell migration,” Essays Biochem. 63, 521–533 (2019). 10.1042/EBC20190017 [DOI] [PubMed] [Google Scholar]
- 219. Fortier A. M., Asselin E., and Cadrin M., “ Keratin 8 and 18 loss in epithelial cancer cells increases collective cell migration and cisplatin sensitivity through claudin1 up-regulation,” J. Biol. Chem. 288, 11555–11571 (2013). 10.1074/jbc.M112.428920 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 220. Kölsch A., Windoffer R., and Leube R. E., “ Actin-dependent dynamics of keratin filament precursors,” Cell Motil. Cytoskeleton 66, 976–985 (2009). 10.1002/cm.20395 [DOI] [PubMed] [Google Scholar]
- 221. Vasileva E. and Citi S., “ The role of microtubules in the regulation of epithelial junctions,” Tissue Barriers 6, 1539596 (2018). 10.1080/21688370.2018.1539596 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 222. Shahbazi M. N., Megias D., Epifano C., Akhmanova A., Gundersen G. G., Fuchs E., and Perez-Moreno M., “ CLASP2 interacts with p120-catenin and governs microtubule dynamics at adherens junctions,” J. Cell Biol. 203, 1043–1061 (2013). 10.1083/jcb.201306019 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 223. Stehbens S. J., Paterson A. D., Crampton M. S., Shewan A. M., Ferguson C., Akhmanova A., Parton R. G., and Yap A. S., “ Dynamic microtubules regulate the local concentration of E-cadherin at cell-cell contacts,” J. Cell Sci. 119, 1801–1811 (2006). 10.1242/jcs.02903 [DOI] [PubMed] [Google Scholar]
- 224. Chen X., Kojima S.-I., Borisy G. G., and Green K. J., “ P120 catenin associates with kinesin and facilitates the transport of cadherin–catenin complexes to intercellular junctions,” J. Cell Biol. 163, 547–557 (2003). 10.1083/jcb.200305137 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 225. Ligon L. A. and Holzbaur E. L. F., “ Microtubules tethered at eepithelial cell junctions by dynein facilitate efficient junction assembly,” Traffic 8, 808–819 (2007). 10.1111/j.1600-0854.2007.00574.x [DOI] [PubMed] [Google Scholar]
- 226. Tsukita K., Kitamata M., Kashihara H., Yano T., Fujiwara I., Day T. F., Katsuno T., Kim J., Takenaga F., Tanaka H., Park S., Miyata M., Watanabe H., Kondoh G., Takahashi R., Tamura A., and Tsukita S., “ Phase separation of an actin nucleator by junctional microtubules regulates epithelial function,” Sci. Adv. 9, eadf6358 (2023). 10.1126/sciadv.adf6358 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 227. Revenu C., Streichan S., Donà E., Lecaudey V., Hufnagel L., and Gilmour D., “ Quantitative cell polarity imaging defines leader-to-follower transitions during collective migration and the key role of microtubule-dependent adherens junction formation,” Development 141, 1282–1291 (2014). 10.1242/dev.101675 [DOI] [PubMed] [Google Scholar]
- 228. Law R. A., Kiepas A., Desta H. E., Perez Ipiña E., Parlani M., Lee S. J., Yankaskas C. L., Zhao R., Mistriotis P., Wang N., Gu Z., Kalab P., Friedl P., Camley B. A., and Konstantopoulos K., “ Cytokinesis machinery promotes cell dissociation from collectively migrating strands in confinement,” Sci. Adv. 9, eabq6480 (2023). 10.1126/sciadv.abq6480 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 229. Maker A. and Gumbiner B. M., “ Reconstitution of the full transmembrane cadherin-catenin complex,” Protein Expression Purif. 193, 106056 (2022). 10.1016/j.pep.2022.106056 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 230. Rimm D. L., Koslov E. R., Kebriaei P., Cianci C. D., and Morrow J. S., “ Alpha 1(E)-catenin is an actin-binding and -bundling protein mediating the attachment of F-actin to the membrane adhesion complex,” Proc. Natl. Acad. Sci. U. S. A. 92, 8813–8817 (1995). 10.1073/pnas.92.19.8813 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 231. Rangarajan E. S., Smith E. W., and Izard T., “ Distinct inter-domain interactions of dimeric versus monomeric α-catenin link cell junctions to filaments,” Commun. Biol. 6, 276 (2023). 10.1038/s42003-023-04610-x [DOI] [PMC free article] [PubMed] [Google Scholar]
- 232. Yamada S., Pokutta S., Drees F., Weis W. I., and Nelson W. J., “ Deconstructing the cadherin-catenin-actin complex,” Cell 123, 889–901 (2005). 10.1016/j.cell.2005.09.020 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 233. Buckley C. D., Tan J., Anderson K. L., Hanein D., Volkmann N., Weis W. I., Nelson W. J., and Dunn A. R., “ The minimal cadherin-catenin complex binds to actin filaments under force,” Science 346, 1254211 (2014). 10.1126/science.1254211 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 234. Arbore C., Sergides M., Gardini L., Bianchi G., Kashchuk A. V., Pertici I., Bianco P., Pavone F. S., and Capitanio M., “ α-catenin switches between a slip and an asymmetric catch bond with f-actin to cooperatively regulate cell junction fluidity,” Nat. Commun. 13, 1146 (2022). 10.1038/s41467-022-28779-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 235. Hansen S. D., Kwiatkowski A. V., Ouyang C.-Y., Liu H., Pokutta S., Watkins S. C., Volkmann N., Hanein D., Weis W. I., Mullins R. D., and Nelson W. J., “ αE-catenin actin-binding domain alters actin filament conformation and regulates binding of nucleation and disassembly factors,” Mol. Bio. Cell 24, 3710–3720 (2013). 10.1091/mbc.e13-07-0388 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 236. Harrison O. J., Brasch J., Lasso G., Katsamba P. S., Ahlsen G., Honig B., and Shapiro L., “ Structural basis of adhesive binding by desmocollins and desmogleins,” Proc. Natl. Acad. Sci. U. S. A. 113, 7160–7165 (2016). 10.1073/pnas.1606272113 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 237. Fuchs M., Kugelmann D., Schlegel N., Vielmuth F., and Waschke J., “ Desmoglein 2 can undergo Ca2+-dependent interactions with both desmosomal and classical cadherins including E-cadherin and N-cadherin,” Biophys. J. 121, 1322–1335 (2022). 10.1016/j.bpj.2022.02.023 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 238. Mathur M., Goodwin L., and Cowin P., “ Interactions of the cytoplasmic domain of the desmosomal cadherin Dsg1 with plakoglobin,” J. Biol. Chem. 269, 14075–14080 (1994). 10.1016/S0021-9258(17)36756-X [DOI] [PubMed] [Google Scholar]
- 239. Kowalczyk A. P., Bornslaeger E. A., Borgwardt J. E., Palka H. L., Dhaliwal A. S., Corcoran C. M., Denning M. F., and Green K. J., “ The amino-terminal domain of desmoplakin binds to plakoglobin and clusters desmosomal cadherin–plakoglobin complexes,” J. Cell Biol. 139, 773–784 (1997). 10.1083/jcb.139.3.773 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 240. Fontao L., Favre B., Riou S., Geerts D., Jaunin F., Saurat J.-H., Green K. J., Sonnenberg A., and Borradori L., “ Interaction of the bullous pemphigoid antigen 1 (BP230) and desmoplakin with intermediate filaments is mediated by distinct sequences within their COOH terminus,” Mol. Biol. Cell 14, 1978–1992 (2003). 10.1091/mbc.e02-08-0548 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 241. Favre B., Begré N., Bouameur J.-E., Lingasamy P., Conover G. M., Fontao L., and Borradori L., “ Desmoplakin interacts with the coil 1 of different types of intermediate filament proteins and displays high affinity for assembled intermediate filaments,” PLoS One 13, e0205038 (2018). 10.1371/journal.pone.0205038 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 242. Brown J. W., Cho C. J., and Mills J. C., “ Paligenosis: Cellular remodeling during tissue repair,” Annu. Rev. Physiol. 84, 461–483 (2021). 10.1146/annurev-physiol-061121-035954 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 243. Savagner P., “ Epithelial–mesenchymal transitions: From cell plasticity to concept elasticity,” Curr. Top. Dev. Biol. 112, 273–300 (2015). 10.1016/bs.ctdb.2014.11.021 [DOI] [PubMed] [Google Scholar]
- 244. Poltavets V., Kochetkova M., Pitson S. M., and Samuel M. S., “ The role of the extracellular matrix and its molecular and cellular regulators in cancer cell plasticity,” Front. Oncol. 8, 431 (2018). 10.3389/fonc.2018.00431 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 245. Ross T. D., Coon B. G., Yun S., Baeyens N., Tanaka K., Ouyang M., and Schwartz M. A., “ Integrins in mechanotransduction,” Curr. Opin. Cell Biol. 25, 613–618 (2013). 10.1016/j.ceb.2013.05.006 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 246. Calderwood D. A., Campbell I. D., and Critchley D. R., “ Talins and kindlins: Partners in integrin-mediated adhesion,” Nat. Rev. Mol. Cell Biol. 14, 503–517 (2013). 10.1038/nrm3624 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 247. Puleo J. I., Parker S. S., Roman M. R., Watson A. W., Rahmani Eliato K., Peng L., Saboda K., Roe D. J., Ros R., Gertler F. B., and Mouneimne G., “ Mechanosensing during directed cell migration requires dynamic actin polymerization at focal adhesions,” J. Cell Biol. 218, 4215–4235 (2019). 10.1083/jcb.201902101 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 248. Friedl P. and Wolf K., “ Plasticity of cell migration: A multiscale tuning model,” J. Cell Biol. 188, 11–19 (2010). 10.1083/jcb.200909003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 249. Mataraza J. M., Briggs M. W., Li Z., Entwistle A., Ridley A. J., and Sacks D. B., “ IQGAP1 promotes cell motility and invasion,” J. Biol. Chem. 278, 41237–41245 (2003). 10.1074/jbc.M304838200 [DOI] [PubMed] [Google Scholar]
- 250. Belletti B., Nicoloso M. S., Schiappacassi M., Berton S., Lovat F., Wolf K., Canzonieri V., Zucchetto A., Friedl P., Colombatti A., and Baldassarre G., “ Stathmin activity influences sarcoma cell shape, motility, and metastatic potential,” Mol. Biol. Cell 19, 2003–2013 (2003). 10.1091/mbc.E07-09-0894 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 251. Ridge K. M., Eriksson J. E., Pekny M., and Goldman R. D., “ Roles of vimentin in health and disease,” Genes Dev. 36, 391–407 (2022). 10.1101/gad.349358.122 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 252. Qi L., Knifley T., Chen M., and O'Connor K. L., “ Integrin α6β4 requires plectin and vimentin for adhesion complex distribution and invasive growth,” J. Cell Sci. 135, jcs258471 (2022). 10.1242/jcs.258471 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 253. Wang Y.-J., Liang H., Liu Y., Bao Q., Yang S., Xu X.-X., Chen Y.-C., Liu W., Shi X., Shi Y., Liu X., Liu B., Gao H., Jiu Y., and Liu Y.-J., “ Lamin A/C and vimentin as a coordinated regulator during amoeboid migration in microscale confined microenvironments,” Nano Lett. 23, 6735 (2023). 10.1021/acs.nanolett.3c02096 [DOI] [PubMed] [Google Scholar]
- 254. Schiller H. B., Friedel C. C., Boulegue C., and Fässler R., “ Quantitative proteomics of the integrin adhesome show a myosin II-dependent recruitment of LIM domain proteins,” EMBO Rep. 12, 259–266 (2011). 10.1038/embor.2011.5 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 255. Ciobanasu C., Faivre B., and Le Clainche C., “ Actomyosin-dependent formation of the mechanosensitive talin–vinculin complex reinforces actin anchoring,” Nat. Commun. 5, 3095 (2014). 10.1038/ncomms4095 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 256. Franz F., Tapia-Rojo R., Winograd-Katz S., Boujemaa-Paterski R., Li W., Unger T., Albeck S., Aponte-Santamaria C., Garcia-Manyes S., Medalia O., Geiger B., and Gräter F., “ Allosteric activation of vinculin by talin,” Nat. Commun. 14, 4311 (2023). 10.1038/s41467-023-39646-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 257. Boujemaa-Paterski R., Martins B., Eibauer M., Beales C. T., Geiger B., and Medalia O., “ Talin-activated vinculin interacts with branched actin networks to initiate bundles,” eLife 9, e53990 (2020). 10.7554/eLife.53990 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 258. Huang D. L., Bax N. A., Buckley C. D., Weis W. I., and Dunn A. R., “ Vinculin forms a directionally asymmetric catch bond with F-actin,” Science 357, 703–706 (2017). 10.1126/science.aan2556 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 259. Pernier J., Cardoso Dos Santos M., Souissi M., Joly A., Narassimprakash H., Rossier O., Giannone G., Helfer E., Sengupta K., and Clainche C. L., “ Talin and kindlin cooperate to control the density of integrin clusters,” J. Cell Sci. 136, jcs260746 (2023). 10.1242/jcs.260746 [DOI] [PubMed] [Google Scholar]
- 260. Kelley C. F., Litschel T., Schumacher S., Dedden D., Schwille P., and Mizuno N., “ Phosphoinositides regulate force-independent interactions between talin, vinculin, and actin,” eLife 9, e56110 (2020). 10.7554/eLife.56110 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 261. Ehrlicher A. J., Nakamura F., Hartwig J. H., Weitz D. A., and Stossel T. P., “ Mechanical strain in actin networks regulates FilGAP and integrin binding to filamin A,” Nature 478, 260–263 (2011). 10.1038/nature10430 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 262. Spiliotis E. T. and McMurray M. A., “ Masters of asymmetry - lessons and perspectives from 50 years of septins,” Mol. Biol. Cell 31, 2289–2297 (2020). 10.1091/mbc.E19-11-0648 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 263. Nakos K., Alam M. N. A., Radler M. R., Kesisova I. A., Yang C., Okletey J., Tomasso M. R., Padrick S. B., Svitkina T. M., and Spiliotis E. T., “ Septins mediate a microtubule-actin crosstalk that enables actin growth on microtubules,” Proc. Natl. Acad. Sci. U. S. A. 119, e2202803119 (2022). 10.1073/pnas.2202803119 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 264. Zhovmer A., Manning A., Smith C., Nguyen A., Prince O., Sáez P., Ma X., Tsygankov D., Cartagena-Rivera A., Singh N., Singh R., and Tabdanov E. D., “ Septins provide microenvironment sensing and cortical actomyosin partitioning in motile amoeboid T lymphocytes,” Sci Adv. 10, eadi1788 (2024). 10.1126/sciadv.adi1788 [DOI] [PubMed] [Google Scholar]
- 265. Kim J., Mooren O. L., Onken M. D., and Cooper J. A., “ Septin and actin contributions to endothelial cell–cell junctions and monolayer integrity,” Cytoskeleton 80, 228–241 (2022). 10.1002/cm.21732 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 266. Wang X., Wang W., Wang X., Wang M., Zhu L., Garba F., Fu C., Zieger B., Liu X., Liu X., and Yao X., “ The septin complex links the catenin complex to the actin cytoskeleton for establishing epithelial cell polarity,” J. Mol. Cell Biol. 13, 395 (2021). 10.1093/jmcb/mjab036 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 267. Vendel K. J. A., Alkemade C., Andrea N., Koenderink G. H., and Dogterom M., “ In vitro reconstitution of dynamic co-organization of microtubules and actin filaments in emulsion droplets,” Methods Mol. Biol. 2101, 53–75 (2020). 10.1007/978-1-0716-0219-5_5 [DOI] [PubMed] [Google Scholar]
- 268. Siton-Mendelson O. and Bernheim-Groswasser A., “ Toward the reconstitution of synthetic cell motility,” Cell Adhes. Migr. 10, 461–474 (2016). 10.1080/19336918.2016.1170260 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 269. Fink A., Doll C. R., Yagüe Relimpio A., Dreher Y., Spatz J. P., Göpfrich K., and Cavalcanti-Adam E. A., “ Extracellular cues govern shape and cytoskeletal organization in giant unilamellar lipid vesicles,” ACS Synth. Biol. 12, 369–374 (2023). 10.1021/acssynbio.2c00516 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 270. Wittmann T., Dema A., and van Haren J., “ Lights, cytoskeleton, action: Optogenetic control of cell dynamics,” Curr. Opin. Cell Biol. 66, 1–10 (2020). 10.1016/j.ceb.2020.03.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 271. Bermudez J. G., Deiters A., and Good M. C., “ Patterning microtubule network organization reshapes cell-like compartments,” ACS Synth. Biol. 10, 1338–1350 (2021). 10.1021/acssynbio.0c00575 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 272. Brückner D. B., Arlt N., Fink A., Ronceray P., Rädler J. O., and Broedersz C. P., “ Learning the dynamics of cell–cell interactions in confined cell migration,” Proc. Natl. Acad. Sci. U. S. A. 118, e2016602118 (2021). 10.1073/pnas.2016602118 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 273. Adikes R. C., Hallett R. A., Saway B. F., Kuhlman B., and Slep K. C., “ Control of microtubule dynamics using an optogenetic microtubule plus end–F-actin cross-linker,” J. Cell Biol. 217, 779–793 (2018). 10.1083/jcb.201705190 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 274. Amiri S., Muresan C., Shang X., Huet-Calderwood C., Schwartz M. A., Calderwood D. A., and Murrell M., “ Intracellular tension sensor reveals mechanical anisotropy of the actin cytoskeleton,” Nat. Commun. 14, 8011 (2023). 10.1038/s41467-023-43612-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Data Availability Statement
The data that support the findings of this study are available from the corresponding author upon reasonable request.