Cell reorientation under cyclic stretching

Ariel Livne, Eran Bouchbinder, Benjamin Geiger, Ariel Livne, Eran Bouchbinder, Benjamin Geiger

Abstract

Mechanical cues from the extracellular microenvironment play a central role in regulating the structure, function and fate of living cells. Nevertheless, the precise nature of the mechanisms and processes underlying this crucial cellular mechanosensitivity remains a fundamental open problem. Here we provide a novel framework for addressing cellular sensitivity and response to external forces by experimentally and theoretically studying one of its most striking manifestations--cell reorientation to a uniform angle in response to cyclic stretching of the underlying substrate. We first show that existing approaches are incompatible with our extensive measurements of cell reorientation. We then propose a fundamentally new theory that shows that dissipative relaxation of the cell's passively-stored, two-dimensional, elastic energy to its minimum actively drives the reorientation process. Our theory is in excellent quantitative agreement with the complete temporal reorientation dynamics of individual cells measured over a wide range of experimental conditions, thus elucidating a basic aspect of mechanosensitivity.

Figures

Figure 1. Cyclic stretching reorients cells and…
Figure 1. Cyclic stretching reorients cells and SFs along two mirror-image angles
a-b, Phase-contrast images of REF-52 fibroblast cells on a fibronectin-coated PDMS substrate, before (a) and after (b) 6 hours of cyclic stretching (10% strain at 1.2Hz), show reorientation from random cell alignments to two, well defined, mirror-image angles. The largest principal strain (stretch) was applied in the horizontal direction as shown by double sided arrow. Bar = 100μm. c-c’, Closeup of reoriented cells (c) shows SF alignment (c′) at a similar angle to the cell body. SFs were imaged after being stained with fluorescently labeled phalloidin. Bar = 40μm. d, The mean SF orientation of individual, polarized, cells (< θSF >) matches the cell body orientation (θcell body). Data from different experiments (red squares correspond to cells from (b)) yield a linear relation of slope ~1 between the two angles (black line is the best fit: θcell body ≈ 1.02·θSF). The final orientation angle varies due to the cyclic stretching conditions (see Fig. 2 for more details). Inset shows how θcell body was determined in phase-contrast images as the long axis angle of the dark, actin-rich, cell core. Error bars represent 95% confidence intervals. Bar = 40μm. e, Analysis of individual SF orientations, θSF, at the end of the cyclic stretching (~1000 SFs from (b) and its vicinity) reveals an angular distribution with two sharp peaks, as contrasted with initial random configuration (dashed red line).
Figure 2. The final orientation angle is…
Figure 2. The final orientation angle is determined by the strains in the underlying substrate and differs from previous theoretical models
a, Cartoon of a single cell (light blue ellipse) on a deformable 2D substrate (magenta), that is stretched with principal strains: εxx & εyy (in our experiments εxx is extensional and εyy compressive). ρ marks the direction of cell body, SF (red) and FA (yellow) polarization which is at angle, θ, relative to the direction of the principal strain εxx. b, Schematic presentation of different loading and clamping conditions (left) and phase-contrast images of typical cell orientations at the end of the stretch cycles (right). The 2cm × 2cm PDMS substrate depicted in magenta, is stretched (red arrows) via clamps (black solid lines) attached at its boundaries. By adjusting the clamps’ location and size we could tune the strains transferred to the ~1 mm2 region of interest (inner dashed box), at the substrate’s center. In this manner we could control the final cell orientation (dashed yellow lines are guides to the eye). Bar = 100μm. c, The measured final orientation angle, θ‒, as a function of the biaxiality ratio, r. Each point (blue circles = SF orientations, red squares = cell body orientations) was extracted from a different experiment (1.2Hz, 4-24% strain) and represents the mean angle for the relevant cell population (n>30) in the region of interest. The green and black lines are respectively the zero strain (Eq. 1) and zero stress (Eq. 2) theoretical predictions. The dashed black line is the minimal stress prediction which extends the zero stress prediction to regions where the latter has no solution. Error bars represent 95% confidence intervals. Inset shows the SF angular distribution of a single experiment. Note that the zero strain prediction (dashed red line) is an outlier in the measured distribution, and cannot account for the discrepancy observed.
Figure 3. Final orientation angle is correctly…
Figure 3. Final orientation angle is correctly captured by the proposed theory
As predicted by the theory (Eq. 7), a linear relation between cos2(θ‒) and (r+1)−1 is clearly observed, where θ‒ is the measured final orientation angle (blue circles are data from Fig. 2c). This excellent agreement (solid black line is the best fit to Eq. 7) depends on a single parameter, b, where both the slope and the intercept are uniquely determined by it (b =1.13 ± 0.04 extracted from fit). In comparison, the zero strain prediction (Eq. 1) is depicted by the dashed black line. Additional measurements - performed on a much softer substrate (~20 kPa compared to ~1MPa) (red squares) as well as data extracted from the literature for a different cell line (green diamonds) – fall on the same line. This suggests that the elastic properties, associated with the b parameter in our theory, do not depend on substrate stiffness and are possibly cell line independent. Inset shows the same data, best fit and zero strain prediction, as above, with θ‒ plotted directly vs. r. Error bars represent 95% confidence intervals.
Figure 4. Reorientation dynamics are quantitatively explained…
Figure 4. Reorientation dynamics are quantitatively explained by the proposed theory
a, Phase contrast snapshots tracking a single cell reorientation dynamics under cyclic stretching (r=0.38, ε˘xx =0.11 at 1.2Hz). The elapsed time from the beginning of cyclic stretching is marked on each image. b, Cell body orientations, θ, of six cells, originally polarized in different directions, were recorded from the onset of cyclic stretching, t=0 (r=0.36, ε˘xx =0.11 at 1.2Hz). Reorientation takes place by a smooth rotation towards the closer of the two mirror-image final alignment angles (here: ± 64°). The individual dynamics leading up to these set points strongly depends on the initial orientation. The reorientation duration is not a simple function of the total rotation angle. Comparing the recorded reorientations to the theory’s predictions (Eq. 6) we find that individual best fits (solid curves) are not only in excellent agreement with measurements, but also all yield the same τ=6.6 ± 0.4s value (b=1.13, independently extracted from Fig. 3, and c=1, as explained in the Supplementary Note 1, were used in the analysis). c, Cells initially oriented at a similar initial angle rotate towards different final orientations according to the applied biaxiality ratio (triangles: r = 0.25, diamonds: r = 0.48, circles: r = 0.69; ε˘xx ≈ 0.10 at 1.2Hz for all three). The theory accurately describes the reorientation dynamics and predicts the same τ ~6.6s value for different cells under a wide range of experimental conditions (solid curves are single parameter fits to Eq. 6). Therefore, analysis of the smooth reorientation of a single cell towards the final orientation predicts the rotational dynamics of all other cells, even when stretched under widely different experimental conditions. d, Phase contrast snapshots tracking a single cell initially co-aligned with the stretching direction. The cell loses polarity shortly after the onset of cyclic stretching (t=400s). This is quickly followed (t=1000s) by a de-novo polarization at an angle close to the final orientation from which the cell smoothly rotates to θ‒ (stretch parameters as in (a)). In a & d, the largest principal strain (stretch) was applied in the horizontal direction and bar = 50μm.

References

    1. Discher DE, Janmey P, Wang Y. Tissue Cells Feel and Respond to the Stiffness of Their Substrate. Science. 2005;310:1139–1143.
    1. Kumar S, Weaver VM. Mechanics, malignancy, and metastasis: The force journey of a tumor cell. Cancer Metastasis Rev. 2009;28:113–127.
    1. Vogel V, Sheetz M. Local force and geometry sensing regulate cell functions. Nat Rev Mol Cell Biol. 2006;7:265–275.
    1. Geiger B, Spatz JP, Bershadsky AD. Environmental sensing through focal adhesions. Nat Rev Mol Cell Biol. 2009;10:21–33.
    1. Adamo L, et al. Biomechanical forces promote embryonic haematopoiesis. Nature. 2009;459:1131–1135.
    1. Behrndt M, et al. Forces Driving Epithelial Spreading in Zebrafish Gastrulation. Science. 2012;338:257–260.
    1. Jaalouk DE, Lammerding J. Mechanotransduction gone awry. Nat. Rev. Mol. Cell Biol. 2009;10:63–73.
    1. Orr AW, Helmke BP, Blackman BR, Schwartz MA. Mechanisms of Mechanotransduction. Dev. Cell. 2006;10:11–20.
    1. Huiskes R, Ruimerman R, van Lenthe GH, Janssen JD. Effects of mechanical forces on maintenance and adaptation of form in trabecular bone. Nature. 2000;405:704–706.
    1. Boerckel JD, Uhrig BA, Willett NJ, Huebsch N, Guldberg RE. Mechanical regulation of vascular growth and tissue regeneration in vivo. Proc. Natl. Acad. Sci. U. S. A. 2011;108:E674–E680.
    1. Buck RC. Reorientation response of cells to repeated stretch and recoil of the substratum. Exp. Cell Res. 1980;127:470–474.
    1. Wang JH-C, Goldschmidt-Clermont P, Wille J, Yin FC-P. Specificity of endothelial cell reorientation in response to cyclic mechanical stretching. J. Biomech. 2001;34:1563–1572.
    1. Jungbauer S, Gao H, Spatz JP, Kemkemer R. Two Characteristic Regimes in Frequency-Dependent Dynamic Reorientation of Fibroblasts on Cyclically Stretched Substrates. Biophys. J. 2008;95:3470–3478.
    1. Faust U, et al. Cyclic Stress at mHz Frequencies Aligns Fibroblasts in Direction of Zero Strain. PLoS ONE. 2011;6:e28963.
    1. Nagayama K, Kimura Y, Makino N, Matsumoto T. Strain waveform dependence of stress fiber reorientation in cyclically stretched osteoblastic cells: effects of viscoelastic compression of stress fibers. Am. J. Physiol. - Cell Physiol. 2012;302:C1469–C1478.
    1. Krishnan R, et al. Fluidization, resolidification, and reorientation of the endothelial cell in response to slow tidal stretches. Am. J. Physiol. - Cell Physiol. 2012;303:C368–C375.
    1. Goldyn AM, Kaiser P, Spatz JP, Ballestrem C, Kemkemer R. The kinetics of force-induced cell reorganization depend on microtubules and actin. Cytoskeleton. 2010;67:241–250.
    1. De R, Zemel A, Safran SA. Dynamics of cell orientation. Nat Phys. 2007;3:655–659.
    1. Wang JH-C. Substrate Deformation Determines Actin Cytoskeleton Reorganization: A Mathematical Modeling and Experimental Study. J. Theor. Biol. 2000;202:33–41.
    1. Safran SA, De R. Nonlinear dynamics of cell orientation. Phys Rev E Stat Nonlin Soft Matter Phys. 2009;80:060901.
    1. Kaunas R, Hsu HJ, Deguchi S. Sarcomeric model of stretch-induced stress fiber reorganization. Cell Health Cytoskelet. 2011;3:13–22. doi:10.2147/CHC.S14984.
    1. Stamenović D, Lazopoulos KA, Pirentis A, Suki B. Mechanical Stability Determines Stress Fiber and Focal Adhesion Orientation. Cell. Mol. Bioeng. 2009;2:475–485.
    1. Balaban NQ, et al. Force and focal adhesion assembly: a close relationship studied using elastic micropatterned substrates. Nat Cell Biol. 2001;3:466–472.
    1. Civelekoglu-Scholey G, et al. Model of coupled transient changes of Rac, Rho, adhesions and stress fibers alignment in endothelial cells responding to shear stress. J. Theor. Biol. 2005;232:569–585.
    1. Riveline D, et al. Focal Contacts as Mechanosensors Externally Applied Local Mechanical Force Induces Growth of Focal Contacts by an Mdia1-Dependent and Rock-Independent Mechanism. J. Cell Biol. 2001;153:1175–1186.
    1. Landau LD, Lifshitz EM. Theory of Elasticity. Pergamon Press; 1986.
    1. Solon J, Levental I, Sengupta K, Georges PC, Janmey PA. Fibroblast Adaptation and Stiffness Matching to Soft Elastic Substrates. Biophys. J. 2007;93:4453–4461.
    1. Lu L, Oswald SJ, Ngu H, Yin FC-P. Mechanical Properties of Actin Stress Fibers in Living Cells. Biophys. J. 2008;95:6060–6071.
    1. Schwarz US, Gardel ML. United we stand – integrating the actin cytoskeleton and cell–matrix adhesions in cellular mechanotransduction. J. Cell Sci. 2012;125:3051–3060.
    1. Wolfenson H, et al. A Role for the Juxtamembrane Cytoplasm in the Molecular Dynamics of Focal Adhesions. PLoS ONE. 2009;4:e4304.
    1. Fritzsche M, Lewalle A, Duke T, Kruse K, Charras G. Analysis of turnover dynamics of the submembranous actin cortex. Mol. Biol. Cell. 2013;24:757–767.
    1. Trepat X, et al. Universal physical responses to stretch in the living cell. Nature. 2007;447:592–595.
    1. Kaunas R, Nguyen P, Usami S, Chien S. Cooperative effects of Rho and mechanical stretch on stress fiber organization. Proc. Natl. Acad. Sci. U. S. A. 2005;102:15895–15900.
    1. Franz CM, Müller DJ. Analyzing focal adhesion structure by atomic force microscopy. J. Cell Sci. 2005;118:5315–5323.
    1. Patla I, et al. Dissecting the molecular architecture of integrin adhesion sites by cryo-electron tomography. Nat. Cell Biol. 2010;12:909–915.
    1. Plotnikov SV, Pasapera AM, Sabass B, Waterman CM. Force Fluctuations within Focal Adhesions Mediate ECM-Rigidity Sensing to Guide Directed Cell Migration. Cell. 2012;151:1513–1527.
    1. Aratyn-Schaus Y, Oakes PW, Gardel ML. Dynamic and structural signatures of lamellar actomyosin force generation. Mol. Biol. Cell. 2011;22:1330–1339.
    1. Shutova M, Yang C, Vasiliev JM, Svitkina T. Functions of Nonmuscle Myosin II in Assembly of the Cellular Contractile System. PLoS ONE. 2012;7:e40814.
    1. Zaidel-Bar R, Itzkovitz S, Ma’ayan A, Iyengar R, Geiger B. Functional atlas of the integrin adhesome. Nat. Cell Biol. 2007;9:858–867.
    1. Xu W, Baribault H, Adamson ED. Vinculin knockout results in heart and brain defects during embryonic development. Development. 1998;125:327–337.
    1. Bershadsky A, Kozlov M, Geiger B. Adhesion-mediated mechanosensitivity: a time to experiment, and a time to theorize. Curr. Opin. Cell Biol. 2006;18:472–481.
    1. Kumar S, et al. Viscoelastic Retraction of Single Living Stress Fibers and Its Impact on Cell Shape, Cytoskeletal Organization, and Extracellular Matrix Mechanics. Biophys. J. 2006;90:3762–3773.
    1. Hinz B, Gabbiani G. Mechanisms of force generation and transmission by myofibroblasts. Curr. Opin. Biotechnol. 2003;14:538–546.
    1. Zaidel-Bar R, Milo R, Kam Z, Geiger B. A paxillin tyrosine phosphorylation switch regulates the assembly and form of cell-matrix adhesions. J. Cell Sci. 2007;120:137–148.
    1. Zemel A, Rehfeldt F, Brown AEX, Discher DE, Safran SA. Optimal matrix rigidity for stress-fibre polarization in stem cells. Nat. Phys. 2010;6:468–473.

Source: PubMed

3
Předplatit