Skip to main content
Log in

Continuum modeling of non-conservative fluid membrane for simulating long-term cell dynamics

  • Regular Article - Living Systems
  • Published:
The European Physical Journal E Aims and scope Submit manuscript

Abstract

Living cells actively deform and move by their force generations in three-dimensional (3D) space. These 3D cell dynamics occur over a long-term time scale, ranging from tens of minutes to days. On such a time scale, turnover of cell membrane constituents due to endocytosis and exocytosis cannot be ignored, i.e., the surface membrane dynamically deforms without mass conservation. Although membrane turnover is essential for large deformation of cells, there is no computational framework yet to simulate long-term cell dynamics with a non-conservative fluidic membrane. In this paper, we proposed a computational framework for simulating the long-term dynamics of a cell membrane in 3D space. For this purpose, in the proposed framework, the cell surface membrane is treated as a viscous fluid membrane without mass conservation. Cell shape is discretized by a triangular mesh, and its dynamics are expressed by effective energy and dissipation function. The mesh structure, distorted by membrane motion, is dynamically optimized by introducing a modified dynamic remeshing method. To validate the proposed framework, numerical simulations were performed, showing that the membrane flow is reproduced in a physically consistent manner and that the artificial effects of the remeshing method were negligible. To further demonstrate the applicability of the proposed framework, numerical simulations of cell migration induced by a mechanism similar to the Marangoni effect, i.e., the polarized surface tension actively generated by the cell, were performed. The observed cell behaviors agreed with existing analytical solutions, indicating that the proposed computational framework can quantitatively reproduce long-term active cell dynamics with membrane turnover. Based on the simple description of cell membrane dynamics, this framework provides a useful basis for analyzing various cell shaping and movement.

Graphical abstract

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7

Similar content being viewed by others

Data Availability Statement

The datasets generated during and/or analyzed during the current study are partially available from the corresponding author on reasonable request.

References

  1. S.F. Gilbert, Developmental Biology, 10th edn. (Sinauer Associates, 2014)

  2. J. Davies, Mechanisms of Morphogenesis (Academic Press, Cambridge, 2013)

    Google Scholar 

  3. B. Lim, P.A.J. Bascom, R.S.C. Cobbold, Simulation of red blood cell aggregation in shear flow. Biorheology 34, 423–441 (1997)

    Article  Google Scholar 

  4. Y. Liu, W.K. Liu, Rheology of red blood cell aggregation by computer simulation. J. Comput. Phys. 220, 139–154 (2006)

    Article  ADS  MathSciNet  Google Scholar 

  5. K. Tsubota, S. Wada, T. Yamaguchi, Particle method for computer simulation of red blood cell motion in blood flow. Comput. Methods Programs Biomed. 83, 139–146 (2006)

    Article  Google Scholar 

  6. J. Mauer, S. Mendez, L. Lanotte, F. Nicoud, M. Abkarian, G. Gompper, D.A. Fedosov, Flow-induced transitions of red blood cell shapes under shear. Phys. Rev. Lett. 121, 118103 (2018)

    Article  ADS  Google Scholar 

  7. N. Geekiyanage, E. Sauret, S. Saha, R. Flower, Y. Gu, Modelling of red blood cell morphological and deformability changes during in-vitro storage. Appl. Sci. 10, 3209 (2020)

    Article  Google Scholar 

  8. I. Jančigová, K. Kovalčíková, R. Weeber, I. Cimrák, PyOIF: computational tool for modelling of multi-cell flows in complex geometries. PLoS Comput. Biol. 16, e1008249 (2020)

    Article  ADS  Google Scholar 

  9. I. Jančigová, K. Kovalčíková, A. Bohiniková, I. Cimrák, Spring-network model of red blood cell: from membrane mechanics to validation. Int. J. Numer. Meth. Fluids 92, 1368–1393 (2020)

    Article  MathSciNet  Google Scholar 

  10. C. Pozrikidis et al., Boundary Integral and Singularity Methods for Linearized Viscous Flow (Cambridge University Press, Cambridge, 1992)

    Book  Google Scholar 

  11. S. Kim, S.J. Karrila, Microhydrodynamics: principles and selected applications, Courier Corporation, (2013)

  12. R.M. MacMECCAN, J.R. Clausen, G.P. Neitzel, C.K. Aidun, Simulating deformable particle suspensions using a coupled lattice-Boltzmann and finite-element method. J. Fluid Mech. 618, 13 (2009)

    Article  ADS  MathSciNet  Google Scholar 

  13. B.D. Nichols, C.W. Hirt, Methods for calculating multidimensional, transient free surface flows past bodies, in Proceedings of the First International Conference on Numerical Ship Hydrodynamics, (1975)

  14. S. Osher, J.A. Sethian, Fronts propagating with curvature-dependent speed: algorithms based on Hamilton-Jacobi formulations. J. Comput. Phys. 79, 12–49 (1988)

    Article  ADS  MathSciNet  Google Scholar 

  15. D.E. Fyfe, E.S. Oran, M.J. Fritts, Surface tension and viscosity with Lagrangian hydrodynamics on a triangular mesh. J. Comput. Phys. 76, 349–384 (1988)

    Article  ADS  Google Scholar 

  16. M. Dai, D.P. Schmidt, Adaptive tetrahedral meshing in free-surface flow. J. Comput. Phys. 208, 228–252 (2005)

    Article  ADS  Google Scholar 

  17. G. Wang, T. Galli, Reciprocal link between cell biomechanics and exocytosis. Traffic 19, 741–749 (2018)

    Article  Google Scholar 

  18. M. Bergert, A. Erzberger, R.A. Desai, I.M. Aspalter, A.C. Oates, G. Charras, G. Salbreux, E.K. Paluch, Force transmission during adhesion-independent migration. Nat. Cell Biol. 17, 524–529 (2015)

    Article  Google Scholar 

  19. A. Farutin, J. Etienne, C. Misbah, P. Recho, Crawling in a fluid. Phys. Rev. Lett. 123, 118101 (2019)

    Article  ADS  MathSciNet  Google Scholar 

  20. C. Bächer, D. Khoromskaia, G. Salbreux, S. Gekle, A three-dimensional numerical model of an active cell cortex in the viscous limit. Front. Phys. 9, 562 (2021)

    Article  Google Scholar 

  21. V. Cristini, J. Bławzdziewicz, M. Loewenberg, An adaptive mesh algorithm for evolving surfaces: simulations of drop breakup and coalescence. J. Comput. Phys. 168, 445–463 (2001)

    Article  ADS  Google Scholar 

  22. S. Kwak, C. Pozrikidis, Adaptive triangulation of evolving, closed, or open surfaces by the advancing-front method. J. Comput. Phys. 145, 61–88 (1998)

    Article  ADS  MathSciNet  Google Scholar 

  23. S. Okuda, M. Eiraku, Role of molecular turnover in dynamic deformation of a three-dimensional cellular membrane. Biomech. Model. Mechanobiol. 16, 1805–1818 (2017)

    Article  Google Scholar 

  24. H. Noguchi, G. Gompper, Shape transitions of fluid vesicles and red blood cells in capillary flows. Proc. Natl. Acad. Sci. 102, 14159–14164 (2005)

    Article  ADS  Google Scholar 

  25. H. Ni, G.A. Papoian, Membrane-medyan: Simulating deformable vesicles containing complex cytoskeletal networks. J. Phys. Chem. B 125, 10710--10719 (2021)

    Article  Google Scholar 

  26. A. Laadhari, P. Saramito, C. Misbah, An adaptive finite element method for the modeling of the equilibrium of red blood cells. Int. J. Numer. Meth. Fluids 80, 397–428 (2016)

    Article  MathSciNet  Google Scholar 

  27. J. Vorsatz, Ch. Rössl, H.-P. Seidel, Dynamic remeshing and applications. J. Comput. Inf. Sci. Eng. 3, 338–344 (2003)

    Article  Google Scholar 

  28. M. Belkin, J. Sun, Y. Wang, Discrete Laplace operator on meshed surfaces, in Proceedings of the Twenty-Fourth Annual Symposium on Computational Geometry, pp. 278–287 (2008)

  29. M. Bušík, M. Slavík, I. Cimrák, Dissipative coupling of fluid and immersed objects for modelling of cells in flow. Comput. Math. Methods Med. 2018, 7842857 (2018)

    Article  Google Scholar 

  30. S. Okuda, Y. Inoue, M. Eiraku, T. Adachi, Y. Sasai, Vertex dynamics simulations of viscosity-dependent deformation during tissue morphogenesis. Biomech. Model Mechanobiol. 14, 413–425 (2015)

    Article  Google Scholar 

  31. S. Okuda, K. Sato, Polarized interfacial tension induces collective migration of cells, as a cluster, in a 3D tissue. Biophys. J. 121, 1856-1867 (2022)

    Article  ADS  Google Scholar 

  32. A. Saha, M. Nishikawa, M. Behrndt, C.-P. Heisenberg, F. Jülicher, S.W. Grill, Determining physical properties of the cell cortex. Biophys. J. 110, 1421–1429 (2016)

    Article  ADS  Google Scholar 

  33. P. van Liedekerke, J. Neitsch, T. Johann, E. Warmt, I. Gonzàlez-Valverde, S. Hoehme, S. Grosser, J. Kaes, D. Drasdo, A quantitative high-resolution computational mechanics cell model for growing and regenerating tissues. Biomech. Model Mechanobiol. 19, 189–220 (2020)

    Article  Google Scholar 

  34. M. Klima, M. Kucharik, M. Shashkov, Combined swept region and intersection-based single-material remapping method. Int. J. Numer. Meth. Fluids 85, 363–382 (2017)

    Article  MathSciNet  Google Scholar 

  35. M. Drechsler, F. Giavazzi, R. Cerbino, I.M. Palacios, Active diffusion and advection in Drosophila oocytes result from the interplay of actin and microtubules. Nat. Commun. 8, 1–11 (2017)

    Article  Google Scholar 

  36. Y. Qin, Y. Li, L.-Y. Zhang, G.-K. Xu, Stochastic fluctuation-induced cell polarization on elastic substrates: A cytoskeleton-based mechanical model. J. Mech. Phys. Solids 137, 103872 (2020)

    Article  MathSciNet  Google Scholar 

  37. J.-T. Hang, Y. Kang, G.-K. Xu, H. Gao, A hierarchical cellular structural model to unravel the universal power-law rheological behavior of living cells, Nature. Communications 12, 1–7 (2021)

    Google Scholar 

  38. O. Chaudhuri, J. Cooper-White, P.A. Janmey, D.J. Mooney, V.B. Shenoy, Effects of extracellular matrix viscoelasticity on cellular behaviour. Nature 584, 535–546 (2020)

    Article  ADS  Google Scholar 

  39. F. Jülicher, The morphology of vesicles of higher topological genus: conformal degeneracy and conformal modes. J. Phys. II 6, 1797–1824 (1996)

    Google Scholar 

Download references

Acknowledgements

We would like to thank Dr. Alexander S. Mikhailov at Kanazawa University, Dr. Yasuhiro Inoue at Kyoto University, and Dr. Ken-ichi Tsubota at Chiba University for useful discussions. This work was supported by the Japan Science and Technology Agency (JST), CREST Grant No. JPMJCR1921 (SO); the Japan Agency for Medical Research and Development (AMED), Grant No. 21bm0704065h0002 (SO); the Japan Society for the Promotion of Science (JSPS), KAKENHI Grant Nos. 21H01209 (SO), 20K03871 (KS); Global Station for Soft Matter at Hokkaido University (KS); the Research Program of "Five-star Alliance" in "NJRC Mater. and Dev." (KS); and the World Premier International Research Center Initiative, Ministry of Education, Culture, Sports, Science and Technology (MEXT), Japan.

Author information

Authors and Affiliations

Authors

Contributions

SO conceived of the project, developed the software, and performed the analyses; SO and KS calculated the analytical solutions; SO and TH wrote the manuscript. All authors contributed to the final manuscript.

Corresponding author

Correspondence to Satoru Okuda.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Dynamics of cell migration induced by polarized surface tension. Arrows indicate local velocities of the membrane, while colors indicate the magnitude. This result was obtained from the simulation with \(\Gamma_{1} = 0.4\Gamma_{0}\) and \(\mu = 0\) (M4V 1628 kb)

Appendix

Appendix

1.1 A1: Membrane flow under area constraint

The motion equation of the vertices in Eq. (7) is derived from that of the continuum in Eq. (1); however, whether membrane dynamics can be reproduced in a physically consistent manner remains to be determined. Therefore, whether planar membrane flow can be reproduced when the total membrane surface was conserved was determined. Numerical simulations of planar membrane flow were performed, and the obtained velocity profiles were evaluated and compared with theoretical solutions.

In \(xyz\)-coordinates, a rectangular system was considered whose lengths along the individual axes are represented by \(L_{x}\), \(L_{y}\), and \(L_{z}\) (Fig. 

Fig. 8
figure 8

Hagen–Poiseuille flow of membrane. a. Snapshot of membrane flow resulting from numerical simulations. Triangles outside boundaries are not shown. Colors represent membrane velocity along the \(x\)-axis, \(v_{x}\). b. Spatial profile of membrane velocity along the \(x\)-axis, \(v_{x}\), under the condition with \(F = 0.01\,\mu l^{2} /\tau\). Black dots indicate individual vertex velocities at \(t = 100\tau\). The red curve represents the theoretical solution of Eq. (26). ce. Average maximum velocity, \(v_{{{\text{max}}}}\), as a function of pressure gradient, discretization length \(l_{{\text{d}}}\), and threshold ratio \(\lambda\), respectively. The red line in c-e represents the theoretical solutions of Eq. (27). Error bars in c-e indicate standard deviations

8a). A periodic boundary condition was imposed on \(x = 0, L_{x}\) and \(y = 0, L_{y}\). The membrane spreads in the \(xy\)-plane across the boundaries. By representing the \(x\)-component of the ith vertex location by \(x_{i}\), the energetic force is introduced by the effective energy \(U\) in Eq. (7), which is expressed as follows:

$$ U \equiv \left\{ {\begin{array}{*{20}c} { - \mathop \sum \limits_{i}^{{{\text{vertex}}}} \frac{F}{{\rho^{*} }}x_{i} } & {y_{i} < \frac{{L_{y} }}{2}} \\ {\mathop \sum \limits_{i}^{{{\text{vertex}}}} \frac{F}{{\rho^{*} }}x_{i} } & {y_{i} \ge \frac{{L_{y} }}{2}} \\ \end{array} } \right., $$
(24)

where constant \(F\) is the stress exerted on the membrane that induces the pressure gradient as

$$ \frac{dp}{{dx_{i} }} = - \rho_{*} \nabla_{i} U. $$
(25)

Theoretically, this pressure gradient induces a Hagen–Poiseuille flow, whose velocity profile is given as

$$ v_{x} \left( y \right) = \left\{ {\begin{array}{*{20}c} { - 16v_{{{\text{max}}}} \frac{{y_{i} }}{{L_{y} }}\left( {\frac{{y_{i} }}{{L_{y} }} - \frac{1}{2}} \right)} & {\frac{{y_{i} }}{{L_{y} }} < \frac{1}{2}} \\ {16v_{{{\text{max}}}} \left( {\frac{{y_{i} }}{{L_{y} }} - \frac{1}{2}} \right)\left( {\frac{{y_{i} }}{{L_{y} }} - 1} \right)} & {\frac{{y_{i} }}{{L_{y} }} \ge \frac{1}{2}} \\ \end{array} } \right.. $$
(26)

The maximum velocity along the \(x\)-axis, represented by \(v_{{{\text{max}}}}\), is given by

$$ v_{{{\text{max}}}} = \frac{{L_{y}^{2} F}}{8\mu }. $$
(27)

The system size was set to \(L_{x} = 12.2l\) and \(L_{y} = 21.1l\), where \(l\) is the unit length. Friction from the environment was set to \(\xi = 0\). All parameters were non-dimensionalized by units of length, \(l\); time, \(\tau = 100{\Delta }t\); and energy, \(\mu l^{3} /\tau\). The control parameters, \(F\), \(l_{{\text{d}}}\), and \(\lambda\), were explored.

Numerical simulations were performed with several values of \(F\), and vertex velocities were observed (Fig. 8b). The resulting velocity profile agreed with the theoretical solution obtained by Eq. (26). By fitting Eq. (26) to the vertex velocities using the least-squares method, the maximum velocity, \(v_{{{\text{max}}}}\), was estimated (Fig. 8c). The velocities estimated in the simulations agreed with the theoretical solution. Moreover, the numerical parameters, \(l_{{\text{d}}}\) and \(\lambda\), were also varied (Fig. 8d, e). Deviations of the numerical solutions from the theoretical solutions given as Eq. (27) were very small and nearly independent of \(l_{{\text{d}}}\) and \(\lambda\). These results indicate that the planar membrane flow was reproduced in a physically consistent manner. Note that the membrane dynamics are given deterministically in this simulation, whereas the observed velocities were subject to some variability, expressed as standard deviation. This variability was caused by the heterogeneous structure of the triangular mesh, as well as the variation of velocities in the x-axis direction; because the triangular mesh discretizing the membrane is distorted, its structure is heterogeneous, inducing artificial variations in velocities. Moreover, because friction from the environment was set to zero, membrane velocity was determined as a relative velocity, allowing it to physically vary in the x-axis direction.

1.2 A2: Negligible gaps of geometry and energy in dynamic remeshing

Although planar membrane flow was successfully reproduced in the proposed framework (Sect. A1 in Appendix), the flow in 3D space may be affected by the dynamic remeshing algorithm. To clarify whether the artificial effects introduced by the remeshing algorithm are negligible, numerical simulations of rotational cell dynamics were performed under simple shear flow, and gaps in the membrane geometry and energy induced by the remeshing algorithm were evaluated.

The effective energy, \(U\) in Eq. (7), is expressed as

$$ U \equiv - nk_{B} T\log \left( {\frac{V}{{V_{{{\text{ref}}}} }}} \right) + \mathop \sum \limits_{i}^{{{\text{vertex}}}} \Gamma_{0} A_{i} + \mathop \sum \limits_{i}^{{{\text{vertex}}}} 2K\frac{{M_{i}^{2} }}{{A_{i} }} $$
(28)

The first term indicates the osmotic energy of the cytoplasm [30], where the constant \(nk_{B} T\) is the modulus of the cell volume. Variable \(V\) and constant \(V_{{{\text{ref}}}}\) are the current and reference volumes of the cell, respectively. The second term in Eq. (28) indicates the surface energy of the membrane, where constant \(\Gamma_{0}\) is the surface tension and variable \(A_{i}\) is the local surface area around the ith vertex, given by \(A_{i} = \mathop \sum \nolimits_{j\left( i \right)}^{{{\text{triangle}}}} a_{j} /3\), where \(a_{j}\) is the area of the jth triangle that comprises the ith vertex. Assuming that the first and second terms create a force balance on a spherical geometry when \(V = V_{{{\text{ref}}}}\), \(nk_{B} T\) may be rewritten as a function of \(\Gamma_{0}\) [30]:

$$ nk_{B} T = \Gamma_{0} \left( {\frac{{32\pi V_{{{\text{ref}}}}^{2} }}{3}} \right)^{\frac{1}{3}} $$
(29)

The third term in Eq. (28) indicates the bending energy of the membrane [39], where \(K\) is the bending rigidity. Variable \(M_{i}\) is the total mean curvature around the ith vertex, given by \(M_{i} = \mathop \sum \nolimits_{j\left( i \right)}^{{{\text{edge}}}} l_{j} \theta_{j} /4\), where \(l_{j}\) and \(\theta_{j}\) are the length and angle of the jth edge connected to the ith vertex, respectively.

A system box with a sufficiently large space was considered in \(xyz\)-coordinates. To provide a Couette flow as the environment, a simple shear flow was introduced as \({\varvec{u}}_{{\text{e}}} \equiv \left( {qz, 0, 0} \right)\), where \(q\) is the gradient in velocity along the \(z\)-axis. The shear flow exerts shear stress on the cell via the friction force term in Eq. (7), balancing the surface tension in Eq. (28). The physical parameters were non-dimensionalized by length, \(l = \left( {3V_{{{\text{ref}}}} /4000\pi } \right)^{1/3}\); time, \(\tau = \mu l/\Gamma_{0}\); and energy, \(\Gamma_{0} l^{2}\). The remaining parameters were set to \(\xi = \mu /2l^{2}\), \(K = \Gamma_{0} l^{2}\), \(q = 1/\tau\), and \({\Delta }t = 0.01\tau\), while the control parameters, \(l_{{\text{d}}}\) and \(\lambda\), were explored.

Under shear flow, the cell rotated while maintaining its shape (Fig. 

Fig. 9
figure 9

Geometric and energetic gaps via remeshing. a. Snapshots of a rotating cell, obtained at \(t = 100\tau\). The colors indicate local membrane velocities, \({\varvec{u}}_{i}\). be. Average gap ratios of area, volume, curvature, and energy before and after the remeshing operation, as functions of the discretization length \(l_{{\text{d}}}\). Results in be are from numerical simulations with \(\lambda = 0.01\). fi. Average gaps of area, volume, curvature, and energy before and after the remeshing operation, as functions of the threshold ratio \(\lambda\). Results in fi were obtained from simulations with \(l_{{\text{d}}} = l\). Error bars in bi indicate standard deviations

9a). To evaluate the geometric and energetic gaps, several physical quantities were observed: the volume, \(V\); total area, At \(= \mathop \sum \nolimits_{i} A_{i}\); total mean curvature, \(M_{{\text{t}}} = \mathop \sum \nolimits_{i} M_{i}\); and total energy, \(U\). During cell rotations, the gaps of \(V\), \(A_{{\text{t}}}\), \(M_{{\text{t}}}\), and \(U\) were maintained at low values within certain variances (Fig. 9b–e). Importantly, both the averages and variances of these parameters were reduced by decreasing the discretized length, \(l_{{\text{d}}}\) (Fig. 9b–e). Additionally, to evaluate the effects of the correction term in Eq. (13), the cases were compared with and without the correction (Fig. 9b–e). Gaps in the case with the correction were approximately half those without the correction, indicating that the correction term was effective. Moreover, these gaps were reduced by decreasing the threshold ratio, \(\lambda\) (Fig. 9f–i). For example, the gaps became less than approximately \(0.001\%\) when \(l_{{\text{d}}} \le l\) and \(\lambda \le 0.01\). These results indicate that the gaps in membrane geometry and energy generated by the remeshing algorithm can be made negligible by modulating \(l_{{\text{d}}}\).

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Okuda, S., Sato, K. & Hiraiwa, T. Continuum modeling of non-conservative fluid membrane for simulating long-term cell dynamics. Eur. Phys. J. E 45, 69 (2022). https://doi.org/10.1140/epje/s10189-022-00223-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1140/epje/s10189-022-00223-0

Navigation