Skip to main content
Log in

Geometrical Effects on Nonlinear Electrodiffusion in Cell Physiology

  • Published:
Journal of Nonlinear Science Aims and scope Submit manuscript

Abstract

We report here new electrical laws, derived from nonlinear electrodiffusion theory, about the effect of the local geometrical structure, such as curvature, on the electrical properties of a cell. We adopt the Poisson–Nernst–Planck equations for charge concentration and electric potential as a model of electrodiffusion. In the case at hand, the entire boundary is impermeable to ions and the electric field satisfies the compatibility condition of Poisson’s equation. We construct an asymptotic approximation for certain singular limits to the steady-state solution in a ball with an attached cusp-shaped funnel on its surface. As the number of charge increases, they concentrate at the end of cusp-shaped funnel. These results can be used in the design of nanopipettes and help to understand the local voltage changes inside dendrites and axons with heterogeneous local geometry.

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.

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

Similar content being viewed by others

References

  • Bezanilla, F.: How membrane proteins sense voltage. Nat. Rev. Mol. Cell Biol. 9, 323–332 (2008)

  • Cartailler, J., Schuss, Z., Holcman, D.: Analysis of the Poisson-Nernst-Planck equation in a ball for modeling the voltage-current relation in neurobiological microdomains. Phys. D Nonlinear Phenom. 339, 39–48 (2016)

    Article  MathSciNet  Google Scholar 

  • Courant, R., Hilbert, D.: Methods of Mathematical Physics, vol. 2. Wiley Interscience, New York (1989)

    Book  MATH  Google Scholar 

  • Debye, P., Hückel, E.: Zur Theorie der Elektrolyte. I. Gefrierpunktserniedrigung und verwandte Erscheinungen. Phys. Z. 24(9), 185–206 (1923)

    MATH  Google Scholar 

  • Delgado, M., Coombs, D.: Conditional mean first passage times to small traps in a 3-D domain with a sticky boundary. SIAM J. Multiscale Anal. Simul. 13(4), 1224–1258 (2015)

    Article  MATH  Google Scholar 

  • Eisenberg, R.S.: From structure to function in open ionic channels. J. Membr. Biol. 171, 1–24 (1998a)

    Article  Google Scholar 

  • Eisenberg, R.S.: Ionic channels in biological membranes. Electrostatic analysis of a natural nanotube. Contemp. Phys. 39(6), 447–466 (1998b)

    Article  Google Scholar 

  • Eisenberg, R.S., Klosek, M.M., Schuss, Z.: Diffusion as a chemical reaction: stochastic trajectories between fixed concentrations. J. Chem. Phys. 102(4), 1767–1780 (1995)

    Article  Google Scholar 

  • Hille, B.: Ion Channels of Excitable Membranes, 3rd edn. Sinauer Associates, Sunderland (2001)

  • Holcman, D., Schuss, Z.: Brownian motion in dire straits. SIAM J. Multiscale Model. Simul. 10(4), 1204–1231 (2012a)

    Article  MathSciNet  MATH  Google Scholar 

  • Holcman, D., Schuss, Z.: Brownian motion in dire straits. Multiscale Model. Simul. 10(4), 1204–1231 (2012b)

    Article  MathSciNet  MATH  Google Scholar 

  • Holcman, D., Schuss, Z.: Stochastic Narrow Escape in Molecular an Cellular Biology. Analysis and Applications. Springer, New York (2015)

    Book  MATH  Google Scholar 

  • Holcman, D., Yuste, R.: The new nanophysiology: regulation of ionic flow in neuronal subcompartments. Nat. Rev. Neurosci. 16, 685–692 (2015)

    Article  Google Scholar 

  • Holcman, D., Hoze, N., Schuss, Z.: Narrow escape through a funnel and effective diffusion on a crowded membrane. Phys. Rev. E 84, 021906 (2011). Erratum. Phys. Rev. E 85, 039903 (2012)

    Article  Google Scholar 

  • Horn, R., Roux, B., Aqvist, J.: Permeation redux: thermodynamics and kinetics of ion movement through potassium channels. Biophys J. 106(9), 1859–1863 (2014)

    Article  Google Scholar 

  • Lindsay, A.E., Bernoff, A.J., Ward, M.J.: First passage statistics for the capture of a Brownian particle by a structured spherical target with multiple surface traps. SIAM J. Multiscale Model. Simul. 15(1), 74–109 (2016)

  • Mamonov, A., Coalson, R., Nitzan, A., Kurnikova, M.: The role of the dielectric barrier in narrow biological channels: a novel composite approach to modeling single channel currents. Biophys. J. 84, 3646–3661 (2003)

    Article  Google Scholar 

  • Perry, D., Momotenko, D., Lazenby, R.A., Kang, M., Unwin, P.R.: Characterization of nanopipettes. Anal. Chem. 88(10), 5523–5530 (2016)

    Article  Google Scholar 

  • Pillay, S., Peirce, A., Kolokolnikov, T., Ward, M.: An asymptotic analysis of the mean first passage time for narrow escape problems: part I: two-dimensional domains. SIAM Multiscale Model. Simul. 8(3), 803–835 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  • Qian, N., Sejnowski, T.J.: An electro-diffusion model for computing membrane potentials and ionic concentrations in branching dendrites, spines and axons. Biol. Cybern. 62, 1–15 (1989)

    Article  MATH  Google Scholar 

  • Rall, W.: Cable theory for dendritic neurons. In: Koch, C., Segev, I. (eds.) Methods in Neuronal Modeling: From Synapses to Networks, pp. 9–63. MIT Press, Cambridge (1989)

    Google Scholar 

  • Roux, B., Karplus, M.: Ion transport in the gramicidin channel: free energy of the solvated right-handed dimer in a model membrane. J. Am. Chem. Soc. 115, 3250–3262 (1993)

    Article  Google Scholar 

  • Ruiz, F.J.G., Godoy, A., Gamiz, F., Sampedro, C., Donetti, L.: A comprehensive study of the corner effects in Pi-gate MOSFETs including quantum effects. IEEE Trans. Electron Devices 54(12), 3369–3377 (2007)

    Article  Google Scholar 

  • Savtchenko, L.P., Kulahin, N., Korogod, S.M., Rusakov, D.A.: Electric fields of synaptic currents could influence diffusion of charged neurotransmitter molecules. Synapse 51(4), 270–278 (2004)

    Article  Google Scholar 

  • Schuss, Z., Nadler, B., Eisenberg, R.S.: Derivation of Poisson and Nernst-Planck equations in a bath and channel from a molecular model. Phys. Rev. E 64, 036116 (2001)

    Article  Google Scholar 

  • Singer, A., Norbury, J.: A Poisson-Nernst-Planck model for biological ion channels–an asymptotic analysis in a three-dimensional narrow funnel. SIAM Appl. Math. 7(3), 949–968 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  • Sparreboom, W., van den Berg, A., Eijkel, J.C.T.: Principles and applications of nanofluidic transport. Nat. Nanotechnol 4, 713–720 (2009)

    Article  Google Scholar 

  • Yuste, R.: Dendritic Spines. The MIT Press, Cambridge (2010)

    Book  Google Scholar 

Download references

Acknowledgements

Funding was provided by H2020 Marie Skłodowska-Curie Actions (Grant No. 4328435).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to D. Holcman.

Additional information

Communicated by Philip K. Maini.

Appendix

Appendix

1.1 Regular Expansion of the PNP Solution when There are an Excess of Positive and a Small Number of Negative Charges

We show that for the concentrations of cations and anions found in the literature (Hille 2001), the leading-order solution of the electrical potential in the bulk can be obtained by considering positive charges only. We assume that the charge of an electrolyte confined in \(\tilde{\varOmega }\) consists of identical \(N_{p}\) positive and \(N_m\) negative ions with density \(q_{p}({\varvec{x}})\) and \(q_m({\varvec{x}})\) such as

$$\begin{aligned} N_i=\int _{\tilde{\varOmega }}q_i({\tilde{{\varvec{x}}}})\,\mathrm{d}{\tilde{{\varvec{x}}}},\quad \text{ for } \, i \in \{ p,\,m\}, \end{aligned}$$
(117)

where p and m are positive and negative species, respectively. The total charges in \(\tilde{\varOmega }\) are the sum

$$\begin{aligned} Q =e( N_{p}-N_m). \end{aligned}$$
(118)

The associated charge densities \(\rho _{p}({\varvec{x}},t)\) and \(\rho _m({\varvec{x}},t)\) satisfy the boundary value problem for the Nernst–Planck equation

$$\begin{aligned}&D_i\left[ \Delta \rho _i({\tilde{{\varvec{x}}}},t) +\frac{z_ie}{kT} \nabla \left( \rho _i({\tilde{{\varvec{x}}}},t) \nabla \phi ({\tilde{{\varvec{x}}}},t)\right) \right] =\, \frac{\partial \rho _i({\tilde{{\varvec{x}}}},t)}{\partial t}\quad \text{ for }\ {\tilde{{\varvec{x}}}}\in \tilde{\varOmega } \end{aligned}$$
(119)
$$\begin{aligned}&D_i\left[ \frac{\partial \rho _i({\tilde{{\varvec{x}}}},t)}{\partial n}+\frac{z_ie}{kT}\rho _i({\tilde{{\varvec{x}}}},t)\frac{\partial \phi ({\tilde{{\varvec{x}}}},t)}{\partial n}\right] =\,0\quad \text{ for }\ {\tilde{{\varvec{x}}}}\in \partial \tilde{\varOmega } \end{aligned}$$
(120)
$$\begin{aligned}&\rho _i({\tilde{{\varvec{x}}}},0)=\,q_i({\tilde{{\varvec{x}}}})\quad \text{ for }\ {\tilde{{\varvec{x}}}}\in \tilde{\varOmega }, \end{aligned}$$
(121)

where \(z_i\) is the valence and \(D_i\) is the diffusion coefficient for the ion specie i. The electric potential \(\phi ({\tilde{{\varvec{x}}}},t)\) in \(\tilde{\varOmega }\) is solution of the Neumann problem for the Poisson equation

$$\begin{aligned} \Delta \phi ({\tilde{{\varvec{x}}}},t)&= -\frac{e}{\varepsilon _r\varepsilon _0}\left( \rho _p({\tilde{{\varvec{x}}}})- \rho _m({\tilde{{\varvec{x}}}})\right) \quad \text{ for }\ {\tilde{{\varvec{x}}}}\in \tilde{\varOmega }\nonumber \\ \frac{\partial \phi ({\varvec{x}},t)}{\partial n}&=-\tilde{\sigma }(\tilde{\varvec{x}},t)\quad \text{ for }\ \tilde{\varvec{x}}\in {\partial \tilde{\varOmega }}, \end{aligned}$$
(122)

where \(\tilde{\sigma }({\tilde{{\varvec{x}}}},t)\) is the surface charge density on the boundary \(\partial \tilde{\varOmega }\). At steady state, (119) gives

$$\begin{aligned} \rho _i({\tilde{{\varvec{x}}}})= \rho _{i,0}\exp \left( \displaystyle -\frac{z_i e\phi ({\tilde{{\varvec{x}}}})}{k_\mathrm{BT}} \right) \quad \text{ for } i\in \{p\,,\,m\}, \end{aligned}$$
(123)

where \(\rho _{i,0} \) is obtained from no-flux boundary condition (120), thus

$$\begin{aligned} \rho _i({\tilde{{\varvec{x}}}})= \frac{ N_i\exp \left( \displaystyle -\frac{z_i e\phi ({\tilde{{\varvec{x}}}})}{k_\mathrm{BT}} \right) }{\displaystyle \int _{\tilde{\varOmega }} \exp \left( \displaystyle -\frac{z_i e\phi ({\varvec{s}})}{k_\mathrm{BT}} \right) \mathrm{d}{\varvec{s}}}\quad \text{ for } i\in \{p,\,m\}. \end{aligned}$$
(124)

Using the nondimensionalized potential \(\displaystyle \tilde{u}({\tilde{{\varvec{x}}}})=\frac{e\,\phi ({\tilde{{\varvec{x}}}})}{k_\mathrm{BT}}\), Eq. (123) becomes

$$\begin{aligned} \rho _i({\tilde{{\varvec{x}}}})= \frac{ N_ie^{\displaystyle -z_i\, \tilde{u}({\tilde{{\varvec{x}}}})}}{ \displaystyle \int _{\tilde{\varOmega }} e^{\displaystyle -z_i\, \tilde{u}({\varvec{s}})}\mathrm{d}{\varvec{s}}}\quad \text{ for } i\in \{p,\,m\}. \end{aligned}$$
(125)

Using (122) and (125), we obtain

$$\begin{aligned} -\Delta \tilde{u}({\tilde{{\varvec{x}}}})= & {} \frac{l_\mathrm{B}N_pe^{\displaystyle - \tilde{u}({\tilde{{\varvec{x}}}})}}{ \int _{\tilde{\varOmega }} e^{\displaystyle - \tilde{u}({\varvec{s}})}\, \mathrm{d}{\varvec{s}}}-\frac{l_\mathrm{B} N_me^{\displaystyle \tilde{u}({\tilde{{\varvec{x}}}})}}{ \int _{\tilde{\varOmega }} e^{\displaystyle \tilde{u}({\varvec{s}})}\, \mathrm{d}{\varvec{s}}}\, \text{ in } \, \tilde{\varOmega }\nonumber \\ \frac{\partial u({\tilde{{\varvec{x}}}})}{\partial n}= & {} -\frac{(N_p-N_m) }{|\partial \tilde{\varOmega }| }l_\mathrm{B} \, \text{ on } \, \partial \tilde{\varOmega }, \end{aligned}$$
(126)

where \(l_\mathrm{B}\) is the Bjerrum length. Using \(\displaystyle {\varvec{x}}=\frac{{\tilde{{\varvec{x}}}}}{R_\mathrm{c}}\) and \(\tilde{u}(\tilde{x})=u(x)\) where \(R_\mathrm{c}\) is the cusp curvature radius, (126) becomes

$$\begin{aligned} -\Delta {u}({\varvec{x}})= & {} \frac{l_\mathrm{B}N_pe^{\displaystyle - {u}( {\varvec{x}})}}{R_\mathrm{c} \int _{ \varOmega } e^{\displaystyle - {u}({\varvec{s}})}\, \mathrm{d}{\varvec{s}}}-\frac{l_\mathrm{B} N_me^{\displaystyle {u}({\varvec{x}})}}{R_\mathrm{c} \int _{ \varOmega } e^{\displaystyle {u}({\varvec{s}})}\, \mathrm{d}{\varvec{s}}} \quad \text{ in } \, {\varOmega }\nonumber \\ \frac{\partial u( {\varvec{x}})}{\partial n}= & {} -\frac{l_\mathrm{B}(N_p-N_m) }{R_\mathrm{c}|\partial \varOmega | } \quad \text{ on } \, \partial {\varOmega }. \end{aligned}$$
(127)

The small parameter is \(\zeta =\displaystyle \frac{N_m}{N_p}\ll 1\) because in the bulk, the concentration of negative charges such as chloride (about 4 mM) is much smaller than positive ones (potassium and sodium account together roughly for 167 mM Hille 2001). A regular expansion of \(u({\varvec{x}})\) is

$$\begin{aligned} u({\varvec{x}})= & {} u_0({\varvec{x}})+\zeta u_1({\varvec{x}})+\cdots \end{aligned}$$
(128)

Using (128) in (127), in small \(\zeta \) limit, we have

$$\begin{aligned} -\Delta {u_0}({\varvec{x}})= & {} \frac{l_\mathrm{B} N_p e^{\displaystyle - {u_0}( {\varvec{x}})}}{R_\mathrm{c} \int _{ \varOmega } e^{\displaystyle - {u_0}({\varvec{s}})}\, \mathrm{d}{\varvec{s}}}\, \text{ in } \, {\varOmega }\nonumber \\ \frac{\partial u_0( {\varvec{x}})}{\partial n}= & {} -\frac{l_\mathrm{B} N_p }{R_\mathrm{c}|\partial \varOmega | }\, \text{ on } \, \partial {\varOmega }, \end{aligned}$$
(129)

and in \(\varOmega \)

$$\begin{aligned} \Delta {u_1}({\varvec{x}})= & {} \frac{l_\mathrm{B} N_{p} }{R_\mathrm{c} } \left( \frac{e^{\displaystyle - u_0({\varvec{x}})}}{\int _{ \varOmega } e^{\displaystyle - {u_0}({\varvec{s}})}\, \mathrm{d}{\varvec{s}}}\left( u_1({\varvec{x}})- \frac{\int _{ \varOmega } u_1({\varvec{s}})e^{\displaystyle - {u_0}({\varvec{s}})}\, \mathrm{d}{\varvec{s}}}{\int _{ \varOmega } e^{\displaystyle - {u_0}({\varvec{s}})}\, \mathrm{d}{\varvec{s}}} \right) + \frac{e^{\displaystyle u_0({\varvec{x}})}}{\int _{ \varOmega } e^{\displaystyle {u_0}({\varvec{s}})}\, \mathrm{d}{\varvec{s}}}\right) \nonumber \\ \frac{\partial u_1( {\varvec{x}})}{\partial n}= & {} \frac{l_\mathrm{B}N_{p} }{R_\mathrm{c}|\partial \varOmega | }\, \text{ on } \, \partial {\varOmega }, \end{aligned}$$
(130)

which admit a regular solution. This result shows that the limit of the PNP equation when \(\zeta \) tends to zero (small charge limit) gives \(v_0({\varvec{x}})\), and thus we conclude that a small amount of negative charges does not perturb the distribution of the total excess of positive charge in the bulk.

1.2 The Numerical Procedure

Numerical solutions were constructed by the COMSOL Multiphysics 5.0 (BVP problems), Maple 2015 (shooting problems) and MATLAB R2015 (conformal mapping). The boundary value problems in 1D, 2D and 3D were solved by the finite elements method in the COMSOL ‘Mathematics’ package. We used an adaptive mesh refinement to ensure numerical convergence for large value of the parameter \(\lambda \).

We solved the PDEs by the shooting procedure for boundary value problems using Runge–Kutta fourth-order method, as well as solvers from Maple packages.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cartailler, J., Schuss, Z. & Holcman, D. Geometrical Effects on Nonlinear Electrodiffusion in Cell Physiology. J Nonlinear Sci 27, 1971–2000 (2017). https://doi.org/10.1007/s00332-017-9393-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00332-017-9393-2

Keywords

Mathematics Subject Classification

Navigation