Skip to main content
Log in

Sensitivity of the electrical response of a node of Ranvier model to alterations of the myelin sheath geometry

  • Published:
Journal of Mathematical Biology Aims and scope Submit manuscript

Abstract

Nodes of Ranvier play critical roles in the generation and transmission of action potentials. Alterations in node properties during pathology and/or development are known to affect the speed and quality of electrical transmission. From a modelling standpoint, nodes of Ranvier are often described by systems of ordinary differential equations neglecting or greatly simplifying their geometric structure. These approaches fail to accurately describe how fine scale alteration in the node geometry or in myelin thickness in the paranode region will impact action potential generation and transmission. Here, we rely on a finite element approximation to describe the three dimensional geometry of a node of Ranvier. With this, we are able to investigate how sensitive is the electrical response to alterations in the myelin sheath and paranode geometry. We could in particular investigate irregular loss of myelin, which might be more physiologically relevant than the uniform loss often described through simpler modelling approaches.

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
Fig. 8
Fig. 9

Similar content being viewed by others

References

  • Adams RA, Fournier JJF (2003) Sobolev spaces. Pure and applied mathematics, vol 140. Academic Press, New York, London

    Google Scholar 

  • Arancibia-Carcamo IL, Attwell D (2014) The node of Ranvier in CNS pathology. Acta Neuropathol 128(2):161–175

    Article  Google Scholar 

  • Bathe K (1996) Finite element procedures. Prentice-Hall, New Jersey

    MATH  Google Scholar 

  • Belhamadia Y, Fortin A, Bourgault Y (2014) On the performance of anisotropic mesh adaptation for scroll wave turbulence dynamics in reaction–diffusion systems. J Comput Appl Math 271:233–246

    Article  MATH  Google Scholar 

  • Boahen F, Doyon N (2020) Modelling dendritic spines with the finite element method, investigating the impact of geometry on electric and calcic responses. J Math Biol 81(2):517–547

    Article  MATH  Google Scholar 

  • Bois R, Fortin M, Fortin A (2012) A fully optimal anisotropic mesh adaptation method based on a hierarchical error estimator. Comput Methods Appl Mech Eng 209–212:12–27

    Article  MATH  Google Scholar 

  • Bolintineanu DS, Sayyed-Ahmad A, Davis HT, Kaznessis YN (2009) Poisson–Nernst–Planck models of nonequilibrium ion electrodiffusion through a protegrin transmembrane pore. PLOS Comput Biol 5(1):1–12

    Article  Google Scholar 

  • Boucher P-A, Joós B, Morris CE (2012) Coupled left-shift of Nav channels: modeling the Na+-loading and dysfunctional excitability of damaged axons. J Comput Neurosci 33(2):301–319

    Article  MATH  Google Scholar 

  • Ciarlet P, Luneville E (2009) La méthode des éléments finis: de la théorie à la pratique. Concepts généraux. I. Cours (ENSTA). Les Presses de l’ENSTA

  • Dione I, Deteix J, Briffard T, Chamberland E, Doyon N (2016) Improved simulation of electrodiffusion in the node of Ranvier by mesh adaptation. PLoS ONE 11(8):e0161318

    Article  Google Scholar 

  • Dione I, Doyon N, Deteix J (2019) Sensitivity analysis of the Poisson Nernst–Planck equations: a finite element approximation for the sensitive analysis of an electrodiffusion model. J Math Biol 78(1):21–56

    Article  MATH  Google Scholar 

  • Gramse G, Dols-Perez A, Edwards MA, Fumagalli L, Gomila G (2013) Nanoscale measurement of the dielectric constant of supported lipid bilayers in aqueous solutions with electrostatic force microscopy. Biophys J 104(6):1257–1262

    Article  Google Scholar 

  • Hairer E, Nørsett SP, Wanner G (1993) Solving ordinary differential equations I (2nd Revised. Ed.): nonstiff problems, Revised. Springer-Verlag, New York

    MATH  Google Scholar 

  • Hille B (1992) Ionic channels of excitable membranes. Sinauer Associates, Sunderland

    Google Scholar 

  • Hodgkin AL, Huxley AF (1952) A quantitative description of membrane current and its application to conduction and excitation in nerve. J Physiol 117(4):500–544

    Article  Google Scholar 

  • Lopreore CL, Bartol TM, Coggan JS, Keller DX, Sosinsky GE, Ellisman MH, Sejnowski TJ (2008) Computational modeling of three-dimensional electrodiffusion in biological systems: application to the node of Ranvier. Biophys J 95(6):2624–2635

    Article  Google Scholar 

  • Lu B, Holst MJ, McCammon JA, Zhou Y (2010) Poisson–Nernst–Planck equations for simulating biomolecular diffusion–reaction processes I: finite element solutions. J Comput Phys 229(19):6979–6994

    Article  MATH  Google Scholar 

  • Peters A (1961) A radial component of central myelin sheaths. J Biophys Biochem Cytol 11(3):733–5

    Article  Google Scholar 

  • Pods J, Schonke J, Bastian P (2013) Electrodiffusion models of neurons and extracellular space using the Poisson–Nernst–Planck equations—numerical simulation of the intra- and extracellular potential for an axon model. Biophys J 105(1):242–254

    Article  Google Scholar 

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

    Article  MATH  Google Scholar 

  • Quarteroni AM, Valli A (2008) Numerical approximation of partial differential equations. Springer, Cham

    MATH  Google Scholar 

  • Rosenbluth J, Nave K, Mierzwa A, Schiff R (2006) Subtle myelin defects in PLP-null mice. Glia 54(3):172–82

    Article  Google Scholar 

  • Stadelmann C, Timmler S, Barrantes-Freer A, Simons M (2019) Myelin in the central nervous system: structure, function, and pathology. Physiol Rev 99(3):1381–1431

    Article  Google Scholar 

  • Susuki K (2013) Node of Ranvier disruption as a cause of neurological diseases. ASN Neuro 5(3):209–19

    Article  Google Scholar 

  • Sylantyev S, Savtchenko L, Ermolyuk Y, Michaluk P, Rusakov D (2013) Spike-driven glutamate electrodiffusion triggers synaptic potentiation via a homer-dependent mGluR-NMDAR link. Neuron 77(3):528–541

    Article  Google Scholar 

  • Tora A, Rovira X, Dione I, Bertrand H, Brabet I, De Koninck Y, Doyon N, Pin J, Acher F, Goudet C (2015) Allosteric modulation of metabotropic glutamate receptors by chloride ions. FASEB J 29(10):4174–4188

    Article  Google Scholar 

  • Volman V, Ng LJ (2014) Primary paranode demyelination modulates slowly developing axonal depolarization in a model of axonal injury. J Comput Neurosci 37(3):439–57

    Article  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Nicolas Doyon.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Jean Deteix and Nicolas Doyon are supported by the Natural Sciences and Engineering Research Council of Canada (http://www.nserc-crsng.gc.ca/).

Supplementary Information

Appendices

A Discrete version of the Poisson Nernst–Planck equations

We denote by \(L^{2}(\Omega )\) the space of square integrable functions defined on \(\Omega \). The Sobolev spaces \(H^{m}(\Omega )\), \(m \ge 0\), are the spaces of functions in \(L^{2}(\Omega )\) with generalized partial derivatives belonging to \(L^{2}(\Omega )\) up to order m (see Adams and Fournier 2003). The functional space \(U_V\) containing the ionic concentrations and electric potential is defined as

$$\begin{aligned} V = \left\{ v \in H^{1}(\Omega ): v = 0 \text{ on } \Gamma _{D}\right\} ,\quad U_V = L^{2}(0, T_F; V)\cap L^{\infty }(0, T_F; L^{\infty }(\Omega )). \end{aligned}$$

1.1 A.1 Weak formulation of the Poisson Nernst–Planck equations

Multiplying the system of Eq. (1) by test functions \(v, w \in H^{1}(\Omega )\) and integrating over the domain \(\Omega \), we obtain the following weak formulation

$$\begin{aligned} \left\{ \begin{aligned}&\int _{\Omega }\left( \frac{\partial c_{i}}{\partial t} + \nabla \cdot {\textbf{J}}_{i}(c_{i},\phi )\right) vdx = 0,\;\;i=1,\ldots ,n_c, \\&- \int _{\Omega }\left( \nabla \cdot (\varepsilon \nabla \phi ) + F \sum _{i = 1}^{n_{c}} z_{i}c_{i}\right) wdx = 0. \end{aligned} \right. \end{aligned}$$

Integrating by parts and taking into account (2), (4) and (5) we have \(\forall v,w \in V\),

$$\begin{aligned} \left\{ \begin{aligned}&\int _{\Omega }\frac{\partial c_{i}}{\partial t}vdx -\int _{\Omega }{\textbf{J}}_i(c_{i},\phi ) \cdot \nabla vdx \\&\quad = - \sum ^{n_{sd}}_{k=1}\int _{\partial \Omega _{k}}\left( {\textbf{J}}_{i}(c_{i},\phi )\cdot {\textbf{n}}_{k}\right) vds,\;\; i=1,\ldots ,n_c, \\&\int _{\Omega }\varepsilon \nabla \phi \cdot \nabla wdx -F\sum _{i = 1}^{n_{c}}\int _{\Omega }z_{i}c_{i}wdx = \sum ^{n_{sd}}_{k=1}\int _{\partial \Omega _{k}}\left( \varepsilon \nabla \phi \cdot {\textbf{n}}_{k}\right) wds. \end{aligned} \right. \end{aligned}$$

Since

$$\begin{aligned} \partial \Omega _k = \left( \partial \Omega \cap \partial \Omega _{k}\right) \cup \left( \partial \Omega _k\cap \Gamma \right) \quad \text{ and }\quad \left( \partial \Omega \cap \partial \Omega _{k}\right) \cap \left( \partial \Omega _k\cap \Gamma \right) = \emptyset \end{aligned}$$

taking into account (6) and introducing \({\tilde{D}}_{i} = D_{i}/\lambda _{i}\) we obtain the weak problem: find \(c_{i} \in U_V\), \(i = 1,\ldots , n_c\) and \(\phi \in U_V\) such that

$$\begin{aligned} \left\{ \begin{aligned}&\int _{\Omega }\frac{\partial c_{i}}{\partial t}vdx + \int _{\Omega }(D_{i}\nabla c_{i} + {\tilde{D}}_{i}c_{i}\nabla \phi )\cdot \nabla vdx \\ {}&\quad = - \int _{\Gamma }f_{i}vds,\; \forall \, v \in V,\;\; i=1,\ldots , n_c,\\&\int _{\Omega }\varepsilon \nabla \phi \cdot \nabla wdx -F\sum _{i = 1}^{n_{c}}\int _{\Omega }z_{i}c_{i}wdx = 0,\;\forall \, w \in V. \end{aligned} \right. \end{aligned}$$
(9)

1.2 A.2 Numerical approximations

In order to numerically approximate solutions of (9) we introduce a time discrete formulation based on a finite difference time scheme and combine it with the finite element method for the spatial approximation. We consider a second order backward difference formula with, for convenience and simplicity, a constant time-step \(\Delta t\). Denoting \(t^n = n\Delta t\) and \(\psi ^n = \psi (\cdot ,t^n)\), the time derivative of \(\psi \) at time \(t^{n+1}\) is approximated by

$$\begin{aligned} \frac{\partial \psi (\cdot ,t^{n+1})}{\partial t} \approx \frac{3\psi ^{n+1}-4\psi ^n+\psi ^{n-1}}{2\Delta t},\qquad n\ge 1. \end{aligned}$$

The first timestep being calculated using a backward Euler approximation.

For the spatial approximation, the finite element method is used. First, we introduce a mesh \({\mathcal {T}}_{h} = \{\top \}\) of simplicial \(\top \) partitioning \(\Omega \):

$$\begin{aligned} {\overline{\Omega }}_h = \bigcup _{\top \in {\mathcal {T}}_{h}}\top \,\subseteq {\overline{\Omega }}, \end{aligned}$$

with the usual restrictions on \({\mathcal {T}}_{h}\) (see Bathe 1996 or Ciarlet and Luneville 2009 for details). From the family \({\mathcal {T}}_{h}\) of triangulations of the domain \(\Omega \) (indexed by h), we construct the family of finite dimensional vector spaces \(V_{h}\). This set of standard continuous and piecewise functions is defined by

$$\begin{aligned} V_{h} = \left\{ v_h \in {\mathscr {C}}^{0}\left( \overline{\Omega _h}\right) ;\, {v_h}_{|\top } \in {\mathscr {P}}_{k}\left( \top \right) , \, \forall \, \top \in {\mathcal {T}}_{h}, \, v_h = 0 \, \text{ on } \Gamma _{D} \right\} \end{aligned}$$

where \({\mathscr {P}}_{k}(\top )\) is the space of polynomials of degree less or equal to k on \(\top \). An index "h" is used to denote a spatial approximation in \(V_h\). An approximation, in \(V_h\), of \(\psi \) at time \(t^n\) will be noted \(\psi ^n_h\).

Since an implicit time scheme is used for an approximation at time \(t^{n+1}\), all terms (apart form the time derivatives) will be expressed at time \(t^{n+1}\). Applied on (9), this results in sequences of strongly coupled (and possibly non linear) problems characterizing time-discrete unknown (ionic concentrations, electric potential, etc.). Details for the resolution of these non linear systems and the resulting algorithm, are presented in Dione et al. (2016).

1.3 A.3 Discrete formulation of the PN–P equations

To simplify the notation, let

$$\begin{aligned} f_{ih}^{n+1} = f_{i}(t^{n+1}, c^{n+1}_{ih}, \phi ^{n+1}_{h}). \end{aligned}$$

The implicit time scheme and finite element method applied on (9) produce a sequence of problems: for \(n \ge 1\), given \(c^{n-1}_{ih}\), \(c^{n}_{ih}\),\(\;\;i=1,\ldots ,n_c\), \(\phi ^{n-1}_{h}\) and \(\phi ^{n}_{h}\), find \(c^{n+1}_{ih} \in V_{h}\) and \(\phi ^{n+1}_{h} \in V_{h}\) such that

$$\begin{aligned} \left\{ \begin{aligned}&\int _{\Omega _{h}} \frac{3c^{n+1}_{ih} - 4c^{n}_{ih} + c^{n-1}_{ih}}{2\Delta t_n}v_{h}dx + \int _{\Omega _{h}}(D_{i}\nabla c^{n+1}_{ih}+{\tilde{D}}_{i}c^{n+1}_{ih}\nabla \phi ^{n+1}_{h})\cdot \nabla v_{h}dx \quad \\&\quad = - \int _{\Gamma _h}f_{ih}^{n+1}v_{h}ds, \; \forall \ v_{h} \in V_{h},\;\;i=1,\ldots ,n_c, \\&\int _{\Omega _{h}}\varepsilon \nabla \phi ^{n+1}_{h}\cdot \nabla w_{h}dx -F\sum _{k = 1}^{n_{c}} \int _{\Omega _{h}}z_{i}c^{n+1}_{ih}w_{h}dx = 0,\forall \; w_{h} \in V_{h}, \end{aligned} \right. \end{aligned}$$
(10)

where \(\Gamma _h\) is the discrete version of the internal interface \(\Gamma \). These systems are non linear, since they contain \({\tilde{D}}_{i}c^{n+1}_{ih}\nabla \phi ^{n+1}_{h}\), moreover, the right-hand side \(f_{ih}^{n+1}\), a priori, depends on \(c_{ih}^{n+1}\). We use a Newton–Raphson method for solving (10) (which amount to a first order Taylor development of the non linear terms), the reader is refered to Dione et al. (2016) for the complete details.

B Voltage-gated channels

The functions \(n(t, \phi )\), \(m(t, \phi )\) and \(h(t, \phi )\) are time and voltage-dependent gating functions, taking values in [0, 1] defined by the following ODEs

$$\begin{aligned} {\left\{ \begin{array}{ll} \displaystyle \frac{dn}{dt} = \alpha _n(\phi )(1-n) - \beta _{n}(\phi )n, &{} \\ \displaystyle \frac{dm}{dt} = \alpha _m(\phi )(1-m) - \beta _m(\phi )m, &{} \\ \displaystyle \frac{dh}{dt} = \alpha _h(\phi )(1-h) - \beta _h(\phi )h, &{} \end{array}\right. } \end{aligned}$$
(11)

where, denoting \(\phi _r\) the resting potential, the rate functions \(\alpha _{(\cdot )}\) and \(\beta _{(\cdot )}\) are defined by

$$\begin{aligned}{} & {} \phi _{r}=-55\; mV,\quad \gamma _v(\phi ) = \left( v + \phi _r - [\phi ]\right) /10 \\{} & {} \beta _n(\phi ) = \displaystyle 0.125 e^{\gamma _0(\phi )/8}, \quad \alpha _n(\phi ) = {\left\{ \begin{array}{ll} 0.1\displaystyle \gamma _{10}(\phi )\left( e^{\gamma _{10}(\phi )} - 1\right) ^{-1} &{} \text{ if } \; \displaystyle \gamma _{10}(\phi ) \ne 0,\\ 0.1 &{} \text{ else }, \end{array}\right. } \\{} & {} \beta _m(\phi )= \displaystyle 4e^{\gamma _0(\phi )/1.8}, \quad \alpha _m(\phi ) = {\left\{ \begin{array}{ll} \displaystyle \gamma _{25}(\phi )\left( e^{\gamma _{25}(\phi )} - 1\right) ^{-1} &{} \text{ if } \; \displaystyle \gamma _{25}(\phi ) \ne 0,\\ 1 &{} \text{ else } , \end{array}\right. } \\{} & {} \beta _h(\phi ) = \left( \displaystyle e^{\gamma _{30}(\phi )} + 1\right) ^{-1}, \quad \alpha _h(\phi ) = 0.07e^{\gamma _0(\phi )/2}. \end{aligned}$$

We use the same rate functions as in Pods et al. (2013) (under the assumption that the temperature factor \(C_T\) is equal to 1) which are themselves equivalent to the rate functions used in Lopreore et al. (2008). The notation is however modified to describe the resting potential \(\phi _r\) as a parameter that could eventually vary and to give explicitly the value of the rate functions when the denominators are equal to zero. Computing the flux in (7) requires to solve the ODE system (11).

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) 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

Boahen, F., Doyon, N. & Deteix, J. Sensitivity of the electrical response of a node of Ranvier model to alterations of the myelin sheath geometry. J. Math. Biol. 86, 17 (2023). https://doi.org/10.1007/s00285-022-01841-0

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00285-022-01841-0

Keywords

Mathematics Subject Classification

Navigation