Skip to main content
Log in

Flux Ratios and Channel Structures

  • Published:
Journal of Dynamics and Differential Equations Aims and scope Submit manuscript

Abstract

We investigate Ussing’s unidirectional fluxes and flux ratios of charged tracers motivated particularly by the insightful proposal of Hodgkin and Keynes on a relation between flux ratios and channel structure. Our study is based on analysis of quasi-one-dimensional Poisson–Nernst–Planck type models for ionic flows through membrane channels. This class of models includes the Poisson equation that determines the electrical potential from the charges present and is in that sense consistent. Ussing’s flux ratios generally depend on all physical parameters involved in ionic flows, particularly, on bulk conditions and channel structures. Certain setups of ion channel experiments result in flux ratios that are universal in the sense that their values depend on bulk conditions but not on channel structures; other setups lead to flux ratios that are specific in the sense that their values depend on channel structures too. Universal flux ratios could serve some purposes better than specific flux ratios in some circumstances and worse in other circumstances. We focus on two treatments of tracer flux measurements that serve as estimators of important properties of ion channels. The first estimator determines the flux of the main ion species from measurements of the flux of its tracer. Our analysis suggests a better experimental design so that the flux ratio of the tracer flux and the main ion flux is universal. The second treatment of tracer fluxes concerns ratios of fluxes and experimental setups that try to determine some properties of channel structure. We analyze the two widely used experimental designs of estimating flux ratios and show that the most widely used method depends on the spatial distribution of permanent charge so this flux ratio is specific and thus allows estimation of (some of) the properties of that permanent charge, even with ideal ionic solutions. The work presented in this paper is a first step showing how measurements of fluxes and flux ratios can give important insights into channel structure and function.

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.

Similar content being viewed by others

Notes

  1. The independence principle played a large part in the early thinking of Hodgkin, Huxley, and Katz [38,39,40,41,42,43,44] and was widely used by physiologists and biochemists to describe bulk solutions perhaps for that reason. Hodgkin was unaware of the evidence (personal communications ALH to RSE) that the Kohlraush principle of independent migration (\(\sim \)1880) had been disproven by measurements showing that properties of dilute sodium chloride solutions depended on the square root of concentration (because of the screening of the ionic atmosphere approximated by Debye-Hückel theory).

References

  1. Abaid, N., Eisenberg, R.S., Liu, W.: Asymptotic expansions of I–V relations via a Poisson–Nernst–Planck system. SIAM J. Appl. Dyn. Syst. 7, 1507–1526 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  2. Barcilon, V.: Ion flow through narrow membrane channels: Part I. SIAM J. Appl. Math. 52, 1391–1404 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  3. Barcilon, V., Chen, D.-P., Eisenberg, R.S.: Ion flow through narrow membrane channels: Part II. SIAM J. Appl. Math. 52, 1405–1425 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  4. Barcilon, V., Chen, D.-P., Eisenberg, R.S., Jerome, J.W.: Qualitative properties of steady-state Poisson–Nernst–Planck systems: perturbation and simulation study. SIAM J. Appl. Math. 57, 631–648 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  5. Bass, L., Bracken, A., Hilden, J.: Flux ratio theorems for nonstationary membrane transport with temporary capture of tracer. J. Theor. Biol. 118, 327–338 (1988)

    Article  Google Scholar 

  6. Bass, L., McNabb, A.: Flux ratio theorems for nonlinear membrane transport under nonstationary conditions. J. Theor. Biol. 133, 185–191 (1988)

    Article  Google Scholar 

  7. Begenisich, T., Busath, D.: Sodium flux ratio in voltage-clamped squid giant axons. J Gen. Physiol. 77(5), 489–502 (1981)

    Article  Google Scholar 

  8. Benos, D.J., Hyde, B.A., Latorre, R.: Sodium flux ratio through the amiloride-sensitive entry pathway in frog skin. J. Gen. Physiol. 81, 667–685 (1993)

    Article  Google Scholar 

  9. Boda, D., Busath, D., Eisenberg, B., Henderson, D., Nonner, W.: Monte Carlo simulations of ion selectivity in a biological \(\text{ Na }^+\) channel: charge–space competition. Phys. Chem. Chem. Phys. 4, 5154–5160 (2002)

    Article  Google Scholar 

  10. Boda, D., Busath, D.D., Henderson, D., Sokolowski, S.: Monte Carlo simulations of the mechanism of channel selectivity: the competition between volume exclusion and charge neutrality. J. Phys. Chem. B. 104, 8903–8910 (2000)

    Article  Google Scholar 

  11. Boda, D., Nonner, W., Valisko, M., Henderson, D., Eisenberg, B., Gillespie, D.: Steric selectivity in Na channels arising from protein polarization and mobile side chains. Biophys. J. 93, 1960–1980 (2007)

    Article  Google Scholar 

  12. Boda, D., Valisko, M., Eisenberg, B., Nonner, W., Henderson, D., Gillespie, D.: Effect of protein dielectric coefficient on the ionic selectivity of a calcium channel. J. Chem. Phys. 125, 034901(1-11) (2006)

    Article  Google Scholar 

  13. Boda, D., Valisko, M., Eisenberg, B., Nonner, W., Henderson, D., Gillespie, D.: The combined effect of pore radius and protein dielectric coefficient on the selectivity of a calcium channel. Phys. Rev. Lett. 98, 168102(1-4) (2007)

    Article  Google Scholar 

  14. Boda, D., Varga, T., Henderson, D., Busath, D., Nonner, W., Gillespie, D., Eisenberg, B.: Monte Carlo simulation study of a system with a dielectric boundary: application to calcium channel selectivity. Mol. Simul. 30, 89–96 (2004)

    Article  MATH  Google Scholar 

  15. Busath, D., Begenisich, T.: Unidirectional sodium and potassium fluxes through the sodium channel of squid giant axons. Biophys. J. 40, 41–49 (1982)

    Article  Google Scholar 

  16. Catterall, W.A.: Ion channel voltage sensors: structure, function, and pathophysiology. Neuron 67, 915–928 (2010)

    Article  Google Scholar 

  17. Catterall, W.A.: Molecular properties of voltage-sensitive sodium channels. Annu. Rev. Biochem. 55, 953–985 (1986)

    Article  Google Scholar 

  18. Chen, X.H., Bezprozvanny, I., Tsien, R.W.: Molecular basis of proton block of L-type \(\text{ Ca }^{2+}\) channels. J. Gen. Physiol. 108, 363–374 (1996)

    Article  Google Scholar 

  19. Crozier, P.S., Henderson, D., Rowley, R.L., Busath, D.D.: Model channel ion currents in NaCl-extended simple point charge water solution with applied-field molecular dynamics. Biophys. J. 81, 3077–3089 (2001)

    Article  Google Scholar 

  20. Crozier, P.S., Rowley, R.L., Holladay, N.B., Henderson, D., Busath, D.D.: Molecular dynamics simulation of continuous current flow through a model biological membrane channel. Phys. Rev. Lett. 86, 2467–2470 (2001)

    Article  Google Scholar 

  21. Curtis, B.A., Eisenberg, R.S.: Calcium influx in contracting and paralyzed frog twitch muscle fibers. J. Gen. Physiol. 85, 383–408 (1985)

    Article  Google Scholar 

  22. Doyle, D.A., Cabral, J.M., Pfuetzner, R.A., Kuo, A., Gulbis, J.M., Cohen, S.L., Chait, B.T., MacKinnon, R.: The structure of the potassium channel: molecular basis of \(\text{ K }^+\) conduction and selectivity. Science 280, 69–77 (1998)

    Article  Google Scholar 

  23. Eisenberg, B.: Ionic interactions are everywhere. Physiology 28, 28–38 (2013)

    Article  Google Scholar 

  24. Eisenberg, B.: Interacting ions in biophysics: real is not ideal. Biophys. J. 104, 1849–1866 (2013)

    Article  Google Scholar 

  25. Eisenberg, B.: Crowded charges in ion channels. In: Rice, S.A. (ed.) Advances in Chemical Physics, pp. 77–223. Wiley, New York (2011)

    Chapter  Google Scholar 

  26. Eisenberg, B.: Proteins, channels, and crowded ions. Biophys. Chem. 100, 507–517 (2003)

    Article  Google Scholar 

  27. Eisenberg, R.S.: From structure to function in open ionic channels. J. Membr. Biol. 171, 1–24 (1999)

    Article  Google Scholar 

  28. Eisenberg, B., Liu, W.: Poisson–Nernst–Planck systems for ion channels with permanent charges. SIAM J. Math. Anal. 38, 1932–1966 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  29. Eisenberg, B., Liu, W., Xu, H.: Reversal charge and reversal potential: case studies via classical Poisson–Nernst–Planck models. Nonlinearity 28, 103–128 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  30. Favre, I., Moczydlowski, E., Schild, L.: On the structural basis for ionic selectivity among \(\text{ Na }^+, \text{ K }^+\), and \(\text{ Ca }^{2+}\) in the voltage-gated sodium channel. Biophys. J. 71, 3110–3125 (1996)

    Article  Google Scholar 

  31. Gillespie, D.: A review of steric interactions of ions: why some theories succeed and others fail to account for ion size. Microfluid Nanofluidics 18, 717–738 (2015)

    Article  Google Scholar 

  32. Gillespie, D.: A Singular Perturbation Analysis of the Poisson–Nernst–Planck System: Applications to Ionic Channels. Ph.D. Dissertation, Rush University at Chicago (1999)

  33. Gillespie, D., Nonner, W., Henderson, D., Eisenberg, R.S.: A physical mechanism for large-ion selectivity of ion channels. Phys. Chem. Chem. Phys. 4, 4763–4769 (2002)

    Article  Google Scholar 

  34. Gillespie, D., Nonner, W., Eisenberg, R.S.: Coupling Poisson–Nernst–Planck and density functional theory to calculate ion flux. J. Phys. Condens. Matter 14, 12129–12145 (2002)

    Article  Google Scholar 

  35. Gillespie, D., Nonner, W., Eisenberg, R.S.: Density functional theory of charged, hard-sphere fluids. Phys. Rev. E 68, 0313503(1-10) (2003)

    Article  Google Scholar 

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

    Google Scholar 

  37. Hille, B.: Transport across cell membranes: carrier mechanisms, Chapter 2. In: Patton, H.D., Fuchs, A.F., Hille, B., Scher, A.M., Steiner, R.D. (eds.) Textbook of Physiology. Philadelphia, Saunders 1, 24–47 (1989)

  38. Hodgkin, A.L.: The ionic basis of electrical activity in nerve and muscle. Biol. Rev. 26, 339–409 (1951)

    Article  Google Scholar 

  39. Hodgkin, A.L.: Ionic movements and electrical activity in giant nerve fibres. Proc. R. Soc. Lond. Ser. B 148, 1–37 (1958)

    Article  Google Scholar 

  40. Hodgkin, A.L., Huxley, A.F.: Currents carried by sodium and potassium ions through the membrane of the giant axon of Loligo. J. Physol. 116, 449–472 (1952)

    Google Scholar 

  41. Hodgkin, A.L., Huxley, A.F.: The components of membrane conductance in the giant axon of Loligo. J. Physiol. 116, 473–496 (1952)

    Article  Google Scholar 

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

    Article  Google Scholar 

  43. Hodgkin, A.L., Huxley, A., Katz, B.: Ionic currents underlying activity in the giant axon of the squid. Arch. Sci. Physiol. 3, 129–150 (1949)

    Google Scholar 

  44. Hodgkin, A.L., Katz, B.: The effect of sodium ions on the electrical activity of the giant axon of the squid. J. Physiol. 108, 37–77 (1949)

    Article  Google Scholar 

  45. Hodgkin, A.L., Keynes, R.D.: The potassium permeability of a giant nerve fibre. J. Physiol. 128, 61–88 (1955)

    Article  Google Scholar 

  46. Horowicz, P., Gage, P.W., Eisenberg, R.S.: The role of the electrochemical gradient in determining potassium fluxes in frog striated muscle. J. Gen. Physiol. 51, 193s–203s (1968)

    Google Scholar 

  47. Im, W., Beglov, D., Roux, B.: Continuum solvation model: electrostatic forces from numerical solutions to the Poisson–Boltzmann equation. Comput. Phys. Commun. 111, 59–75 (1998)

    Article  MATH  Google Scholar 

  48. Im, W., Roux, B.: Ion permeation and selectivity of OmpF porin: a theoretical study based on molecular dynamics, Brownian dynamics, and continuum electrodiffusion theory. J. Mol. Biol. 322, 851–869 (2002)

    Article  Google Scholar 

  49. Jackson, M.: Molecular and Cellular Biophysics. University Press, Cambridge (2006)

    Book  Google Scholar 

  50. Ji, S., Liu, W.: Poisson–Nernst–Planck systems for ion flow with density functional theory for hard-sphere potential: I–V relations and critical potentials. Part I: analysis. J. Dyn. Differ. Equ. 24, 955–983 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  51. Ji, S., Liu, W., Zhang, M.: Effects of (small) permanent charge and channel geometry on ionic flows via classical Poisson–Nernst–Planck models. SIAM J. Appl. Math. 75, 114–135 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  52. Kedem, O., Essig, A.: Isotope flows and flux ratios in biological membranes. J. Gen. Physiol. 48(6), 1047 (1965)

    Article  Google Scholar 

  53. Kurnikova, M.G., Coalson, R.D., Graf, P., Nitzan, A.: A lattice relaxation algorithm for 3D Poisson–Nernst–Planck theory with application to ion transport through the gramicidin A channel. Biophys. J. 76, 642–656 (1999)

    Article  Google Scholar 

  54. Lin, G., Liu, W., Yi, Y., Zhang, M.: Poisson–Nernst–Planck systems for ion flow with a local hard-sphere potential for ion sizes. SIAM J. Appl. Dyn. Syst. 12, 1613–1648 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  55. Liu, W.: Geometric singular perturbation approach to steady-state Poisson–Nernst–Planck systems. SIAM J. Appl. Math. 65, 754–766 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  56. Liu, W.: One-dimensional steady-state Poisson–Nernst–Planck systems for ion channels with multiple ion species. J. Differ. Equ. 246, 428–451 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  57. Liu, W., Wang, B.: Poisson–Nernst–Planck systems for narrow tubular-like membrane channels. J. Dyn. Diff. Equ. 22, 413–437 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  58. Liu, W., Tu, X., Zhang, M.: Poisson–Nernst–Planck systems for ion flow with density functional theory for hard-sphere potential: I–V relations and critical potentials. Part II: numerics. J. Dyn. Differ. Equ. 24, 985–1004 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  59. Liu, W., Xu, H.: A complete analysis of a classical Poisson–Nernst–Planck model for ionic flow. J. Differ. Equ. 258, 1192–1228 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  60. Long, S.B., Campbell, E.B., Mackinnon, R.: Crystal structure of a mammalian voltage-dependent Shaker family \(\text{ K }^+\) channel. Science 309, 897–903 (2005)

    Article  Google Scholar 

  61. Long, S.B., Tao, X., Campbell, E.B., MacKinnon, R.: Atomic structure of a voltage-dependent \(\text{ K }^+\) channel in a lipid membrane-like environment. Nature 450, 376–382 (2007)

    Article  Google Scholar 

  62. McNabb, A., Bass, L.: Flux-ratio theorems for nonlinear equations of generalized diffusion. IMA J. Appl. Math. 43, 1–9 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  63. McNabb, A., Bass, L.: Flux theorems for linear multicomponent diffusion. IMA J. Appl. Math. 44, 155–161 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  64. Nadler, B., Schuss, Z., Singer, A., Eisenberg, B.: Diffusion through protein channels: from molecular description to continuum equations. Nanotech 3, 439–442 (2003)

    Google Scholar 

  65. Naranjo, D., Moldenhauer, H., Pincuntureo, M., Díaz-Franulic, I.: Pore size matters for potassium channel conductance. J. Gen. Physiol. 148, 277–291 (2016)

    Article  Google Scholar 

  66. Nonner, W., Eisenberg, B.: Electrodiffusion in ionic channels of biological membranes. J. Mol. Fluids 87, 149–162 (2000)

    Google Scholar 

  67. Nonner, W., Eisenberg, R.S.: Ion permeation and glutamate residues linked by Poisson–Nernst–Planck theory in L-type Calcium channels. Biophys. J. 75, 1287–1305 (1998)

    Article  Google Scholar 

  68. Noskov, S.Y., Im, W., Roux, B.: Ion permeation through the \(z_1\)-hemolysin channel: theoretical studies based on Brownian dynamics and Poisson–Nernst–Planck electrodiffusion theory. Biophys. J. 87, 2299–2309 (2004)

    Article  Google Scholar 

  69. Noskov, S.Y., Roux, B.: Ion selectivity in potassium channels. Biophys. Chem. 124, 279–291 (2006)

    Article  Google Scholar 

  70. Park, J.-K., Jerome, J.W.: Qualitative properties of steady-state Poisson–Nernst–Planck systems: mathematical study. SIAM J. Appl. Math. 57, 609–630 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  71. Payandeh, J., Scheuer, T., Zheng, N., Catterall, W.A.: The crystal structure of a voltage-gated sodium channel. Nature 475, 353–358 (2011)

    Article  Google Scholar 

  72. Rakowski, R.F., Gadsby, D.C., De Weer, P.: Single ion occupancy and steady-state gating of Na channels in squid giant axon. J. Gen. Physiol. 119, 235–249 (2002)

    Article  Google Scholar 

  73. Rosenfeld, Y.: Free energy model for the inhomogeneous fluid mixtures: Yukawa-charged hard spheres, general interactions, and plasmas. J. Chem. Phys. 98, 8126–8148 (1993)

    Article  Google Scholar 

  74. 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, 1–14 (2001)

    Article  Google Scholar 

  75. Sun, L., Liu, W.: Boundary value problems of Poisson–Nernst–Planck systems with nonlocal excess potentials: a case study. J. Dyn. Differ. Equ. (2017). doi:10.1007/s10884-017-9578-2

  76. Teorell, T.: Membrane electrophoresis in relation to bio-electrical polarization effects. Arch. Sci. Physiol. 3, 205–218 (1949)

    Google Scholar 

  77. Tosteson, D.: Membrane Transport: People and Ideas. American Physiological Society, Bethesda (1989)

    Book  Google Scholar 

  78. Ussing, H.H.: The distinction by means of tracers between active transport and diffusion. Acta Physiol. Scand. 19, 43–56 (1949)

    Article  Google Scholar 

  79. Yang, J., Ellinor, P.T., Sather, W.A., Zhang, J.F., Tsien, R.W.: Molecular determinants of \(\text{ Ca }^{2+}\) selectivity and ion permeation in L-type \(\text{ Ca }^{2+}\) channels. Nature 366, 158–161 (1993)

    Article  Google Scholar 

  80. Ye, S., Li, Y., Jiang, Y.: Novel insights into \(\text{ K }^+\) selectivity from high-resolution structures of an open \(\text{ K }^+\) channel pore. Nat. Struct. Mol. Biol. 17, 1019–1023 (2010)

    Article  Google Scholar 

  81. Yue, L., Navarro, B., Ren, D., Ramos, A., Clapham, D.E.: The cation selectivity filter of the bacterial sodium channel, NaChBac. J. Gen. Physiol. 120, 845–853 (2002)

    Article  Google Scholar 

Download references

Acknowledgements

We thank the anonymous reviewer for his/her invaluable comments that help improve the manuscript. This work is partially supported by the University of Kansas GRF (Grant No. #2301055) and NSF of China (Grant Nos. 11322105 and 11671071).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Weishi Liu.

Additional information

Dedicated to the memory of Professor George R. Sell.

Appendix: Derivations of Results in Sect. 5.2

Appendix: Derivations of Results in Sect. 5.2

In this section, we provide a derivation of formula (5.14) in Theorem 5.1 and formula (5.18) in Theorem 5.2 from the basic formula (4.3) in Theorem 4.1 for flux ratios from the one-isotope-setup (Setup 1). The derivation for both formulas are similar so we will provide the detailed derivation for formula (5.14) in Sect. 7.2 and comment on differences in the derivation for formula (5.18) in Sect. 7.3.

1.1 A Reformulation of (4.3) for Cases in Sect. 5.2

We will first recall the setup of the case specified in Sect. 5.2 and give a new form of formula (4.3) for the special case, which is convenient for us to apply previously established results in late part.

Recall that, for the one-isotope-setup (Setup 1), one tracer is used for two experimental settings with two different sets of boundary conditions for tracers. For one experiment, the tracer is injected from the left boundary with concentration \(L_1^{[t,i]}=\rho \) and zero right boundary concentration \(R_1^{[t,i]}=0\) that produces the influx \(J_1^{[t,i]}\). In another experiment, the same tracer is injected from the right boundary with concentration \(R_1^{[t,o]}=\rho \) and zero left boundary concentration \(L_1^{[t,o]}=0\) that produces the efflux \(J_1^{[t,o]}\).

Formula (4.3) in Theorem 4.1 for the flux ratio between \(J_1^{[t,i]}\) and \(J_1^{[t,o]}\) is

$$\begin{aligned} \begin{aligned} \frac{J_1^{[t,i]}}{J_1^{[t,o]}} =-\frac{L_1^{[t,i]}e^{p_1^{[i]}(0;z_1,d_1)}}{R_1^{[t,o]}e^{p_1^{[o]}(1;z_1,d_1)}}\; \int _0^1\frac{e^{p_1^{[o]}(x;z_1,d_1)}}{D_1(x)h(x)}dx\;\Big (\int _0^1\frac{e^{p_1^{[i]}(x;z_1,d_1)}}{D_1(x)h(x)}dx\Big )^{-1}, \end{aligned} \end{aligned}$$
(7.1)

where \(p_1^{[i]}(x;z_1,d_1)\) defined in (2.7) is determined by the profiles of concentrations from the solution of BVP associated with the boundary condition (BVi), and \(p_1^{[o]}(x;z_1,d_1)\) defined in (2.7) is determined by the profiles of concentrations from the solution of BVP associated with the boundary condition (BVo).

Remark 7.1

Note that, as \(\rho \rightarrow 0\), one has \(p_1^{[i]}(x;z_1,d_1)=p_1^{[o]}(x;z_1,d_1)\), and hence,

$$\begin{aligned} \lim _{\rho \rightarrow 0}\frac{J_1^{[t,i]}}{J_1^{[t,o]}} =-\frac{L_1^{[t,i]}e^{p_1^{[i]}(0;z_1,d_1)}}{R_1^{[t,o]}e^{p_1^{[o]}(1;z_1,d_1)}}; \end{aligned}$$

in particular, this limit does not contain information on permanent charge Q. An technical implication is that all terms involving \(Q_0\) in the approximation of \({J_1^{[t,i]}}/{J_1^{[t,o]}}\) should have a factor of \(\rho \). This explains why there is no \(dQ_0\)-term in the leading order expansions of \({J_1^{[t,i]}}/{J_1^{[t,o]}}\) in formula (5.14) in Theorem 5.1.

For the case studied in Sect. 5.2, we are able to obtain an approximation for the flux ratio, up to leading orders, explicitly in terms of given quantities of the problem.

It follows from (3.1) in Proposition 3.1 that

$$\begin{aligned} \Big (\int _0^1\frac{e^{p_1^{[i]}(x;z_1,d_1)}}{D_1(x)h(x)}dx\Big )^{-1} =\frac{J_1^{[i]}}{(L_1+L_1^{[t,i]})e^{p_1^{[i]}(0;z_1,d_1)}-R_1e^{p_1^{[i]}(1;z_1,d_1)}}, \end{aligned}$$

where \(J_1^{[i]}=J_1^{[m,i]}+J_1^{[t,i]}\) is the sum of the fluxes of the main ion species and its tracer. Similarly,

$$\begin{aligned} \int _0^1\frac{e^{p_1^{[o]}(x;z_1,d_1)}}{D_1(x)h(x)}dx=\frac{L_1e^{p_1^{[o]}(0;z_1,d_1)}-(R_1+R_1^{[t,o]})e^{p_1^{[o]}(1;z_1,d_1)}}{J_1^{[o]}}, \end{aligned}$$

where \(J_1^{[o]}=J_1^{[m,o]}+J_1^{[t,o]}\) is the sum of the fluxes of the main ion species and its tracer. Therefore,

$$\begin{aligned} \frac{J_1^{[t,i]}}{J_1^{[t,o]}}=-(I)\cdot (II)\cdot \frac{J_1^{[i]}}{J_1^{[o]}} \end{aligned}$$
(7.2)

where the factors (I) and (II) are given by

$$\begin{aligned} (I)&=\frac{L_1^{[t,i]}e^{p_1^{[i]}(0;z_1,d_1)}}{R_1^{[t,o]}e^{p_1^{[o]}(1;z_1,d_1)}},\nonumber \\ (II)&= \frac{L_1e^{p_1^{[o]}(0;z_1,d_1)}-(R_1+R_1^{[t,o]})e^{p_1^{[o]}(1;z_1,d_1)}}{(L_1+L_1^{[t,i]})e^{p_1^{[i]}(0;z_1,d_1)}-R_1e^{p_1^{[i]}(1;z_1,d_1)}}. \end{aligned}$$
(7.3)

The factors (I) and (II) on the right-hand side of (7.2) are determined by boundary conditions and are independent of the permanent charge. We will first determine these two factors and determine the last factor \({J_1^{[i]}}/{J_1^{[o]}}\) afterwards.

1.2 Derivation of Formula (5.14)

For this formula, the electroneutrality boundary conditions (5.13) are assumed to hold among all ion species including the tracer. It follows from (5.6) that, for the boundary condition associated to (BVi) in (5.13),

$$\begin{aligned} p_1^{[i]}(0;z_1,d_1)&=z_1\zeta _0V_0+ d\Big (2(L_1+\rho )+(1+\lambda ) (L_2-z_1\rho /{z_2})\Big ) +O(d^2),\\&=z_1\zeta _0V_0+ d\Big (2L_1+(1+\lambda ) L_2\Big ) +O(d^2,d\rho ),\\ p_1^{[i]}(1;z_1,d_1)&= d\Big (2R_1+(1+\lambda ) R_2\Big ) +O(d^2); \end{aligned}$$

and, for the boundary condition associated to (BVo) in (5.13),

$$\begin{aligned} p_1^{[o]}(0;z_1,d_1) =&\,z_1\zeta _0V_0+ d(2L_1+(1+\lambda ) L_2) +O(d^2),\\ p_1^{[o]}(1;z_1,d_1) =&\, d(2(R_1+\rho )+(1+\lambda ) (R_2-z_1\rho /{z_2})) +O(d^2)\\ =&\, d(2R_1+(1+\lambda ) R_2) +O(d^2,d\rho ). \end{aligned}$$

Thus, the factor (I) on the right-hand side of (7.2) is

$$\begin{aligned} \begin{aligned} (I)=&\,e^{p_1^{[i]}(0;z_1,d_1)-p_1^{[o]}(1,z_1,d_1)}\\ =&\,e^{z_1\zeta _0V_0} \Big (1+\big (2 (L_1-R_1) +(1+\lambda ) (L_2- R_2) \big )d\Big ) +O(d^2,d\rho ), \end{aligned} \end{aligned}$$
(7.4)

and the factor (II) on the right-hand side of (7.2) is, with \(s=L_1/{R_1}\),

$$\begin{aligned} (II)&=\frac{L_1e^{z_1\zeta _0V_0}-(R_1+\rho )+\left[ (2L_1^2+(1+\lambda )L_1L_2) e^{z_1\zeta _0V_0}-(2R_1^2+(1+\lambda ) R_1R_2) \right] d}{(L_1+\rho )e^{z_1\zeta _0V_0}-R_1 +\left[ (2L_1^2+(1+\lambda )L_1L_2)e^{z_1\zeta _0V_0} -(2R_1^2+(1+\lambda )R_1R_2)\right] d}\nonumber \\&\quad +O(\rho ^2, \rho d, d^2)\nonumber \\&= 1 -\frac{ e^{z_1\zeta _0V_0}+1}{se^{z_1\zeta _0V_0}-1}\frac{\rho }{R_1} +O(\rho ^2, \rho d, d^2). \end{aligned}$$
(7.5)

It remains to approximate the factor \(J_1^{[i]}/{J_1^{[o]}}\) in (7.2) that depends on the permanent charge Q and hence the full profile of the electric potential and the ion concentrations. Its approximation is complicated even for the simple case we considered in Sect. 5. We start with a general expression

$$\begin{aligned} \frac{J_1^{[i]}}{J_1^{[o]}}=S_0(\rho )+S_1(\rho ) d+S_2(\rho ) Q_0+O(d^2, \rho d Q_0, \rho Q_0^2), \end{aligned}$$

and determine the quantities \(S_k(\rho )\)’s for \(k=0,1,2\). In particular, we are interested in \(S_0(\rho )\) up to \(O(\rho )\) order, \(S_1(\rho )=S_1(0)+O(\rho )\) due to the factor d in \(S_1(\rho )d\), and \(S_2(\rho )\) up to \(O(\rho )\). The reason for the latter is, from Remark 7.1, any term involving \(Q_0\) should have a factor \(\rho \). The advantage of this expression is as follows. One can assume the ionic solution is ideal to get the term \(S_0(\rho ) +S_2(\rho ) Q_0\) (Sect. 7.2.1) and then assume the ionic solution is nonideal but the permanent charge is zero to get \(S_0(\rho )+S_1(\rho ) d\) (Sect. 7.2.2). The superposition of them then provides the approximation for \(J_1^{[i]}/{J_1^{[o]}}\).

1.2.1 Case with Ideal Ionic Solution for \(S_0(\rho ) +S_2(\rho ) Q_0\)

We will apply results from [51] for the classical PNP for this case.

It follows from results in [51] (displays (4.4) and (4.5) in [51]) that the influx of the tracer for the boundary concentrations in (BVi) is

$$\begin{aligned} J_1^{[i]}=&J_{10}^{[i]}+J_{11}^{[i]}Q_0+O(Q_0^2), \end{aligned}$$
(7.6)

where

$$\begin{aligned} \begin{aligned} J_{10}^{[i]}&= \frac{D_1( L_1+\rho -R_1)}{H(1)(\ln (L_1+\rho )-\ln R_1)}\frac{ \mu _1^{\delta ,i}}{k_BT},\\ J_{11}^{[i]}&=\frac{D_1A^{[i]}(z_2(1-B^{[i]})\zeta _0 V_0 +\ln (L_1+\rho )-\ln R_1 )}{(z_1-z_2)H(1)\big (\ln (L_1+\rho )-\ln R_1\big )^2}\frac{ \mu _1^{\delta ,i}}{k_BT}, \end{aligned} \end{aligned}$$
(7.7)

and

$$\begin{aligned} \frac{1}{k_BT} \mu _1^{\delta ,i} :=\frac{1}{k_BT}\left( \mu ^{[i]}(0)-\mu ^{[i]}(1)\right) = z_1 \zeta _0V_0+ \ln (L_1+\rho )-\ln R_1, \end{aligned}$$
(7.8)

is the difference between the boundary electrochemical potentials for (BVi), and

$$\begin{aligned} \begin{aligned} A^{[i]}&=-\frac{(\beta -\alpha )(L_1+\rho -R_1)^2}{((1-\alpha )(L_1+\rho )+\alpha R_1)((1-\beta )(L_1+\rho )+\beta R_1)\ln \frac{L_1+\rho }{ R_1}}, \\ B^{[i]}&=\frac{1}{A^{[i]}}\ln \frac{ (1-\beta )(L_1+\rho )+\beta R_1}{(1-\alpha )(L_1+\rho )+\alpha R_1}. \end{aligned} \end{aligned}$$
(7.9)

Similarly, the efflux of the tracer for the boundary concentrations in (BVo) is

$$\begin{aligned} J_1^{[o]}= J_{10}^{[o]}+J_{11}^{[o]}Q_0+O(Q_0^2), \end{aligned}$$
(7.10)

where

$$\begin{aligned} \begin{aligned} J_{10}^{[o]}&= \frac{D_1( L_1-(R_1+\rho ))}{H(1)(\ln L_1-\ln (R_1+\rho ))}\frac{ \mu _1^{\delta ,o}}{k_BT}, \\ J_{11}^{[o]}&=\frac{D_1A^{[o]}(z_2(1-B^{[o]})\zeta _0 V_0 +\ln L_1-\ln (R_1+\rho ) ) }{(z_1-z_2)H(1)\big (\ln L_1-\ln (R_1+\rho )\big )^2}\frac{ \mu _1^{\delta ,o}}{k_BT}, \end{aligned} \end{aligned}$$
(7.11)

and

$$\begin{aligned} \frac{1}{k_BT} \mu _1^{\delta ,o}:=\frac{1}{k_BT}\left( \mu ^{[o]}(0)-\mu ^{[o]}(1)\right) = z_1 \zeta _0V_0+ \ln L_1-\ln (R_1+\rho ) \end{aligned}$$
(7.12)

is the differences between the boundary electrochemical potentials (BVo), and

$$\begin{aligned} \begin{aligned} A^{[o]}&=-\frac{(\beta -\alpha )(L_1-R_1-\rho )^2}{((1-\alpha )L_1+\alpha (R_1+\rho ))((1-\beta )L_1+\beta (R_1+\rho ))\ln \frac{L_1}{R_1+\rho }}, \\ B^{[o]}&=\frac{1}{A^{[o]}}\ln \frac{ (1-\beta )L_1+\beta (R_1+\rho )}{(1-\alpha )L_1+\alpha (R_1+\rho )}. \end{aligned} \end{aligned}$$
(7.13)

Therefore,

$$\begin{aligned} \begin{aligned} \frac{J_1^{[i]}}{J_1^{[o]}}&= \frac{J_{10}^{[i]}+J_{11}^{[i]}Q_0}{J_{10}^{[o]}+J_{11}^{[o]}Q_0}+O(Q_0^2)\\&= F_0\left( 1+\frac{F_1}{F_0}Q_0\right) +O(Q_0^2), \end{aligned} \end{aligned}$$
(7.14)

where

$$\begin{aligned} \begin{aligned} F_0&=\frac{J_{10}^{[i]}}{J_{10}^{[o]}} = \frac{ L_1+\rho -R_1}{ L_1-(R_1+\rho )} \frac{\ln L_1-\ln (R_1+\rho )}{\ln (L_1+\rho )-\ln R_1}\frac{ \mu _1^{\delta ,i}}{ \mu _1^{\delta ,o}},\\ \frac{F_1}{F_0}&=\frac{J_{11}^{[i]}}{J_{10}^{[i]}}-\frac{J_{11}^{[o]}}{J_{10}^{[o]}} =\frac{ A^{[i]}\Big (z_2(1-B^{[i]}) \zeta _0V_0 +\ln (L_1+\rho )-\ln R_1\Big )}{(z_1-z_2)(L_1+\rho -R_1) \big (\ln (L_1+\rho )-\ln R_1\big )}\\&\qquad -\frac{ A^{[o]}\Big (z_2(1-B^{[o]}) \zeta _0V_0 +\ln L_1-\ln (R_1+\rho )\Big )}{(z_1-z_2)(L_1-(R_1+\rho )) \big (\ln L_1-\ln (R_1+\rho )\big )}. \end{aligned} \end{aligned}$$
(7.15)

Lemma 7.2

With \(s=L_1/{R_1}\), one has

$$\begin{aligned} F_0 = 1+ \left( f_1(s)+\frac{s+1}{s(z_1\zeta _0V_0+\ln s)}\right) \frac{\rho }{R_1}+O(\rho ^2), \end{aligned}$$

where \(f_1(s)\) is defined in (5.8), and

$$\begin{aligned} \frac{F_1}{F_0}=g_1(V_0,s,\alpha ,\beta )\frac{\rho }{R_1^2}+O(\rho ^2), \end{aligned}$$

where \(g_1(V_0,s,\alpha ,\beta )\) is given in (5.9). As a consequence,

$$\begin{aligned} S_0(\rho )=1+ \Big (f_1(s)+\frac{s+1}{s(z_1\zeta _0V_0+\ln s)}\Big )\frac{\rho }{R_1}\; \text{ and } \; S_2(\rho ) =g_1(V_0,s,\alpha ,\beta )\frac{\rho }{R_1^2}. \end{aligned}$$
(7.16)

Proof

For the term \(F_0\), it follows (7.15), (7.8) and (7.12) that

$$\begin{aligned} F_0&= \frac{s-1+\frac{\rho }{R_1}}{s-1- \frac{\rho }{R_1}}\cdot \frac{\ln s-\ln (1+\frac{\rho }{R_1})}{\ln s+\ln (1+\frac{\rho }{sR_1})}\frac{z_1\zeta _0V_0+\ln (s+\frac{\rho }{R_1})}{z_1\zeta _0V_0+\ln s-\ln (1+\frac{\rho }{R_1})},\\&=1+ \left( \frac{2s\ln s-s^2+1}{s(s-1)\ln s}+\frac{s+1}{s(z_1\zeta _0V_0+\ln s)}\right) \frac{\rho }{R_1}+O(\rho ^2), \end{aligned}$$

which is the claimed formula for \(F_0\).

The approximate formula for \(F_1/F_0\) is lengthy but otherwise straightforward.

With \(s=L_1/{R_1}\) and \(L_1^{[t,i]}=R_1^{[t,o]}=\rho \), it follows from (7.15) that

$$\begin{aligned} \frac{F_1}{F_0}&=\frac{ z_2(A^{[i]}-A^{[i]}B^{[i]}) \zeta _0V_0 }{(z_1-z_2)R_1(s-1+\rho /{ R_1})\big (\ln s+ \rho / (sR_1)\big )}\\&\quad +\frac{A^{[i]}}{(z_1-z_2)R_1(s-1+\rho /{ R_1})}\\&\quad -\frac{ z_2(A^{[o]}-A^{[o]}B^{[o]}) \zeta _0V_0 }{(z_1-z_2)R_1(s-1-\rho /{ R_1})\big (\ln s- \rho /{ R_1}\big )}\\&\quad -\frac{A^{[o]}}{(z_1-z_2)R_1(s-1-\rho /{ R_1})} +O(\rho ^2). \end{aligned}$$

Next, we fix \(A^{[i]}, B^{[i]}, A^{[o]}\) and \(B^{[o]}\), and expand the above in \(\rho \) to get

$$\begin{aligned} \frac{F_1}{F_0}&=\frac{ z_2 \zeta _0V_0 (A^{[i]}-A^{[i]}B^{[i]}-A^{[o]}+A^{[o]}B^{[o]}) }{(z_1-z_2)R_1(s-1) \ln s }\\&\quad +\frac{(s-1)(A^{[i]}-A^{[o]})}{(z_1-z_2)R_1(s-1)^2 }\\&\quad -\frac{ z_2 \zeta _0V_0[(A^{[i]}-A^{[i]}B^{[i]})(\ln s+1-1/s)+(A^{[o]}-A^{[o]}B^{[o]})(\ln s+s-1)]}{(z_1-z_2) (s-1 )^2(\ln s)^2} \frac{\rho }{ R_1^2}\\&\quad -\frac{A^{[i]}+A^{[o]}}{(z_1-z_2) (s-1)^2 }\frac{\rho }{ R_1^2} +O(\rho ^2). \end{aligned}$$

Now, we expand \(A^{[i]}, A^{[o]}, B^{[i]}\) and \(B^{[o]}\) defined in (7.9) and (7.13) in \(\rho \) as

$$\begin{aligned} A^{[i]}&=A_0^{[i]}+A_1^{[i]}\frac{\rho }{R_1}+O(\rho ^2),\\ A^{[o]}&=A_0^{[o]}+A_1^{[o]}\frac{\rho }{R_1}+O(\rho ^2),\\ A^{[i]}B^{[i]}&=(A^{[i]}B^{[i]})_0+(A^{[i]}B^{[i]})_1\frac{\rho }{R_1}+O(\rho ^2),\\ A^{[o]}B^{[o]}&=(A^{[o]}B^{[o]})_0+(A^{[o]}B^{[o]})_1\frac{\rho }{R_1}+O(\rho ^2 ). \end{aligned}$$

It follows directly that

$$\begin{aligned} \begin{aligned} A_0^{[i]}&=A_0^{[o]}=-(\beta -\alpha )\frac{(s-1)^2}{[(1-\alpha )s+\alpha ][(1-\beta )s+\beta ]\ln s},\\ A_1^{[i]}&=-(\beta -\alpha )\frac{2(s-1) }{[(1-\alpha )s+\alpha ][(1-\beta )s+\beta ]\ln s}\\&\quad +(\beta -\alpha )\frac{(s-1)^2[2(1-\alpha )(1-\beta )s+\alpha +\beta -2\alpha \beta ] }{[(1-\alpha )s+\alpha ]^2[(1-\beta )s+\beta ]^2\ln s}\\&\quad +(\beta -\alpha )\frac{(s-1)^2 }{[(1-\alpha )s+\alpha ][(1-\beta )s+\beta ]s(\ln s)^2},\\ A_1^{[o]}&=(\beta -\alpha )\frac{2(s-1) }{[(1-\alpha )s+\alpha ][(1-\beta )s+\beta ]\ln s}\\&\quad +(\beta -\alpha )\frac{(s-1)^2[(\alpha +\beta -2\alpha \beta )s+ 2\alpha \beta ] }{[(1-\alpha )s+\alpha ]^2[(1-\beta )s+\beta ]^2\ln s}\\&\quad -(\beta -\alpha )\frac{(s-1)^2 }{[(1-\alpha )s+\alpha ][(1-\beta )t+\beta ] (\ln s)^2},\\ (A^{[i]}B^{[i]})_0&=(A^{[o]}B^{[o]})_0=\ln \frac{(1-\beta )s+\beta }{(1-\alpha )s+\alpha },\\ (A^{[i]}B^{[i]})_1&= \frac{1-\beta }{(1-\beta ) s+\beta } - \frac{1-\alpha }{(1-\alpha ) s+\alpha },\\ (A^{[o]}B^{[o]})_1&=\frac{\beta }{(1-\beta ) s+\beta } - \frac{\alpha }{(1-\alpha ) s+\alpha }. \end{aligned} \end{aligned}$$
(7.17)

In particular,

$$\begin{aligned} A_0^{[i]}-A_0^{[o]}&=0,\quad (A^{[i]}B^{[i]})_0-(A^{[o]}B^{[o]})_0=0,\\ A_0^{[i]}-(A^{[i]}B^{[i]})_0&= -\frac{(\beta -\alpha )(s-1)^2}{[(1-\alpha )s+\alpha ][(1-\beta )s+\beta ]\ln s}-\ln \frac{(1-\beta )s+\beta }{(1-\alpha )s+\alpha },\\ A_1^{[i]}-A_1^{[o]}&=-\frac{4(\beta -\alpha )(s-1) }{[(1-\alpha )s+\alpha ][(1-\beta )s+\beta ]\ln s}\\&\quad +\frac{(\beta -\alpha )(s-1)^2[(2-3\alpha -3\beta +4\alpha \beta )s+\alpha +\beta -4\alpha \beta ] }{[(1-\alpha )s+\alpha ]^2[(1-\beta )s+\beta ]^2\ln s}\\&\quad +\frac{ (\beta -\alpha )(s+1)(s-1)^2}{[(1-\alpha )s+\alpha ][(1-\beta )s+\beta ]s(\ln s)^2},\\ (A^{[i]}B^{[i]})_1-(A^{[o]}B^{[o]})_1&=\frac{1-2\beta }{(1-\beta ) s+\beta } - \frac{1-2\alpha }{(1-\alpha )s+\alpha }. \end{aligned}$$

Combining the above estimates, one has

$$\begin{aligned} \frac{F_1}{ F_0}&=-\frac{4(\beta -\alpha )z_2 \zeta _0V_0 }{(z_1-z_2) [(1-\alpha )s+\alpha ][(1-\beta )s+\beta ](\ln s)^2}\frac{\rho }{(R_1)^2}\\&\quad +\frac{(\beta -\alpha )z_2 \zeta _0V_0(s-1) [(2-3\alpha -3\beta +4\alpha \beta )s+\alpha +\beta -4\alpha \beta ] }{(z_1-z_2) [(1-\alpha )s+\alpha ]^2[(1-\beta )s+\beta ]^2(\ln s)^2}\frac{\rho }{(R_1)^2}\\&\quad +\frac{ (\beta -\alpha )z_2 \zeta _0V_0(s+1)(s-1)}{(z_1-z_2) [(1-\alpha )s+\alpha ][(1-\beta )s+\beta ]s(\ln s)^3}\frac{\rho }{(R_1)^2}\\&\quad -\frac{ z_2 \zeta _0V_0 }{(z_1-z_2) (s-1)\ln s}\Big [\frac{1-2\beta }{(1-\beta )s+\beta } - \frac{1-2\alpha }{(1-\alpha )s+\alpha }\Big ]\frac{\rho }{(R_1)^2} \\&\quad +\frac{ (\beta -\alpha )z_2 \zeta _0V_0(2\ln s+s-1/s) }{(z_1-z_2)[(1-\alpha )s+\alpha ][(1-\beta )s+\beta ] (\ln s)^3} \frac{\rho }{(R_1)^2} \\&\quad +\frac{ z_2 \zeta _0V_0(2\ln s+s-1/s)}{(z_1-z_2) (s-1 )^2(\ln s)^2} \ln \frac{(1-\beta )s+\beta }{(1-\alpha )s+\alpha }\frac{\rho }{(R_1)^2} \\&\quad -\frac{(\beta -\alpha ) [(\alpha +\beta -2\alpha \beta )(s-1)^2+2s] }{(z_1-z_2)[(1-\alpha )s+\alpha ]^2[(1-\beta )s+\beta ]^2\ln s} \frac{\rho }{(R_1)^2}\\&\quad +\frac{ (\beta -\alpha )(s+1)(s-1)}{(z_1-z_2)[(1-\alpha )s+\alpha ][(1-\beta )s+\beta ]s(\ln s)^2} \frac{\rho }{(R_1)^2} +O(\rho ^2). \end{aligned}$$

The fifth term above can be split and combine with the first and the third terms to get the formula for \(F_1/{F_0}\) claimed in the statement of the lemma. \(\square \)

1.2.2 A Case with a Hard-Sphere Component for \(S_0(\rho )+S_1(\rho )d\)

In this section, we examine the flux ratio for PNP models with a hard-sphere component (point-charge with volume exclusion) and assume zero permanent charge.

Recall the local hard-sphere potential (5.5)

$$\begin{aligned} \frac{1}{k_BT}\mu _k^{LHS}(x)&= - \ln \Big (1-\sum _{j=1}^{n}d_jc_j(x)\Big )+ \frac{d_k\sum _{j=1}^nc_j(x)}{1-\sum _{j=1}^{n}d_jc_j(x)}\\&= \sum _{j=1}^{n}(d_j+d_k)c_j(x) +O(d_kd_j) \end{aligned}$$

where \(d_j\) is the diameter of the jth ion species.

For two ion species with diameters \(d_1\) and \(d_2\), set \(d=d_1\) and \(d_2=\lambda d\), and expand the fluxes as \(J_k=J_{k0}+J_{k1}d+O(d^2), k=1,2\). As observed in [29], the Nernst–Planck equations imply that the flux \(J_i\) is proportional to the transmembrane electrochemical potential \(\mu _i^{\delta }=\mu _i(0)-\mu _i(1)\) where

$$\begin{aligned} \begin{aligned} \frac{1}{k_BT}\mu _1^{\delta }&=z_1\zeta _0V_0 +\ln L_1-\ln R_1 \\&\quad +\frac{2z_2-(1+\lambda )z_1}{z_2} (L_1-R_1)d+O(d^2),\\ \frac{1}{k_BT}\mu _2^{\delta }&=z_2\zeta _0V_0 +\ln L_2-\ln R_2\\&\quad -\frac{(1+\lambda )z_2-2\lambda z_1}{z_1} (L_2-R_2)d+O(d^2). \end{aligned} \end{aligned}$$
(7.18)

It follows from results in [50, 54], for example, from Corollary 3.6 in [54] (with different but equivalent expressions) that the flux \(J_1^{[i]}\) is given by

$$\begin{aligned} J_1^{[i]} = \left( 1+ \frac{2(\lambda z_1-z_2)}{z_1z_2} w_1^{[i]} d\right) \frac{D_1w_0^{[i]}}{z_1H(1)}\frac{\mu _1^{\delta ,i}}{k_BT} +O(d^2), \end{aligned}$$

where, with \(\rho =L_1^{[t,i]}=R_1^{[t,o]}\),

$$\begin{aligned} \frac{1}{k_BT}\mu _1^{\delta ,i}&=z_1\zeta _0V_0 +\ln (L_1+\rho )-\ln R_1\\&\quad +\frac{2z_2-(1+\lambda )z_1}{z_2} (L_1+\rho - R_1)d+O(d^2) \end{aligned}$$

is the difference between the boundary electrochemical potentials for (BVi), and

$$\begin{aligned} w_0^{[i]}&=\frac{z_1(L_1+\rho - R_1)}{\ln (L_1+\rho )-\ln R_1}\; \text{ and } \\ w_1^{[i]}&= \frac{z_1(L_1+\rho - R_1)}{\ln (L_1+\rho )-\ln R_1}-\frac{z_1(L_1+\rho +R_1)}{2}. \end{aligned}$$

Similarly, the flux \(J_1^{[o]}\) is given by

$$\begin{aligned} J_1^{[o]} = \left( 1+ \frac{2(\lambda z_1-z_2)}{z_1z_2} w_1^{[o]} d\right) \frac{D_1w_0^{[o]}}{z_1H(1)}\frac{\mu _1^{\delta ,o}}{k_BT} +O(d^2), \end{aligned}$$

where

$$\begin{aligned} \frac{1}{k_BT}\mu _1^{\delta ,o}= & {} z_1\zeta _0V_0 +\ln L_1-\ln (R_1+\rho )\\&+\frac{2z_2-(1+\lambda )z_1}{z_2} (L_1- R_1-\rho )d+O(d^2) \end{aligned}$$

is the difference between the boundary electrochemical potentials for (BVo), and

$$\begin{aligned} w_0^{[o]}= & {} \frac{z_1(L_1- R_1-\rho )}{\ln L_1-\ln (R_1+\rho )} \; \text{ and } \\ w_1^{[o]}= & {} \frac{z_1(L_1- R_1-\rho )}{\ln L_1-\ln (R_1+\rho )}-\frac{z_1(L_1+R_1+\rho )}{2}. \end{aligned}$$

Recall that \(s=L_1/{R_1}\). One gets

$$\begin{aligned} \frac{1}{k_BT}\mu _1^{\delta ,i}&=z_1\zeta _0V_0 +\ln s +\frac{\rho }{sR_1} +\frac{2z_2-(1+\lambda )z_1}{z_2} (s-1) R_1d +O(\rho d, d^2),\\ w_0^{[i]}&= \frac{s-1}{\ln s}z_1R_1+\frac{s\ln s-(s-1)}{s(\ln s)^2} z_1\rho +O(\rho ^2),\\ w_1^{[i]}&=\frac{s-1}{\ln s}{z_1R_1}-\frac{s+1}{2}{z_1R_1}+\left( \frac{s\ln s-(s-1)}{s(\ln s)^2}-\frac{1}{2}\right) {z_1\rho }+O(\rho ^2);\\ \frac{1}{k_BT}\mu _1^{\delta ,o}&=z_1\zeta _0V_0 +\ln s -\frac{\rho }{R_1} +\frac{2z_2-(1+\lambda )z_1}{z_2} (s-1) R_1d +O(\rho d, d^2),\\ w_0^{[o]}&= \frac{s-1}{\ln s}{z_1R_1}+\frac{s-1-\ln s}{(\ln s)^2}z_1 \rho +O(\rho ^2),\\ w_1^{[o]}&=\frac{s-1}{\ln s}{z_1R_1}-\frac{s+1}{2}{z_1R_1}+\left( \frac{s-1-\ln s}{(\ln s)^2}-\frac{1}{2}\right) {z_1\rho }+O(\rho ^2). \end{aligned}$$

Thus,

$$\begin{aligned} \frac{\mu _1^{\delta ,i}}{\mu _1^{\delta ,o}}&=1+\frac{s+1}{s(z_1\zeta _0V_0 +\ln s)}\frac{\rho }{R_1}+O(\rho ^2, \rho d, d^2),\\ \frac{w_0^{[i]}}{w_0^{[o]}}&= 1+\frac{2s\ln s-s^2+1}{s(s-1)\ln s}\frac{\rho }{R_1}+O(\rho ^2),\\ w_1^{[i]}-w_1^{[o]}&=\frac{2s\ln s-s^2+1}{s(\ln s)^2}z_1\rho +O(\rho ^2). \end{aligned}$$

Therefore, the ratio between the fluxes \(J_1^{[i]}\) and \(J_1^{[o]}\) is given by

$$\begin{aligned} \frac{J_1^{[i]}}{J_1^{[o]}}&=\frac{z_1z_2+2(\lambda z_1-z_2)w_1^{[i]}d}{z_1z_2+2(\lambda z_1-z_2)w_1^{[o]} d}\cdot \frac{w_0^{[i]} }{w_0^{[o]}}\frac{\mu _1^{\delta ,i}}{\mu _1^{\delta ,o}}+O( d^2)\\&=\left( 1+\frac{2(\lambda z_1-z_2)}{z_1z_2}(w_1^{[i]}-w_1^{[o]})d\right) \frac{w_0^{[i]}}{w_0^{[o]}}\frac{\mu _1^{\delta ,i}}{\mu _1^{\delta ,o}} +O(d^2) \\&=1+\left( \frac{2s\ln s-s^2+1}{s(s-1)\ln s}+\frac{s+1}{t(z_1\zeta _0V_0 +\ln s)}\right) \frac{\rho }{R_1}+O(\rho ^2, \rho d,d^2). \end{aligned}$$

As a consequence,

$$\begin{aligned} \begin{aligned} S_0(\rho ) =1+\left( \frac{2s\ln s-s^2+1}{s(s-1)\ln s}+\frac{s+1}{s(z_1\zeta _0V_0 +\ln s)}\right) \frac{\rho }{R_1}\; \text{ and } \; S_1(0)=0. \end{aligned} \end{aligned}$$
(7.19)

Note that \(S_0(\rho )\) obtained here agrees with that in (7.16), as expected. The conclusion \(S_1(0)=0\) implies that there is not linear term in d in the approximation for \(J_1^{[i]}/{J_1^{[o]}}\).

1.2.3 Approximation of \(J_1^{[i]}/{J_1^{[o]}}\)

The superposition of (7.16) and (7.19) then gives

$$\begin{aligned} \begin{aligned} \frac{ J_1^{[i]}}{J_1^{[o]}}&=1+\left( \frac{2s\ln s-s^2+1}{s(s-1)\ln s}+\frac{s+1}{s(z_1\zeta _0V_0 +\ln s)}\right) \frac{\rho }{R_1}\\&\quad +g_1(V_0,s,\alpha ,\beta )\frac{\rho Q_0}{R_1^2}+O(\rho ^2, d^2, \rho d Q_0). \end{aligned} \end{aligned}$$
(7.20)

Finally, formula (5.14) in Theorem 5.1 can be obtained from the formula (7.2) by multiplying the three factors estimated in (7.4), (7.5) and (7.20).

1.3 Derivation of Formula (5.18)

We now derive formula (5.18) for the case where approximate electroneutrality boundary conditions are assumed; that is, we assume the boundary concentrations \(L_1\) and \(R_1\) for the main ion species and those \(L_2\) and \(R_2\) for the counterion species stay the same when tracer is added. Therefore, the electroneutrality boundary conditions are only approximate with the tracer concentration \(\rho \) being small. In this approximate electroneutrality boundary conditions, there will be boundary layers to compensate the imperfect electroneutrality.

More precisely, for (BVi) with tracer concentrations \(c_1^{[t,i]}(0)=L_1^{[t,i]}=\rho >0\) at \(x=0\) and \(c_1^{[t,i]}(1)=R_1^{[t,i]}=0\) at \(x=1\), one has

$$\begin{aligned} \text{(BVi): }\quad z_1(L_1+\rho )+z_2L_2=z_1\rho \approx 0\; \text{ and } \; z_1R_1+z_2R_2=0,\end{aligned}$$

and there will be one boundary layer at \(x=0\).

For (BVo) with tracer concentrations \(c_1^{[t,o]}(0)=L_1^{[t,o]}=0\) at \(x=0\) and \(c_1^{[t,o]}(1)=R_1^{[t,o]}=\rho >0\) at \(x=1\), one has

$$\begin{aligned} \text{(BVo): }\quad z_1L_1+z_2L_2= 0\; \text{ and } \; z_1(R_1+\rho )+z_2R_2=z_1\rho \approx 0, \end{aligned}$$

and there will be one boundary layer at \(x=1\).

For (BVi), let \((\phi _{\infty }^{[i]}, c_{1,\infty }^{[i]}, c_{1,\infty }^{[t,i]},c_{2,\infty }^{[i]})\) be the corresponding values of \((\phi ^{[i]}, c_1^{[i]},c_1^{[t,i]},c_2^{[i]})\) at the limiting point of the boundary layer at \(x=0\). Then, it follows from [56] (Lemma 3.2 and Proposition 3.3 in [56]) and (5.17) that

$$\begin{aligned} \begin{aligned} \phi _{\infty }^{[i]}&=V_0-\frac{1}{\zeta _0(z_1-z_2)}\ln \frac{-z_2L_2}{z_1(L_1+\rho )}=V_0+\frac{1}{\zeta _0(z_1-z_2)}\frac{\rho }{L_1}+O(\rho ^2),\\ c_{1,\infty }^{[i]}&=L_1\left( \frac{L_1}{L_1+\rho } \right) ^{\frac{z_1}{z_1-z_2}} =L_1-\frac{z_1}{z_1-z_2} \rho +O(\rho )^2, \\ c_{1,\infty }^{[t,i]}&=\rho \left( \frac{L_1}{L_1+\rho } \right) ^{\frac{z_1}{z_1-z_2}}=\rho -\frac{z_1}{z_1-z_2} \frac{\rho ^2}{L_1} +O(\rho )^3\\ c_{2,\infty }^{[i]}&=L_2 \left( \frac{L_1}{L_1+\rho } \right) ^{\frac{z_2}{z_1-z_2}}=L_2-\frac{z_1L_2}{(z_1-z_2)L_1}\rho +O(\rho )^2. \end{aligned} \end{aligned}$$
(7.21)

It is easy to check that \(z_1(c_{1,\infty }^{[i]}+ c_{1,\infty }^{[t,i]})+z_2c_{2,\infty }^{[i]}=0\), which is known to be true for the limiting point [56].

Similarly, for (BVo), let \((\phi _{\infty }^{[o]}, c_{1,\infty }^{[o]}, c_{1,\infty }^{[t,o]},c_{2,\infty }^{[o]})\) be the corresponding values of \((\phi ^{[o]}, c_1^{[o]},c_1^{[t,o]},c_2^{[o]})\) at the limiting point of the boundary layer at \(x=1\). Then,

$$\begin{aligned} \begin{aligned} \phi _{\infty }^{[o]}&= -\frac{1}{\zeta _0(z_1-z_2)}\ln \frac{-z_2R_2}{z_1(R_1+\rho )}=\frac{1}{\zeta _0(z_1-z_2)}\frac{\rho }{R_1}+O(\rho ^2),\\ c_{1,\infty }^{[o]}&=R_1\left( \frac{R_1}{R_1+\rho } \right) ^{\frac{z_1}{z_1-z_2}}=R_1-\frac{z_1}{z_1-z_2} \rho +O(\rho )^2, \\ c_{1,\infty }^{[t,o]}&=\rho \left( \frac{R_1}{R_1+\rho } \right) ^{\frac{z_1}{z_1-z_2}}=\rho -\frac{z_1}{z_1-z_2}\frac{\rho ^2}{R_1} +O(\rho )^3,\\ c_{2,\infty }^{[o]}&=R_2\left( \frac{R_1 }{R_1+\rho } \right) ^{\frac{z_2}{z_1-z_2}}=R_2-\frac{z_1R_2}{(z_1-z_2)R_1} \rho +O(\rho )^2. \end{aligned} \end{aligned}$$
(7.22)

One can derive the formula (5.18) following exactly the same procedure as in Sect. 7.2. Another approach is to make use of the derivation in Sect. 7.2 with necessary modifications. We have used both approaches and, not surprisingly, got the same approximate formula (5.18). The latter is simpler and is presented below.

More precisely, we need some modifications in the derivation in Sect. 7.2.

For (BVi), the boundary conditions should be replaced by

$$\begin{aligned} (\phi _{\infty }^{[i]}, c_{1,\infty }^{[i]}, c_{1,\infty }^{[t,i]}, c_{2,\infty }^{[i]})\; \text{ at } \; x=0;\quad (0, R_1, 0, R_2)\; \text{ at } \; x=1.\end{aligned}$$

For (BVo), the boundary conditions should be replaced by

$$\begin{aligned} (V_0, L_1, 0, L_2)\; \text{ at } \; x=0;\quad (\phi _{\infty }^{[o]}, c_{1,\infty }^{[o]}, c_{1,\infty }^{[t,o]}, c_{2,\infty }^{[o]})\; \text{ at } \; x=1.\end{aligned}$$

With these modifications, it is not hard to get, from (5.6) that, for the boundary condition associated to (BVi) in (5.17),

$$\begin{aligned} p_1^{[i]}(0^+;z_1,d_1)&=z_1\zeta _0\phi _{\infty }^{[i]}+ d\Big (2(c_{1,\infty }^{[i]}+c_{1,\infty }^{[t,i]}) +(1+\lambda ) c_{2,\infty }^{[i]}\Big ) +O(d^2)\\&=z_1\zeta _0V_0+\frac{z_1}{z_1-z_2}\frac{\rho }{L_1}+ d\Big (2L_1 +(1+\lambda ) L_2\Big ) +O(d^2,d\rho ),\\ p_1^{[i]}(1;z_1,d_1)&= d\Big (2R_1+(1+\lambda ) R_2\Big ) +O(d^2); \end{aligned}$$

and, for the boundary condition associated to (BVo) in (5.17),

$$\begin{aligned} p_1^{[o]}(0;z_1,d_1)&=z_1\zeta _0V_0+ d(2L_1+(1+\lambda ) L_2) +O(d^2),\\ p_1^{[o]}(1^{-};z_1,d_1)&=z_1\zeta _0\phi _{\infty }^{[o]}+ d(2(c_{1,\infty }^{[o]}+c_{1,\infty }^{[t,o]})+(1+\lambda ) c_{2,\infty }^{[o]}) +O(d^2)\\&= \frac{z_1}{z_1-z_2}\frac{\rho }{R_1}+ d(2R_1+(1+\lambda ) R_2) +O(d^2,d\rho ). \end{aligned}$$

Thus, it follows from the definitions of the factors in (7.2) that

$$\begin{aligned} \begin{aligned} (I)&=\frac{c_{1,\infty }^{[t,i]}}{c_{1,\infty }^{[t,o]}}\left( e^{p_1^{[i]}(0;z_1,d_1)-p_1^{[o]}(1^-;z_1,d_1)}\right) =\left( 1+\frac{z_1(s-1)}{(z_1-z_2)s}\frac{\rho }{R_1}\right) e^{z_1\zeta _0V_0} \\&\quad \times \Big (1+\frac{z_1(1-s)}{(z_1-z_2)s}\frac{\rho }{R_1}+\big (2 (L_1-R_1) +(1+\lambda ) (L_2- R_2) \big )d\Big )\\&\quad +O(d^2,d\rho , \rho ^2)\\&=e^{z_1\zeta _0V_0} \Big (1+\big (2 (L_1-R_1) +(1+\lambda ) (L_2- R_2) \big )d\Big )+O(d^2,d\rho , \rho ^2), \end{aligned} \end{aligned}$$
(7.23)

and

$$\begin{aligned} \begin{aligned} (II)&=\frac{L_1e^{p_1^{[o]}(0;z_1,d_1)}-(c_{1,\infty }^{[o]}+c_{1,\infty }^{[t,o]})e^{p_1^{[o]}(1^-;z_1,d_1)}}{(c_{1,\infty }^{[i]}+c_{1,\infty }^{[t,i]})e^{p_1^{[i]}(0^+;z_1,d_1)}-R_1e^{p_1^{[i]}(1;z_1,d_1)}}+O(\rho ^2, \rho d, d^2)\\&=1 -\frac{ e^{z_1\zeta _0V_0}+1}{se^{z_1\zeta _0V_0}-1}\frac{\rho }{R_1} +O(\rho ^2, \rho d, d^2). \end{aligned} \end{aligned}$$
(7.24)

Also, \(F_0\) in (7.15) should be replaced by

$$\begin{aligned} F_0&=\frac{c_{1,\infty }^{[i]}+c_{1,\infty }^{[t,i]}-R_1}{L_1-(c_{1,\infty }^{[o]}+c_{1,\infty }^{[t,o]})} \frac{\ln L_1-\ln (c_{1,\infty }^{[o]}+c_{1,\infty }^{[t,o]})}{\ln (c_{1,\infty }^{[i]}+c_{1,\infty }^{[t,i]})-\ln R_1} \\&\quad \times \frac{z_1\zeta _0 \phi _{\infty }^{[i]}+\ln (c_{1,\infty }^{[i]}+c_{1,\infty }^{[t,i]})-\ln R_1}{z_1\zeta _0 \phi _{\infty }^{[o]}+\ln L_1-\ln (c_{1,\infty }^{[o]}+c_{1,\infty }^{[t,o]})}. \end{aligned}$$

A straightforward calculations gives

$$\begin{aligned} F_0 =1+ \left( \frac{-z_2}{z_1-z_2}f_1(s)+\frac{s+1}{s(z_1\zeta _0V_0+\ln s)}\right) \frac{\rho }{R_1}+O(\rho ^2), \end{aligned}$$

where \(f_1(s)\) is defined in (5.8).

The most complicate term is \(F_0/{F_1}\) in (7.15). By a close examination of the terms involved in the expression, the following modifications are needed. One is to replace \(\rho \) with \(-z_2\rho /{(z_1-z_2)}\). The other is, in the last expression of \(F_0/{F_1}\) in (7.15), to replace \(V_0\) by \(\phi _{\infty }^{[i]}\) in the first term and to replace \(V_0\) by \(V_0-\phi _{\infty }^{[o]}\) in the second term. The combined effect of the latter is the extra term

$$\begin{aligned} (E)=&\frac{A^{[i]}(1-B^{[i]})}{(z_1-z_2)\left( L_1-\frac{z_2}{z_1-z_2}\rho -R_1\right) \left( \ln (L_1-\frac{z_2}{z_1-z_2}\rho )-\ln R_1\right) }\frac{z_2}{z_1-z_2}\frac{\rho }{sR_1}\\&\quad +\frac{A^{[o]}(1-B^{[o]})}{(z_1-z_2)\left( L_1-R_1+\frac{z_2}{z_1-z_2}\rho \right) \left( \ln L_1-\ln (R_1-\frac{z_2}{z_1-z_2}\rho )\right) }\frac{z_2}{z_1-z_2} \frac{\rho }{R_1}\\&=\frac{z_2}{(z_1-z_2)^2}\frac{A_0^{[i]}+A_0^{[0]}s}{s(s-1)\ln s}\frac{\rho }{R_1^2}-\frac{z_2}{(z_1-z_2)^2}\frac{A_0^{[i]}B_0^{[i]}+A_0^{[0]}B_0^{[0]}s}{s(s-1)\ln s}\frac{\rho }{R_1^2}. \end{aligned}$$

Using the expressions of \(A_0^{[i]}, A_0^{[0]}, A_0^{[i]}B_0^{[i]}\) and \(A_0^{[0]}B_0^{[0]}\) in (7.17), one gets

$$\begin{aligned} (E)=g_2(s,\alpha ,\beta )\frac{-z_2}{z_1-z_2}\frac{\rho }{R_1^2}, \end{aligned}$$

where \(g_2(s,\alpha ,\beta )\) is given in (5.9). Therefore, we have

$$\begin{aligned} \frac{F_1}{F_0}&=\frac{-z_2}{z_1-z_2}g_1(V_0,s,\alpha ,\beta )\frac{\rho }{R_1^2}+(E)\\&=\frac{-z_2}{z_1-z_2}\Big (g_1(V_0,s,\alpha ,\beta )+g_2(s,\alpha ,\beta )\Big )\frac{\rho }{R_1^2}+O(\rho ^2), \end{aligned}$$

and hence,

$$\begin{aligned} \begin{aligned} \frac{ J_1^{[i]}}{J_1^{[o]}}&=F_0\left( 1+\frac{F_1}{F_0}Q_0\right) =1\\&\quad +\left( \frac{-z_2}{z_1-z_2}f_1(s)+\frac{s+1}{s(z_1\zeta _0V_0 +\ln s)}\right) \frac{\rho }{R_1}\\&\quad +\Big (g_1(V_0,s,\alpha ,\beta )+g_2(s,\alpha ,\beta )\Big )\frac{-z_2}{z_1-z_2}\frac{\rho Q_0}{R_1^2}+O(\rho ^2, d^2, \rho d Q_0). \end{aligned} \end{aligned}$$
(7.25)

The formula (5.18) in Theorem 5.2 then follows from (7.2), and the estimates (7.23), (7.24) and (7.25).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ji, S., Eisenberg, B. & Liu, W. Flux Ratios and Channel Structures. J Dyn Diff Equat 31, 1141–1183 (2019). https://doi.org/10.1007/s10884-017-9607-1

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10884-017-9607-1

Keywords

Mathematics Subject Classification

Navigation