Skip to main content
Log in

Resolving molecular contributions of ion channel noise to interspike interval variability through stochastic shielding

  • Original Article
  • Published:
Biological Cybernetics Aims and scope Submit manuscript

A Correction to this article was published on 24 June 2021

This article has been updated

Abstract

Molecular fluctuations can lead to macroscopically observable effects. The random gating of ion channels in the membrane of a nerve cell provides an important example. The contributions of independent noise sources to the variability of action potential timing have not previously been studied at the level of molecular transitions within a conductance-based model ion-state graph. Here we study a stochastic Langevin model for the Hodgkin–Huxley (HH) system based on a detailed representation of the underlying channel state Markov process, the “\(14\times 28\)D model” introduced in (Pu and Thomas in Neural Computation 32(10):1775–1835, 2020). We show how to resolve the individual contributions that each transition in the ion channel graph makes to the variance of the interspike interval (ISI). We extend the mean return time (MRT) phase reduction developed in (Cao et al. in SIAM J Appl Math 80(1):422–447, 2020) to the second moment of the return time from an MRT isochron to itself. Because fixed-voltage spike detection triggers do not correspond to MRT isochrons, the inter-phase interval (IPI) variance only approximates the ISI variance. We find the IPI variance and ISI variance agree to within a few percent when both can be computed. Moreover, we prove rigorously, and show numerically, that our expression for the IPI variance is accurate in the small noise (large system size) regime; our theory is exact in the limit of small noise. By selectively including the noise associated with only those few transitions responsible for most of the ISI variance, our analysis extends the stochastic shielding (SS) paradigm (Schmandt and Galán in Phys Rev Lett 109(11):118101, 2012) from the stationary voltage clamp case to the current clamp case. We show numerically that the SS approximation has a high degree of accuracy even for larger, physiologically relevant noise levels. Finally, we demonstrate that the ISI variance is not an unambiguously defined quantity, but depends on the choice of voltage level set as the spike detection threshold. We find a small but significant increase in ISI variance, the higher the spike detection voltage, both for simulated stochastic HH data and for voltage traces recorded in in vitro experiments. In contrast, the IPI variance is invariant with respect to the choice of isochron used as a trigger for counting “spikes.”

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
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14

Similar content being viewed by others

Change history

Notes

  1. Equivalently, “iso-phase-intervals”: the time taken to complete one full oscillation, from a given isochron back to the same isochron.

  2. For the \(14\times 28\)D Langevin Hodgkin–Huxley model, \(\varOmega \) may be thought of as the space of continuous vector functions on \([0,\infty )\) with 28 components – one for each independent noise source.

  3. In this paper, we focus on a Langevin equation, rather than models with discrete channel noise. Therefore, our trajectories are diffusions that have continuous sample paths (with probability one). Therefore, the FPT \(\tau ({\mathbf {x}},{\mathcal {S}})\) is well defined. For discrete channel noise models, a slightly modified definition would be required.

  4. (Gardiner 2004), §6.6, First Exit Time from a Region (Homogeneous Processes), equation 6.6.8.

  5. (Gardiner 2004), Chapter 6.6, First Exit Time from a Region (Homogeneous Processes), equation 6.6.11.

  6. Throughout this section, we use the term “threshold” in the data analysis sense of a Schmitt trigger (Schmitt 1938), rather than the physiological sense of a spike generation mechanism.

  7. Note that the factor b, representing the shift in mean frequency due to noise, does not appear in the expression for the variance of the phase after one period. Similarly, the term \({\overline{T}}_1\) appearing in our assumption A3, Eq. (38), represents the first-order shift in mean period of the stochastic oscillator upon introducing nonzero noise. Observe that \({\overline{T}}_1\) does not appear in our expression for the inter-phase interval variance, consistent with the results in Giacomin et al. (2018).

References

  • Aminzare Z, Holmes P, Srivastava V (2019) On phase reduction and time period of noisy oscillators. In: 2019 IEEE 58th conference on decision and control (CDC), pp 4717–4722

  • Anderson DF, Ermentrout B, Thomas PJ (2015) Stochastic representations of ion channel kinetics and exact stochastic simulation of neuronal dynamics. J Comput Neurosci 38(1):67–82. https://doi.org/10.1007/s10827-014-0528-2

    Article  PubMed  Google Scholar 

  • Azouz R, Gray CM (2000) Dynamic spike threshold reveals a mechanism for synaptic coincidence detection in cortical neurons in vivo. Proc Natl Acad Sci 97(14):8110–8115

    CAS  PubMed  PubMed Central  Google Scholar 

  • Bezanilla F (1987) Single sodium channels from the squid giant axon. Biophys J 52(6):1087–1090

    CAS  PubMed  PubMed Central  Google Scholar 

  • Bressloff PC, Newby JM (2014) Path integrals and large deviations in stochastic hybrid systems. Phys Rev E 89(4):042701

    Google Scholar 

  • Brown E, Moehlis J, Holmes P (2004) On the phase reduction and response dynamics of neural oscillator populations. Neural Comput 16(4):673–715. https://doi.org/10.1162/089976604322860668

    Article  PubMed  Google Scholar 

  • Brunel N, Latham PE (2003) Firing rate of the noisy quadratic integrate-and-fire neuron. Neural Comput 15(10):2281–2306

    PubMed  Google Scholar 

  • Buckwar E, Riedler MG (2011) An exact stochastic hybrid model of excitable membranes including spatio-temporal evolution. J Math Biol 63(6):1051–1093

    PubMed  Google Scholar 

  • Butera RJ Jr, Rinzel J, Smith JC (1999) Models of respiratory rhythm generation in the pre-Bötzinger complex. I. Bursting pacemaker neurons. J Neurophysiol 82(1):382–397

    PubMed  Google Scholar 

  • Cao A (2017) Dimension reduction for stochastic oscillators: investigating competing generalizations of phase and isochrons. Master’s thesis, Case Western Reserve University, Cleveland, Ohio

  • Cao A, Lindner B, Thomas PJ (2020) A partial differential equation for the mean-return-time phase of planar stochastic oscillators. SIAM J Appl Math 80(1):422–447

    Google Scholar 

  • Chacron MJ, Lindner B, Longtin A (2004) Noise shaping by interval correlations increases information transfer. Phys Rev Lett 92(8):080601

    PubMed  Google Scholar 

  • Dorval AD, White JA (2005) Channel noise is essential for perithreshold oscillations in entorhinal stellate neurons. J Neurosci 25(43):10025–10028

    CAS  PubMed  PubMed Central  Google Scholar 

  • Ermentrout GB, Beverlin B 2nd, Troyer T, Netoff TI (2011) The variance of phase-resetting curves. J Comput Neurosci. https://doi.org/10.1007/s10827-010-0305-9

  • Ermentrout GB, Terman DH (2010) Foundations of mathematical neuroscience. Springer, Berlin

    Google Scholar 

  • Faisal AA, White JA, Laughlin SB (2005) Ion-channel noise places limits on the miniaturization of the brain’s wiring. Curr Biol 15(12):1143–1149

    CAS  PubMed  Google Scholar 

  • Fenwick EM, Marty A, Neher E (1982) Sodium and calcium channels in bovine chromaffin cells. J Physiol 331(1):599–635

    CAS  PubMed  PubMed Central  Google Scholar 

  • Fisch K, Schwalger T, Lindner B, Herz AV, Benda J (2012) Channel noise from both slow adaptation currents and fast currents is required to explain spike-response variability in a sensory neuron. J Neurosci 32(48):17332–17344

    CAS  PubMed  PubMed Central  Google Scholar 

  • Forrest MD (2015) Simulation of alcohol action upon a detailed Purkinje neuron model and a simpler surrogate model that runs\(>\)400 times faster. BMC Neurosci 16(1):27

    PubMed  PubMed Central  Google Scholar 

  • Fox Lu (1994) Emergent collective behavior in large numbers of globally coupled independently stochastic ion channels. Phys Rev E Stat Phys Plasmas Fluids Relat Interdiscip Topics 49(4):3421–3431

    CAS  PubMed  Google Scholar 

  • Gardiner CW (2004) Handbook of stochastic methods for physics, chemistry, and the natural sciences, 2nd edn. Springer, Berlin

    Google Scholar 

  • Gerstein GL (1960) Analysis of firing patterns in single neurons. Science 131(3416):1811–1812

    CAS  PubMed  Google Scholar 

  • Gerstner W, Kempter R, Van Hemmen JL, Wagner H (1996) A neuronal learning rule for sub-millisecond temporal coding. Nature 383(6595):76–78

    CAS  PubMed  Google Scholar 

  • Giacomin G, Poquet C, Shapira A (2018) Small noise and long time phase diffusion in stochastic limit cycle oscillators. J Differ Equ 264(2):1019–1049

    Google Scholar 

  • Goldwyn JH, Shea-Brown E (2011) The what and where of adding channel noise to the Hodgkin–Huxley equations. PLoS Comput Biol 7(11):e1002247. https://doi.org/10.1371/journal.pcbi.1002247

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Goldwyn JH, Shea-Brown E, Rubinstein JT (2010) Encoding and decoding amplitude-modulated cochlear implant stimuli: a point process analysis. J Comput Neurosci 28(3):405–424

    PubMed  PubMed Central  Google Scholar 

  • Guckenheimer J (1975) Isochrons and phaseless sets. J Math Biol 1:259–273. https://doi.org/10.1007/BF01273747

    Article  CAS  PubMed  Google Scholar 

  • Guckenheimer J, Holmes P (1990) Nonlinear oscillations, dynamical systems, and bifurcations of vector fields. Applied Mathematical Sciences, vol. 42, third edn. Springer

  • Gutkin BS, Ermentrout GB (1998) Dynamics of membrane excitability determine interspike interval variability: a link between spike generation mechanisms and cortical spike train statistics. Neural Comput 10(5):1047–1065

    CAS  PubMed  Google Scholar 

  • Kuramoto Y (1984) Chemical Oscillations, Waves, and Turbulence, Springer-Verlag, Berlin

  • Kurtz TG (1978) Strong approximation theorems for density dependent Markov chains. Stoch Process Appl 6(3):223–240

    Google Scholar 

  • Lasota A, Mackey MC (1994) Chaos, fractals, and noise: stochastic aspects of dynamics, applied mathematical sciences, vol 97. Springer, New York

    Google Scholar 

  • Linaro D, Storace M, Giugliano M (2011) Accurate and fast simulation of channel noise in conductance-based model neurons by diffusion approximation. PLoS Comput Biol 7(3):e1001102

    CAS  PubMed  PubMed Central  Google Scholar 

  • Lindner B (2004) Interspike interval statistics of neurons driven by colored noise. Phys Rev E 69(2):022901

    Google Scholar 

  • Lindner B, Chacron MJ, Longtin A (2005) Integrate-and-fire neurons with threshold noise: a tractable model of how interspike interval correlations affect neuronal signal transmission. Phys Rev E 72(2):021911

    Google Scholar 

  • Llinás R, Sugimori M (1980) Electrophysiological properties of in vitro Purkinje cell dendrites in mammalian cerebellar slices. J Physiol 305(1):197–213

    PubMed  PubMed Central  Google Scholar 

  • Llinás R, Sugimori M (1980) Electrophysiological properties of in vitro Purkinje cell somata in mammalian cerebellar slices. J Physiol 305(1):171–195

    PubMed  PubMed Central  Google Scholar 

  • Mainen ZF, Sejnowski TJ (1995) Reliability of spike timing in neocortical neurons. Science 268(5216):1503–1506

    CAS  PubMed  Google Scholar 

  • Makielski JC, Sheets MF, Hanck DA, January CT, Fozzard HA (1987) Sodium current in voltage clamped internally perfused canine cardiac Purkinje cells. Biophys J 52(1):1–11

    CAS  PubMed  PubMed Central  Google Scholar 

  • McKane AJ, Newman TJ (2005) Predator-prey cycles from resonant amplification of demographic stochasticity. Phys Rev Lett 94(21):218102

    CAS  PubMed  Google Scholar 

  • Moiseff A, Konishi M (1981) The owl’s interaural pathway is not involved in sound localization. J Comput Physiol 144(3):299–304

    Google Scholar 

  • Mukhametov L, Rizzolatti G, Tradardi V (1970) Spontaneous activity of neurones of nucleus reticularis thalami in freely moving cats. J Physiol 210(3):651–667

    CAS  PubMed  PubMed Central  Google Scholar 

  • Neher E, Sakmann B (1976) Single-channel currents recorded from membrane of denervated frog muscle fibres. Nature 260(5554):799

    CAS  PubMed  Google Scholar 

  • Netoff T, Schwemmer MA, Lewis TJ (2012) Experimentally estimating phase response curves of neurons: theoretical and practical issues. In: Phase Response curves in neuroscience, pp 95–129. Springer

  • Øksendal B (2007) Stochastic differential equations: an introduction with applications, 6th edn. Springer, Berlin

    Google Scholar 

  • Orio P, Soudry D (2012) Simple, fast and accurate implementation of the diffusion approximation algorithm for stochastic ion channels with multiple states. PLoS ONE 7(5):e36670

    CAS  PubMed  PubMed Central  Google Scholar 

  • Ovsepian SV, Friel DD (2008) The leaner P/Q-type calcium channel mutation renders cerebellar Purkinje neurons hyper-excitable and eliminates Ca\(^{2+}\)-Na\(^+\) spike bursts. Eur J Neurosci 27(1):93–103

    PubMed  Google Scholar 

  • Pakdaman K, Thieullen M, Wainrib G (2010) Fluid limit theorems for stochastic hybrid systems with application to neuron models. Adv Appl Probab 42(3):761–794

    Google Scholar 

  • Pezo D, Soudry D, Orio P (2014) Diffusion approximation-based simulation of stochastic ion channels: which method to use? Front Comput Neurosci 8:139

    PubMed  PubMed Central  Google Scholar 

  • Pietrobon D (2010) Ca v 2.1 channelopathies. Pflügers Archiv-Eur J Physiol 460(2):375–393

    CAS  Google Scholar 

  • Pikovsky A (2015) Comment on “asymptotic phase for stochastic oscillators.” Phys. Rev. Lett. 115:069401. https://doi.org/10.1103/PhysRevLett.115.069401

  • Pu S (2020) Noise decomposition for stochastic Hodgkin-Huxley models. Ph.D. thesis, Case Western Reserve University, Cleveland, Ohio

  • Pu S, Thomas PJ (2020) Fast and accurate Langevin simulations of stochastic Hodgkin–Huxley dynamics. Neural Comput 32(10):1775–1835

    PubMed  Google Scholar 

  • Rajakulendran S, Kaski D, Hanna MG (2012) Neuronal P/Q-type calcium channel dysfunction in inherited disorders of the CNS. Nat Rev Neurol 8(2):86

    CAS  PubMed  Google Scholar 

  • Risken H (1996) The Fokker–Planck equation: methods of solution and applications, 2nd edn. Springer Series in Synergistics. Springer, Berlin

  • Rowat P (2007) Interspike interval statistics in the stochastic Hodgkin-Huxley model: coexistence of gamma frequency bursts and highly irregular firing. Neural Comput 19(5):1215–1250

    PubMed  Google Scholar 

  • Rowat PF, Greenwood PE (2011) Identification and continuity of the distributions of burst-length and interspike intervals in the stochastic Morris-Lecar neuron. Neural Comput 23(12):3094–3124

    PubMed  Google Scholar 

  • Saks S (1937) Theory of the integral. http://eudml.org/doc/219302

  • Schleimer JH (2013) Spike statistics and coding properties of phase models. Ph.D. thesis, Humboldt-Universität zu Berlin, Mathematisch-Naturwissenschaftliche Fakultät I

  • Schmandt NT, Galán RF (2012) Stochastic-shielding approximation of Markov chains and its application to efficiently simulate random ion-channel gating. Phys Rev Lett 109(11):118101

    PubMed  Google Scholar 

  • Schmid G, Goychuk I, Hänggi P (2001) Stochastic resonance as a collective property of ion channel assemblies. Europhys Lett 56(1):22

    CAS  Google Scholar 

  • Schmidt DR, Galán RF, Thomas PJ (2018) Stochastic shielding and edge importance for Markov chains with timescale separation. PLoS Comput Biol 14(6):e1006206

    PubMed  PubMed Central  Google Scholar 

  • Schmidt DR, Thomas PJ (2014) Measuring edge importance: a quantitative analysis of the stochastic shielding approximation for random processes on graphs. J Math Neurosci 4(1):6

    PubMed  PubMed Central  Google Scholar 

  • Schmitt OH (1938) A thermionic trigger. J Sci Instrum 15(1):24

    Google Scholar 

  • Schneidman E, Freedman B, Segev I (1998) Ion channel stochasticity may be critical in determining the reliability and precision of spike timing. Neural Comput 10(7):1679–1703

    CAS  PubMed  Google Scholar 

  • Schwabedal J, Pikovsky A (2010) Effective phase dynamics of noise-induced oscillations in excitable systems. Phys Rev E: Stat Nonlinear Soft Matter Phys 81(4 Pt 2):046218

    Google Scholar 

  • Schwabedal J, Pikovsky A (2013) Phase description of stochastic oscillations. Phys Rev Lett 110:4102

    Google Scholar 

  • Schwalger T, Fisch K, Benda J, Lindner B (2010) How noisy adaptation of neurons shapes interspike interval histograms and correlations. PLoS Comput Biol 6(12):e1001026

    PubMed  PubMed Central  Google Scholar 

  • Schwalger T, Tiana-Alsina J, Torrent M, Garcia-Ojalvo J, Lindner B (2012) Interspike-interval correlations induced by two-state switching in an excitable system. Europhys Lett 99(1):10004

    Google Scholar 

  • Schwemmer MA, Lewis TJ (2012) The theory of weakly coupled oscillators. In: Phase response curves in neuroscience, pp 3–31. Springer

  • Shadlen MN, Newsome WT (1998) The variable discharge of cortical neurons: implications for connectivity, computation, and information coding. J Neurosci 18(10):3870–3896

    CAS  PubMed  PubMed Central  Google Scholar 

  • Shiau L, Schwalger T, Lindner B (2015) Interspike interval correlation in a stochastic exponential integrate-and-fire model with subthreshold and spike-triggered adaptation. J Comput Neurosci 38(3):589–600

    PubMed  Google Scholar 

  • Shingai R, Quandt FN (1986) Single inward rectifier channels in horizontal cells. Brain Res 369(1–2):65–74

    CAS  PubMed  Google Scholar 

  • Sigworth F (1977) Sodium channels in nerve apparently have two conductance states. Nature 270(5634):265–267

    CAS  PubMed  Google Scholar 

  • Stein RB (1965) A theoretical analysis of neuronal variability. Biophys J 5(2):173–194

    CAS  PubMed  PubMed Central  Google Scholar 

  • T O’Connor D, Mahata SK, Mahata M, Jiang Q, Hook VY, Taupenot L (2007) Primary culture of bovine chromaffin cells. Nat Protoc 2(5):1248–1253

    Google Scholar 

  • Thomas PJ, Lindner B (2014) Asymptotic phase for stochastic oscillators. Phys Rev Lett 113(25):254101

    PubMed  Google Scholar 

  • Thomas PJ, Lindner B (2015) Thomas and lindner reply. Phys. Rev. Lett. 115.6 (2015): 069402. (A reply to “Comment on Asymptotic Phase for Stochastic Oscillators” by A. Pikovsky, same journal.)

  • Thomas PJ, Lindner B (2019) Phase descriptions of a multidimensional Ornstein-Uhlenbeck process. Phys Rev E 99(6):062221

    CAS  PubMed  Google Scholar 

  • Vilela RD, Lindner B (2009) Are the input parameters of white noise driven integrate and fire neurons uniquely determined by rate and CV? J Theor Biol 257(1):90–99

    PubMed  Google Scholar 

  • Walter JT, Alvina K, Womack MD, Chevez C, Khodakhah K (2006) Decreases in the precision of Purkinje cell pacemaking cause cerebellar dysfunction and ataxia. Nat Neurosci 9(3):389–397

    CAS  PubMed  Google Scholar 

  • White JA, Rubinstein JT, Kay AR (2000) Channel noise in neurons. Trends Neurosci 23(3):131–137

    CAS  PubMed  Google Scholar 

  • Wilson D, Ermentrout B (2018) Greater accuracy and broadened applicability of phase reduction using isostable coordinates. J Math Biol 76(1–2):37–66

    PubMed  Google Scholar 

  • Wilson D, Ermentrout B (2018) An operational definition of phase characterizes the transient response of perturbed limit cycle oscillators. SIAM J Appl Dyn Syst 17(4):2516–2543

    Google Scholar 

  • Wilson D, Ermentrout B (2019) Augmented phase reduction of (not so) weakly perturbed coupled oscillators. SIAM Rev 61(2):277–315

    Google Scholar 

  • Winfree A (2000) The geometry of biological time, 2nd edn. Springer, New York

    Google Scholar 

  • Winfree AT (1974) Patterns of phase compromise in biological cycles. J Math Biol 1:73–95

    Google Scholar 

Download references

Acknowledgements

The authors thank Dr. David Friel (Case Western Reserve University, School of Medicine) for introducing us to the problem of channel noise and spontaneous spike time variability in the leaner mouse, for providing access to voltage recordings obtained in his laboratory and for many illuminating discussions on mechanisms at work in Purkinje cells. The authors thank Dr. Daniela Calvetti and Dr. Erkki Somersalo (Case Western Reserve University, Department of Mathematics, Applied Mathematics and Statistics) for helpful advice and discussion. This work was made possible in part by grants from the National Science Foundation (DMS-2052109, DEB-1654989). PT thanks the Oberlin College Department of Mathematics for research support. This research has been supported in part by the Mathematical Biosciences Institute and the National Science Foundation under grant DMS-1440386. Large-scale simulations made use of the High Performance Computing Resource in the Core Facility for Advanced Research Computing at Case Western Reserve University.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Shusen Pu.

Additional information

Communicated by André Longtin.

Publisher's Note

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

The original article has been updated: Due to textual changes.

Appendices

Model parameters, common symbols and notations

Table 2 Parameters used for simulations in this paper

Subunit kinetics for Hodgkin and Huxley parameters are given by

$$\begin{aligned}&\alpha _m(V)= \frac{0.1*(25-V)}{ \exp (2.5-0.1V)-1} \end{aligned}$$
(107)
$$\begin{aligned}&\beta _m(V)= 4*\exp (-V/18) \end{aligned}$$
(108)
$$\begin{aligned}&\alpha _h(V)= 0.07*\exp (-V/20) \end{aligned}$$
(109)
$$\begin{aligned}&\beta _h(V)= \frac{1}{ \exp (3-0.1V)+1} \end{aligned}$$
(110)
$$\begin{aligned}&\alpha _n(V)= \frac{0.01* (10-V)}{\exp (1-0.1V)-1} \end{aligned}$$
(111)
$$\begin{aligned}&\beta _n(V)= 0.125\exp (-V/80) \end{aligned}$$
(112)
$$\begin{aligned}&A_K(V) =\begin{bmatrix} D_1(1)&{} \beta _n(V) &{} 0 &{} 0 &{} 0\\ 4\alpha _n(V)&{} D_1(2)&{} 2\beta _n(V) &{} 0&{} 0\\ 0&{} 3\alpha _n(V)&{} D_1(3)&{} 3\beta _n(V)&{} 0\\ 0&{} 0&{} 2\alpha _n(V)&{} D_1(4)&{} 4\beta _n(V)\\ 0&{} 0&{} 0&{} \alpha _n(V)&{} D_1(5) \end{bmatrix},\nonumber \\&A_\text {Na}\nonumber \\&\quad =\begin{bmatrix} D_2(1) &{} \beta _m&{}0 &{}0 &{}\beta _h&{}0&{}0&{}0\\ 3\alpha _m&{}D_2(2)&{}2\beta _m&{}0&{}0&{}\beta _h&{}0&{}0\\ 0&{}2\alpha _m&{}D_2(3) &{}3\beta _m&{}0&{}0&{}\beta _h&{}0 \\ 0&{}0&{}\alpha _m&{}D_2(4)&{}0&{}0&{}0&{}\beta _h \\ \alpha _h&{}0&{}0&{}0&{}D_2(5)&{}\beta _m&{}0&{}0\\ 0&{}\alpha _h&{}0&{}0&{}3\alpha _m&{}D_2(6)&{}2\beta _m&{}0\\ 0&{}0&{}\alpha _h&{}0&{}0&{}2\alpha _m&{}D_2(7)&{}3\beta _m\\ 0&{}0&{}0&{}\alpha _h&{}0&{}0&{}\alpha _m&{}D_2(8)\\ \end{bmatrix},\nonumber \\ \end{aligned}$$
(113)

where the diagonal elements

$$\begin{aligned} D_k(i)=-\sum _{j\ne i}A_\text {ion}(j,i),\quad k\in \{1,2\} \quad \text {ion}\in \{\text {Na},\text {K}\}. \end{aligned}$$
Table 3 Table of common symbols and notations

Diffusion matrix of the 14D model

Define the state vector for \(\hbox {Na}^+\) and \(\hbox {K}^+\) channels as

$$\begin{aligned}&{\mathbf {M}}=[m_{00},m_{10},m_{20},m_{30},m_{01},m_{11},m_{21},m_{31}]^\intercal , \end{aligned}$$

and \({\mathbf {N}}=[n_0,n_1,n_2,n_3,n_4]^\intercal \), respectively.

The matrices \(S_\text {K}\) and \(S_\text {Na}\) are given by

and

where \(S^{(i:j)}_\text {Na}\) is the \(\hbox {i}^{th}\)-\(\hbox {j}^{th}\) column of \(S_\text {Na}\).

Note that each of the 8 columns of \(S_\text {K}\) corresponds to the flux vector along a single directed edge in the \(\hbox {K}^+\) channel transition graph. Similarly, each of the 20 columns of \(S_\text {Na}\) corresponds to the flux vector along a directed edge in the \(\hbox {Na}^+\) graph (cf. Fig. 2). Factors \(M_\text {ref}=6000\) and \(N_\text {ref}=1800\) represent the reference number of \(\hbox {K}^+\) and \(\hbox {Na}^+\) channels from Goldwyn and Shea-Brown’s model (Goldwyn and Shea-Brown 2011).

Proof of Lemma 1

For the reader’s convenience, we restate

Lemma  1For a conductance-based model of the form (3), and for any fixed applied current \(I_\text {app}\), there exist upper and lower bounds \(v_\text {max}\) and \(v_\text {min}\) such that trajectories with initial voltage condition \(v\in [v_\mathrm{min},v_\mathrm{max}]\) remain within this interval for all times \(t>0\), with probability 1, regardless of the initial channel state, provided the gating variables satisfy \(0\le M_{ij}\le 1\) and \(0\le N_i\le 1\).

Proof

Let \(V_1={\text {min}}_{\mathrm{ion}}\{V_\mathrm{ion}\}\wedge V_\mathrm{leak}\), and \(V_2={\text {max}}_{\mathrm{ion}}\{V_\mathrm{ion}\}\vee V_\mathrm{leak}\), where \(\text {ion}\in \{\text {Na}^+, \text {K}^+\}\). Note that by assumption, for both the \(\hbox {Na}^+\)  and \(\hbox {K}^+\) channel, \(0\le {M_{31}} \le 1\), \(0\le {N_4} \le 1\). Moreover, \(g_i>0,\ g_\mathrm{leak}>0\); therefore, when \(V\le V_1\)

$$\begin{aligned} \frac{\hbox {d}V}{\hbox {d}t}&=\frac{1}{C}\left\{ I_\text {app}(t)-{\bar{g}}_\text {Na}{M_{31}}\left( V-V_\text {Na}\right) \nonumber \right. \\&\left. -{\bar{g}}_\text {K}{N_4}\left( V-V_\text {K}\right) -g_\text {leak}(V-V_\text {leak})\right\} \end{aligned}$$
(114)
$$\begin{aligned}&\ge \frac{1}{C}\left\{ I_\text {app}(t)-{\bar{g}}_\text {Na}{M_{31}}\left( V-V_1\right) \nonumber \right. \\&\left. -{\bar{g}}_\text {K}{N_4}\left( V-V_1\right) -g_\text {leak}(V-V_1)\right\} \end{aligned}$$
(115)
$$\begin{aligned}&\ge \frac{1}{C}\left\{ I_\text {app}(t)-{{\bar{g}}_\text {Na}\times 0\times }\left( V-V_1\right) \nonumber \right. \\&\quad \left. -{{\bar{g}}_\text {K}\times 0\times }\left( V-V_1\right) -g_\text {leak}(V-V_1)\right\} \end{aligned}$$
(116)
$$\begin{aligned}&=\frac{1}{C}\left\{ I_\text {app}(t)-g_\text {leak}(V-V_1)\right\} . \end{aligned}$$
(117)

Inequality (115) holds with probability 1 because \(V_1 ={\text {min}}_{i\in {\mathcal {I}}} \wedge V_\text {leak}\), and inequality (116) follows because \(V-V_1\le 0\), \(g_i>0\) and \({M_{31}}\ge 0\), \({N_4}\ge 0\). Let \(V_\text {min}:=\text {min}\left\{ \frac{I_-}{g_\text {leak}} +V_1,V_1\right\} \). When \(V<V_\text {min}\), \(\frac{\hbox {d}V}{\hbox {d}t}>0\). Therefore, V will not decrease beyond \(V_\text {min}\).

Similarly, when \(V\ge V_2\)

$$\begin{aligned} \frac{\hbox {d}V}{\hbox {d}t}&=\frac{1}{C}\left\{ I_\text {app}(t)-{\bar{g}}_\text {Na}{M_{31}}\left( V-V_\text {Na}\right) \nonumber \right. \\&\quad \left. -{\bar{g}}_\text {K}{N_4}\left( V-V_\text {K}\right) -g_\text {leak}(V-V_\text {leak})\right\} \end{aligned}$$
(118)
$$\begin{aligned}&\le \frac{1}{C}\left\{ I_\text {app}(t)-{\bar{g}}_\text {Na}{M_{31}}\left( V-V_2\right) \nonumber \right. \\&\left. \quad -{\bar{g}}_\text {K}{N_4}\left( V-V_2\right) -g_\text {leak}(V-V_2)\right\} \end{aligned}$$
(119)
$$\begin{aligned}&\le \frac{1}{C}\left\{ I_\text {app}(t)-{{\bar{g}}_\text {Na}\times 0\times }\left( V-V_2\right) \nonumber \right. \\&\left. \quad -{{\bar{g}}_\text {K}\times 0\times }\left( V-V_2\right) -g_\text {leak}(V-V_2)\right\} \end{aligned}$$
(120)
$$\begin{aligned}&=\frac{1}{C}\left\{ I_\text {app}(t)-g_\text {leak}(V-V_2)\right\} . \end{aligned}$$
(121)

Inequality (119) and inequality (120) hold because \(V_2={\text {max}}_{i\in {\mathcal {I}}}\{V_i\}\vee V_\text {leak},\) \(V-V_2\ge 0\), \(g_i>0\) and \({M_{31}}\ge 0\), \({N_4}\ge 0\). Let \(V_\text {max}=\text {max}\left\{ \frac{I_\text {app}}{g_\text {leak}}+V_2, V_2 \right\} \). When \(V>V_\text {max},\) \(\frac{\hbox {d}V}{\hbox {d}t}<0\). Therefore, V will not go beyond \(V_\text {max}\).

We conclude that if V takes an initial condition in the interval \([V_\text {min},V_\text {max}],\) and if \(0\le M_{ij},N_i\le 1\) for all time, then V(t) remains within this interval for all \(t\ge 0\). Thus, we complete the proof of Lemma 1. \(\square \)

Experimental methods

Whole-cell current clamp recordings of Purkinje cells from in vitro cerebellar slice preparations taken from wild type and leaner mice were performed in the laboratory of Dr. David Friel (Case Western Reserve University School of Medicine), as described in Ovsepian and Friel (2008). Experimental procedures conformed to guidelines approved by the Institutional Animal Care and Use Committee at Case Western Reserve University. Voltage signals were sampled at a frequency of 20kHz, filtered at 5–10 kHz, digitized at a resolution of 32/mV and analyzed using custom software written in IgorPro and MATLAB.

Comparison of our main theorem with the related literature

1.1 Comparison with Giacomin et al. (2018)

Giacomin et al. (2018) considered a one-parameter family of strong solutions of the (Ito) stochastic differential equation

$$\begin{aligned} \hbox {d}X_t^\varepsilon =F(X_t^\varepsilon )\,\hbox {d}t+\varepsilon G(X_t^\varepsilon )\,\hbox {d}B_t, \end{aligned}$$
(122)

where the parameter \(\varepsilon > 0\), \(F(\cdot )\) is a locally Lipschitz vector field on \({\mathbb {R}}^d\), \(G(\cdot )\) is a locally Lipschitz function on \({\mathbb {R}}^d\) with values in the \(d\times m\) matrices and \(B_\cdot \) is a standard m dimensional Brownian motion. Note that in our formulation of the Langevin system (3), we use \(\sqrt{\epsilon }\) where Giacomin et al. use \(\varepsilon \). Consequently, our expressions involving \(GG^\intercal \) scale as \(\epsilon \), whereas the analogous expressions scale as \(\varepsilon ^2\) in Giacomin et al. (2018).

If the deterministic asymptotic phase is \(\theta (x)\), then evaluated along a trajectory \(\theta (X^\varepsilon _t)\) the phase obeys Ito’s formula:

$$\begin{aligned} \hbox {d}\theta (x)= & {} \left[ F_i(x)\partial _i\theta (x)\nonumber \right. \\&\left. +\frac{\varepsilon ^2}{2}{\mathcal {D}}_{ij}(x)\partial ^2_{ij}\theta (x)\right] \hbox {d}t\nonumber \\&+\varepsilon (\partial _i\theta (x))G_{ik}(x)\,\hbox {d}W_k(t), \end{aligned}$$
(123)

where \({\mathcal {D}}=GG^\intercal \) and we use Einstein’s summation convention. Given their definition of the asymptotic phase function and their Theorem 2.1 on page 1024, \(F_i(x)\partial _i\theta (x)=1\). As a result of their Theorem 2.3 (page 1026 in Giacomin et al. 2018), if \(\varepsilon \) is small, then over long times, we can establish that (since \(X^\varepsilon _t\) stays “close to” the deterministic limit cycle solution \(q_t=X^0_t\) for times on the order of \(t/\varepsilon ^2\)) approximately

$$\begin{aligned}&d\tilde{\theta }\approx \left( 1+\varepsilon ^2 b\right) \,dt+\varepsilon \sigma \, dW_t \end{aligned}$$
(124)

for some constants \(\sigma \) and b that one may calculate in terms of F and G for any particular system. The notation \({\tilde{\theta }}\) refers to the “lift” of \(\theta \) from the circle \(\theta \in {\mathbb {S}}_T\equiv [0,T)\) to \({\tilde{\theta }}\in {\mathbb {R}}\), i.e., the integral of (124) without the jumps induced by taking \(\theta \) mod the period T. For small noise \(\varepsilon \), integrating up to time t, we have (provided t is \(o(\varepsilon ^{-2})\), and setting \(x_0\equiv X_0^\varepsilon \)),

$$\begin{aligned} {\tilde{\theta }}(X^\varepsilon _t)\approx \theta (x_0)+t(1+\varepsilon ^2 b)+\varepsilon \sigma \,W_t, \end{aligned}$$
(125)

where \(W_t\) is now a 1D standard Brownian motion (with \(\approx \) signifying convergence in distribution, after rescaling time). Upon rescaling time, the authors establish that

$$\begin{aligned} {\tilde{\theta }}(X_{t/\varepsilon ^{2}}^\varepsilon )\approx \theta (x_0)+\frac{t}{\varepsilon ^2}+tb+\sigma \, W_t. \end{aligned}$$
(126)

Therefore, the accumulated variance of \({\tilde{\theta }}(X^\varepsilon _t)\) during one (deterministic) period T is

$$\begin{aligned} {\mathbb {V}}({\tilde{\theta }}(t+T)-{\tilde{\theta }}(t))&\approx {\mathbb {V}}\left( T+\varepsilon ^2bT+\varepsilon \sigma (W_{t+T}-W_t)\right) \end{aligned}$$
(127)
$$\begin{aligned}&=\varepsilon ^2\sigma ^2T \end{aligned}$$
(128)
$$\begin{aligned}&=\varepsilon ^2\int _0^T (D\theta (q_s)G(q_s)G^t(q_s)D\theta ^t(q_s))\,\hbox {d}s. \end{aligned}$$
(129)

Equation (128) holds because the variance of the Wiener process increment over time T is exactly T and because the limit cycle period T is a deterministic quantity.Footnote 7 Equation (129) holds by virtue of the authors’ formula, Eq. (2.20) on page 1026 in Giacomin et al. (2018).

Finally, we note that, for a conductance-based channel noise model of the type we consider in this paper, the expression \(D\theta (q_s)G(q_s)G^t(q_s)D\theta ^t(q_s)\) in Giacomin et al. (2018) corresponds exactly with our expression \(\sum _{ij}({\mathcal {G}}{\mathcal {G}}^\intercal )_{ij}\partial _{i}T_0\partial _{j}T_0\) and \(\sum _{k=1}^{28}\alpha _k(v){\mathbf {X}}_{i(k)} \left( \zeta _k^\intercal {\tilde{{\mathbf {Z}}}}\right) ^2,\) as in our Theorem 1 and Corollary 1. To see the correspondence, note that \(D\theta (q_s)\), the gradient of the phase function evaluated along the limit cycle is none other than the infinitesimal phase response curve and is identical to the gradient of the zero-noise timing function \(\nabla T_0\); the rest follows from the structure of \({\mathcal {G}}\) as shown in the proof of our main theorem.

1.2 Comparison with Aminzare et al. (2019)

In Aminzare et al. (2019), the authors consider the scaled Langevin model (Eq. (10) on page 4718)

$$\begin{aligned} \hbox {d}x={\mathbf {F}}(x)\hbox {d}t+\sigma B(x)\hbox {d}{\mathbf {W}}(t). \end{aligned}$$
(130)

The authors consider the special case in which \(B(x)=\text {diag}(\beta _1,\ldots ,\beta _n)\). Our corresponding matrix \({\mathcal {G}}\) is not diagonal, due to the biophysical origin of channel noise, so direct comparison of our results is somewhat limited. Nevertheless, the authors consider the moments of the stochastic period, defined as the time to complete one oscillation, and arrive at an expression for the variance of the stochastic period that we can compare with the result of our theorem 1. In particular, in Proposition 3 on page 4721 in Aminzare et al. (2019), the authors calculate the first and second moments of the stochastic return time T for system (130) with \(B(x)=\text {diag}(\beta _1,\ldots ,\beta _n)\) as

$$\begin{aligned} {\mathbb {E}}[T]=\frac{2\pi }{\omega }-\frac{\sigma ^2}{\omega ^2}\int _0^{2\pi }\varPi (\alpha )\,\hbox {d}\alpha +o(\sigma ^2), \end{aligned}$$
(131)

and

$$\begin{aligned} {\mathbb {E}}[T^2]= & {} \frac{4\pi ^2}{\omega ^2}-\frac{2\sigma ^2}{\omega ^3}\int _0^{2\pi }(2\pi -\alpha )\varPi (\alpha )\,\hbox {d}\alpha \nonumber \\&+\frac{2\sigma ^2}{\omega ^3}\int _0^{2\pi }\zeta (\alpha )\,\hbox {d}\alpha \nonumber \\&-\frac{2\sigma ^2}{\omega ^3}\int _0^{2\pi }\int _\xi ^{2\pi }\varPi (\alpha )\,\hbox {d}\alpha \, \hbox {d}\xi +o(\sigma ^2), \end{aligned}$$
(132)

where

$$\begin{aligned} \varPi (\theta _0)= & {} \frac{1}{2}\sum _{i=1}^n\beta ^2_iH_{ii}(\theta _0),\quad \text {and }\nonumber \\ \zeta (\theta _0)= & {} \frac{1}{2}\sum _{i=1}^n\beta ^2_iZ_{i}(\theta _0)^2, \end{aligned}$$
(133)

\(H(\theta )\) is the “second-order” PRC (the Hessian of the asymptotic phase function) and \(Z(\theta )\) is the infinitesimal phase response curve. Note that because (Aminzare et al. 2019) scales phase from 0 to \(2\pi \), their expression \(\zeta (\theta _0)\) has units of radians/time. In contrast, our infinitesimal phase response curve \({\mathbf {Z}}(t)\) has units of time; we may convert by multiplying by \(T_0/2\pi \) (Sect. 2.4). Although \({\mathcal {G}}\) is not diagonal for conductance-based neural oscillators driven by channel noise (cf. Eq. (3)), we can still express the term \(\zeta (\theta _0)\) as

$$\begin{aligned} \zeta (\theta _0)= & {} 2\pi ^2\sum _{ij}({\mathcal {G}}{\mathcal {G}}^\intercal )_{ij}\partial _{i}T_0\partial _{j}T_0\nonumber \\= & {} 2\pi ^2\sum _{k=1}^{28}\alpha _k(v){\mathbf {X}}_{i(k)} \left( \zeta _k^\intercal {\tilde{{\mathbf {Z}}}}\right) ^2. \end{aligned}$$
(134)

Given the first and second moments of the time period in Aminzare et al. (2019), the variance is given by

$$\begin{aligned} {\mathbb {V}}[T]&={\mathbb {E}}[T^2]-\left( {\mathbb {E}}[T]\right) ^2 \end{aligned}$$
(135)
$$\begin{aligned}&=\frac{4\pi ^2}{\omega ^2}-\frac{2\sigma ^2}{\omega ^3}\int _0^{2\pi }(2\pi -\alpha )\varPi (\alpha )\hbox {d}\alpha \nonumber \\&\quad +\frac{2\sigma ^2}{\omega ^3}\int _0^{2\pi }\zeta (\alpha )\hbox {d}\alpha -\frac{2\sigma ^2}{\omega ^3}\int _0^{2\pi }\int _\xi ^{2\pi }\varPi (\alpha )\hbox {d}\alpha \hbox {d}\xi \nonumber \\&\quad - \frac{4\pi ^2}{\omega ^2}+\frac{4\pi \sigma ^2}{\omega ^2}\nonumber \\&\quad \int _0^{2\pi }\varPi (\alpha )\hbox {d}\alpha + o(\sigma ^2)\ \end{aligned}$$
(136)
$$\begin{aligned}&=-\frac{4\pi \sigma ^2}{\omega ^3}\int _0^{2\pi }\varPi (\alpha )\hbox {d}\alpha +\frac{2\sigma ^2}{\omega ^3}\int _0^{2\pi }\alpha \varPi (\alpha )\hbox {d}\alpha \nonumber \\&\quad +\frac{2\sigma ^2}{\omega ^3}\int _0^{2\pi }\zeta (\alpha )\hbox {d}\alpha \nonumber \\&\quad -\frac{2\sigma ^2}{\omega ^3}\int _0^{2\pi }\int _\xi ^{2\pi }\varPi (\alpha )\hbox {d}\alpha \hbox {d}\xi \nonumber \\&\quad +\frac{4\pi \sigma ^2}{\omega ^2}\int _0^{2\pi }\varPi (\alpha )\hbox {d}\alpha + o(\sigma ^2) \end{aligned}$$
(137)
$$\begin{aligned}&\quad =\frac{2\sigma ^2}{\omega ^3}\int _0^{2\pi }\alpha \varPi (\alpha )\hbox {d}\alpha +\frac{2\sigma ^2}{\omega ^3}\int _0^{2\pi }\zeta (\alpha )\hbox {d}\alpha \nonumber \\&\quad -\frac{2\sigma ^2}{\omega ^3}\int _0^{2\pi }\int _\xi ^{2\pi }\varPi (\alpha )\hbox {d}\alpha \hbox {d}\xi . \end{aligned}$$
(138)

By Tonelli’s Theorem (Saks 1937),

$$\begin{aligned}&\int _0^{2\pi }\int _\xi ^{2\pi }\varPi (\alpha )\hbox {d}\alpha \hbox {d}\xi \end{aligned}$$
(139)
$$\begin{aligned}&\quad =\int \limits _{\begin{array}{c} \xi \le \alpha \le 2\pi \\ (\alpha ,\xi )\in [0,2\pi ]\times [0,2\pi ] \end{array}}\varPi (\alpha )d(\alpha ,\xi ) \end{aligned}$$
(140)
$$\begin{aligned}&\quad =\int _0^{2\pi }\int _0^{\alpha }\varPi (\alpha )\hbox {d}\xi \hbox {d}\alpha \end{aligned}$$
(141)
$$\begin{aligned}&\quad =\int _0^{2\pi }\left( \int _0^{\alpha }\varPi (\alpha )\hbox {d}\xi \right) \hbox {d}\alpha \end{aligned}$$
(142)
$$\begin{aligned}&\quad =\int _0^{2\pi }\alpha \varPi (\alpha )\hbox {d}\alpha . \end{aligned}$$
(143)

Therefore, Eq. (138) reduces to

$$\begin{aligned} {\mathbb {V}}[T]&=\frac{2\sigma ^2}{\omega ^3}\int _0^{2\pi }\zeta (\alpha )\hbox {d}\alpha \end{aligned}$$
(144)
$$\begin{aligned}&=\frac{2\sigma ^2}{\omega ^3}\times \frac{\omega ^2}{2\pi }\times 2\pi ^2\int _0^{T} \sum _{k=1}^{28}\alpha _k(v){\mathbf {X}}_{i(k)} \left( \zeta _k^\intercal {\tilde{{\mathbf {Z}}}}\right) ^2\hbox {d}t \end{aligned}$$
(145)
$$\begin{aligned}&=\sigma ^2T\sum _{k=1}^{28}\int _0^{T}\alpha _k(v){\mathbf {X}}_{i(k)} \left( \zeta _k^\intercal {\tilde{{\mathbf {Z}}}}\right) ^2\hbox {d}t \end{aligned}$$
(146)

which is consistent with our main result, because \(\sigma ^2=\epsilon \), \(\omega =2\pi /T\), and \(\hbox {d}\alpha \sim \omega \,\hbox {d}t\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pu, S., Thomas, P.J. Resolving molecular contributions of ion channel noise to interspike interval variability through stochastic shielding. Biol Cybern 115, 267–302 (2021). https://doi.org/10.1007/s00422-021-00877-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00422-021-00877-7

Keywords

Navigation