Skip to main content
Log in

Minimal time spiking in various ChR2-controlled neuron models

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

Abstract

We use conductance based neuron models, and the mathematical modeling of optogenetics to define controlled neuron models and we address the minimal time control of these affine systems for the first spike from equilibrium. We apply tools of geometric optimal control theory to study singular extremals, and we implement a direct method to compute optimal controls. When the system is too large to theoretically investigate the existence of singular optimal controls, we observe numerically the optimal bang–bang controls.

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
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19

Similar content being viewed by others

References

  • Adamantidis AR, Zhang F, Aravanis AM, Deisseroth K, de Lecea L (2007) Neural substrates of awakening probed with optogentic control of hypocretin neurons. Nature 450:420–424

    Article  Google Scholar 

  • Bonnard B, Kupka I (1993) Théories des singularités de l’application entrée/sortie et optimalité des trajectoires singulières dans le problème du temps minimal. Forum Math 5:111–159

    Article  MathSciNet  MATH  Google Scholar 

  • Boyden ES (2015) Optogenetics and the future of neuroscience. Nat Neurosci 18:1200–1201

    Article  Google Scholar 

  • Boyden ES, Zhang F, Bamberg E, Nagel G, Deisseroth K (2005) Millisecond-timescale, genetically targeted optical contro of neural activity. Nat Neurosci 8(9):1263–1268

    Article  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

    Article  MATH  Google Scholar 

  • Chen Y, Xiong M, Zhang SC (2015) Illuminating Parkinson’s therapy with optogenetics. Nat Biotechnol 33(2):149–150

    Article  Google Scholar 

  • Deisseroth K (2011) Optogenetics. Nat Methods 8:26–29

    Article  Google Scholar 

  • Deisseroth K (2015) Optogenetics: 10 years of microbial opsins in neuroscience. Nat Neurosci 18:1213–1225

    Article  Google Scholar 

  • Ditlevsen S, Greenwood P (2013) The Morris–Lecar neuron model embeds a leaky integrate-and-fire model. J Math Biol 67(2):239–259

    Article  MathSciNet  MATH  Google Scholar 

  • Feng J, Tuckwell HC (2003) Optimal control of neuronal activity. Phys Rev Lett 91(1): 018101/1–018101/4

  • FitzHugh R (1961) Impulses and physiological states in theoretical models of nerve membrane. Biophys J 1(6):445–466

    Article  Google Scholar 

  • Fourer R, Gay DM, Kernighan BW (2002) AMPL: a modeling language for mathematical programming. Duxbury Press, Belmont

    MATH  Google Scholar 

  • Foutz TJ, Arlow RL, McIntyre CC (2012) Theoretical principles underlying optical stimulation of a channelrhodopsin-2 positive pyramidal neuron. J Neurophysiol 107(12):3235–3245

    Article  Google Scholar 

  • Gaub BMMH, Berry AE, Holt EY Isacoff, Flannery JG (2015) Optogenetics vision restoration using rhodopsin for enhanced sensitivity. Mol Theory 23(10):1562–1571

    Article  Google Scholar 

  • Hegemann P, Ehlenbeck S, Gradmann D (2005) Multiple photocycles of channelrhodopsin. Biophys J 89:3911–3918

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Lecar H, Morris C (1981) Voltage oscillations in the barnacle giant muscle fiber. Biophys J 35:193–213

    Article  Google Scholar 

  • Lobo MK, Nestler EJ, Covington HE (2012) Potential utility of optogenetics in the study of depression. Biol Psychiatry 71(12):1068–1074

    Article  Google Scholar 

  • Lolov A, Ditlevsen S, Longtin A (2014) Stochastic optimal control of single neuron spike trains. J Neural Eng 11: 046004/1-046004/22

  • Nabi A, Moehlis J (2011) Single input optimal control for globally coupled neuron networks. J Neural Eng 8:3911–3918

    Article  Google Scholar 

  • Nagumo J, Arimoto S, Yoshizawa S (1962) An active pulse transmission line simulating nerve axon. Proc IRE 50(10):2061–2070

    Article  Google Scholar 

  • Nikolic K, Degenaar P, Toumazou C (2006) Modeling and engineering aspects of channelrhodopsin2 system for neural photostimulation. In: Proceedings of 28th IEEE engineering in medicine and biology society conference, pp 1626–1629

  • Nikolic K, Grossman N, Grubb MS, Burrone J, Toumazou C, Degenaar P (2009) Photocycles of channelrhodopsin-2. Photochem Photobiol 85:400–411

    Article  Google Scholar 

  • Paz JT, Huguenard JR (2015) Optogenetics and epilepsy: past, present and future. Epilepsy Curr 15(1):34–38

    Article  Google Scholar 

  • Pontryagin L, Boltyanski V, Gamkrelidze R, Michtchenko E (1974) Théorie mathématique des processus optimaux. Editions Mir, Moscou

    Google Scholar 

  • Rubin J, Wechselberger M (2008) The selection of mixed-mode oscillations in a Hodgkin–Huxley model with multiple timescales. Chaos 18(1):015105

    Article  MathSciNet  MATH  Google Scholar 

  • Ryan TJ, Roy DS, Pignatelli M, Arons A, Tonegawa S (2015) Engram cells retain memory under retrograde amnesia. Science 348(6238):1007–1013

    Article  Google Scholar 

  • Saint-Hilaire M, Longtin A (2004) Comparison of coding capabilities of type I and type II neurons. J Comput Neurosci 16:299–313

    Article  Google Scholar 

  • Trélat E (2008) Contrôle optimal: théorie et applications. Vuibert, Paris

    MATH  Google Scholar 

  • Trélat E (2012) Optimal control and applications to aerospace: some results and challenges. J Optim Theory Appl 154(3):713–758

    Article  MathSciNet  MATH  Google Scholar 

  • Wong J, Abilez OJ, Kuhl E (2012) Computational optogenetics: a novel continuum framework for the photoelectrochemistry of living systems. J Mech Phys Solids 60(6):1158–1178

    Article  Google Scholar 

  • Wächter A, Biegler LT (2006) On the implementation of an interior-point filer line-search algorithm for large-scale nonlinear programming. Math Program 106:25–27

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Vincent Renault.

Appendices

Appendix 1: Numerical constants for the Morris–Lecar model

The numerical values of the several constants and their physiological meaning are taken from Ditlevsen and Greenwood (2013) and gathered in Table 5.

Table 5 Meaning and numerical values of the constants appearing in the Morris–Lecar model

Table 6 gathers the numerical values for Fig. 10.

Table 6 Meaning and numerical values of the constants, taken from Saint-Hilaire and Longtin (2004), appearing in the Morris–Lecar model

Appendix 2: Numerical constants for the Hodgkin–Huxley model

$$\begin{aligned} \alpha _n(V)&= \frac{0.1-0.01V}{e^{1-0.1V}-1},&\beta _n(V)&= 0.125e^{-\frac{V}{80}},\\ \alpha _m(V)&= \frac{2.5-0.1V}{e^{2.5-0.1V}-1},&\beta _m(V)&= 4e^{-\frac{V}{18}},\\ \alpha _h(V)&= 0.07e^{-\frac{V}{20}},&\beta _h(V)&= \frac{1}{e^{3-0.1V}+1}. \end{aligned}$$

The following Table 7 gathers the numerical values of the Hodgkin–Huxley model, as given in the original paper Hodgkin and Huxley (1952).

Table 7 Meaning and numerical values of the constants appearing in the Hodgkin–Huxley model

The equilibrium potential \(E_L\) of the leakage current is usually set so that the equilibrium value of the (HH) system is such that \(V=0\).

Appendix 3: Numerical constants for the ChR2 models

1.1 The 3-states model

The constants of the model are the rates \(K_d\) and \(K_r\) of the transitions between the open state and the light adapted closed state and between the two closed states, the maximal conductance \(g_{ChR2}\) and the equilibrium potential \(V_{ChR2}\). As specified in Sect. 3, we assume that these rates are constants during the evolution in order to obtain an affine control system. For the numerical computations, we took the values given in Table 1 of Nikolic et al. (2009):

$$\begin{aligned} K_d = 0.2\text { m\,s}^{-1}, \quad K_r = 0.021 \text { m\,s}^{-1}. \end{aligned}$$

The maximal conductance is given by the formula \(g_{ChR2} = \rho _{ChR2}g^*_{ChR2}\), with \(\rho _{ChR2}\) the density of channels and \(g^*_{ChR2}\) the conductance of a single channel. These values are taken from Foutz et al. (2012) to obtain

$$\begin{aligned} g_{ChR2} = 0.65 \text { mS}\, \text {cm}^{-2}. \end{aligned}$$

As mentioned at the end of “Appendix 2”, the physiological equilibrium membrane potential is mathematically shifted to equal 0. The equilibrium potential of the ChR2 that is usually measured around 0 (Foutz et al. 2012) and very often taken as 0 (Foutz et al. 2012; Nikolic et al. 2009). The exact value 0 would raise a mathematical problem because since we shifted the value of \(E_L\) so that \(V=0\) corresponds to the equilibrium point of the uncontrolled system we start from. Indeed, \(V=0\) would also correspond to an equilibrium point of the controlled system, regardless of the value of the control. For this reason, we shifted the value of \(V_{ChR2}\) and took it equal to 60mV. This value corresponds to the shift of the membrane resting potential for the Morris–Lecar and Hodgkin–Huxley models.

Finally we can give an estimation of the physiological maximal value \(u_{max}\) of the control. Indeed, upon illumination, the transition rate between the dark adapted closed state and the open state in Nikolic et al. (2009) is \(\varepsilon F\) where \(\varepsilon =0.5\) is the quantum efficiency and F is given by the formula

$$\begin{aligned} F = \frac{\sigma _{ret} \phi }{w_{loss}}, \end{aligned}$$

where \(\sigma _{ret} \simeq 10^{-8} \upmu \text {m}^2\) is the retinal cross section (cross section of the photon receptor on the ChR2), \(\phi = 6.2\times 10^{9}\text { ph}\cdot \upmu \text {m}^{-2}\,\text {s}^{-1}\) is the original flux of photons and \(w_{loss} =1.1\) is a loss factor. As for the numerical value of \(K_d\) and \(K_r\) we took the one of Table 1 in Nikolic et al. (2009) for the value of \(\phi \). With these values we get

$$\begin{aligned} u_{max} = 0.028 \text { m\,s}^{-1}. \end{aligned}$$

1.2 The 4-states model

The numerical values for the ChR2-4-states model are taken from Foutz et al. (2012) and gathered in Table 8.

Table 8 Numerical values of the constants appearing in the ChR2-4-States model

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Renault, V., Thieullen, M. & Trélat, E. Minimal time spiking in various ChR2-controlled neuron models. J. Math. Biol. 76, 567–608 (2018). https://doi.org/10.1007/s00285-017-1101-1

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-017-1101-1

Keywords

Mathematics Subject Classification

Navigation