CALIFORNIA STATE UNIVERSITY, NORTHRIDGE AN APPROXIMATE ANALYTICAL SOLUTION FOR THE EXCITATION THRESHOLD IN A ONE-DIMENSIONAL FITZHUGH-NAGUMO SYSTEM A thesis submitted in partial fulfillment of the requirements For the degree of Master of Science in Physics By D’Artagnan G. Greene August 2014 The thesis of D’Artagnan G. Greene is approved: Date Dr. Miroslav Peric Date Dr. Donna Sheng Date Dr. Yohannes Shiferaw, Chair California State University, Northridge ii ACKNOWLEDGEMENTS I would especially like to thank my research advisor Yohannes Shiferaw for his patience and guidance throughout this entire thesis project. I would also like to thank all of my other professors at California State University Northridge for providing me with the background that I needed to be able to handle a project like this. Last, but not least, I would like to thank my parents. Without their sustained encouragement and support throughout my education, this thesis surely would not exist. iii TABLE OF CONTENTS SIGNATURE PAGE . . . . . . . . . ii ACKNOWLEDGEMENTS . . . . . . . . iii LIST OF FIGURES . . . . . . . . . vii ABSTRACT . . . . . . . . . . x INTRODUCTION . . . . . . . . . 1 . . . . 4 . . . 4 . . . 8 Hodgkin and Huxley’s Expressions for the Conductance of the Various Ion Channels . . 10 The Diffusion of the Potential Along a Chain of Cells: The Cable Equation . . 15 . . . 18 . . . 20 . . 29 . 29 Moving in the Direction of an Analytical Theory for the Threshold Surface: Projected Dynamics 36 1. BASIC MODELS OF THE ACTION POTENTIAL Establishing the Resting Potential Across the Plasma Membrane for a Cell Modeling Using the Equivalent Circuit Approach . . The General Form of an Equation Describing an Action Potential . . A Simplified Model of the Action Potential: The Fitzhugh-Nagumo Model 2. ANALYTICAL APPROACHES TO STUDYING THE EXCITATION THRESHOLD . . . . . . . Studying Threshold Phenomena for a Linear Chain of Cells: The McKean and Moll Approach The Formal Derivation of Projected Dynamics . iv . . . . . 38 . . . . . 43 Analysis of the Projected Dynamics Using a Gaussian Pulse . . . . 48 . . . 53 Applying Projected Dynamics Using a Gaussian Pulse Discussion of the Results of the Projected Dynamics Using a Gaussian Pulse 3. FORMULATION OF THE THRESHOLD EQUATION FOR A SQUARE PULSE USING INITIAL CONDITIONS . . . . . . . 55 A New Formulation: Changing the Initial Voltage Profile to a Square Pulse . . . 55 Applying Projected Dynamics for a Square Pulse Using Initial Conditions . . . 62 Analysis of the Threshold Equation for a Square Pulse Using Initial Conditions . . 64 . . 71 . . 71 . . . 74 . . . 77 4. FORMULATION OF THE THRESHOLD EQUATION FOR A CONSTANT SQUARE PULSE . . . . . . . From Using the Initial Conditions for a Square Pulse to a Constant Square Pulse Reformulating the Projected Dynamics Using a Constant Square Pulse Analysis of the Threshold Equation for a Constant Square Pulse . 5. FORMULATION OF THE THRESHOLD EQUATION FOR A TIME-DEPENDENT SQUARE PULSE . . . . . . . . . 82 . . 82 . . 86 6. A SIMPLIFIED APPROXIMATE SOLUTION FOR THE SQUARE PULSE . 92 Generalizing the Approach to find the Threshold for a Time-Dependent Square Pulse Obtaining the Threshold Equation for a Time-Dependent Square Pulse . Introducing a Simplified Initial Voltage Profile for the Constant Square Pulse . v . . 92 The Projected Dynamics and the Threshold Equation for the Simple Constant Square Pulse The Threshold Equation for the Simple Time-Dependent Square Pulse . 94 . . . 97 CONCLUSION . . . . . . . . . 102 REFERENCES . . . . . . . . . 104 . . . . . . 106 APPENDIX: NUMERICAL METHODS vi LIST OF FIGURES . . 14 2. The fast phase plane of the reduced Hodgkin and Huxley system as plotted by Fitzhugh. . 22 . . 23 1. The solution of the Hodgkin and Huxley equations displaying an action potential. 3. Plots of the time course of the various gating variables during an action potential. 4. The fast-slow phase plane for the Hodgkin and Huxley model. . . . . 24 5. The equivalent circuit for the Fitzhugh-Nagumo system. . . . . . 26 6. A propagating wave “front”; a propagating wave “pulse.” . . . . . 30 . . . 31 7. The cubic function used for excitation in the Fitzhugh-Nagumo system. . . . . . . 32 9. The initial waveform studied by McKean and Moll. . . . . . 33 10. The threshold surface visualized by McKean and Moll. . . . . . 34 . . . 35 . . . . 47 . . . . 48 . . 50 . 51 8. The piecewise-linear function used by McKean. 11. Numerical plots of the critical amplitude (height) vs. the pulse width (base). 12. The phase plane for the projected dynamics with a Gaussian pulse. 13. Comparison of Figures 10-12. . . . . 14. Comparison of the amplitude nullclines obtained using projected dynamics vs. the numerical threshold solution. . . . . . . 15. Plot of the critical amplitude (πΌπ ) vs. the ion channel conductivity (π) for the projected dynamics with a Gaussian pulse. . . . vii . . . . 16. Plot of the critical amplitude (πΌπ ) vs. the longitudinal conductivity (π·) for the projected . . . . . . . . 52 17. A sample plot of a square pulse. . . . . . . . . 55 . . . 64 . . . 67 . . . 68 . . 69 . . 72 . . 77 . . 79 . . 79 dynamics with a Gaussian pulse. 18. Plot of the critical amplitude (πΌπ ) vs. the pulse width (π€) for the projected dynamics with a square pulse. . . . . . 19. Plot of the critical amplitude (πΌπ ) vs. the potassium ion channel conductivity (ππ ) for the projected dynamics with a square pulse. . . . 20. Plot of the critical amplitude (πΌπ ) vs. the sodium ion channel conductivity (πππ ) for the projected dynamics with a square pulse. . . . 21. Plot of the critical amplitude (πΌπ ) vs. the longitudinal conductivity (π·) for the projected dynamics with a square pulse. . . . . . . 22. A comparison of the numerical results for the critical amplitude vs. the pulse width from two different approaches to applying a square pulse. . . . 23. A plot of the critical amplitude vs. pulse width for a constant applied square pulse. 24. A plot of the critical amplitude vs. potassium ion channel conductivity for a constant applied square pulse. . . . . . . . 25. A plot of the critical amplitude vs. sodium ion channel conductivity for a constant applied square pulse. . . . . viii . . . 26. A plot of the critical amplitude vs. longitudinal conductivity for a constant applied square pulse. . . . . . . . . . . 80 . . 85 . . 90 . . 96 . 99 27. A plot of the critical amplitude vs. the time that the pulse was applied for a uniformly applied square pulse. . . . . . . . 28. A plot of the critical amplitude vs. the time that the square pulse was applied is provided for both a 100 cell pulse width and a 5 cell pulse width. . . . 29. A plot of the critical amplitude vs. pulse width for the simple constant applied square pulse. . . . . . . . . . 30. A plot of the critical amplitude vs. the time that the square pulse was applied is provided for both a 100 cell pulse width and a 5 cell pulse width. Here, the simple square pulse approach is compared with the numerical result. . . ix . . . . ABSTRACT AN APPROXIMATE ANALYTICAL SOLUTION FOR THE EXCITATION THRESHOLD IN A ONE-DIMENSIONAL FITZHUGH-NAGUMO SYSTEM By D’Artagnan G. Greene Master of Science in Physics Understanding the nature of electrical excitation of a group of cells is important both in examining the onset of a cardiac arrhythmia and in designing the treatment for sudden cardiac arrest. In the past, several attempts have been made to understand the threshold for the excitation of a one-dimensional chain of cells from a mathematical viewpoint. However, obtaining an analytical solution to describe threshold phenomena has proven to be difficult as the equations in this problem are highly non-linear and resist solution by standard mathematical techniques. Here, we apply a method developed by Neu et al. where the time evolution of the width and amplitude of a pulse is approximately described by a gradient flow on a two-dimensional phase plane. Using this approach, we obtain a mathematical expression that successfully models the excitation threshold for an applied square current pulse in a simplified Fitzhugh-Nagumo system. We then analyze our solution to reveal how the excitation threshold depends on key physiological parameters. x INTRODUCTION Certain cells are termed “excitable” if they respond to an externally applied electric current in a characteristic way. If the applied electric current is large enough, the membrane potential for the cell will carry out a characteristic rise and fall pattern that is called an action potential. If the applied current is too small, an action potential will not be observed, and the membrane potential will quickly return to rest. Furthermore, many cells can be grouped together whereby if some of the cells are excited they can induce nearby cells to excite. This can lead to a propagating electric signal across an entire group of cells which is referred to as an excitation wave. The excitation threshold is defined as the minimum amount of electric current that must be applied to excite the cell(s). This threshold plays a particularly important role in cardiac systems. For example, cardiac arrhythmias often arise from unintended excitation of regions of the heart. At the other extreme, medical defibrillators attempt to manually excite cells in regions of the heart that have become completely inactive during sudden cardiac arrest. Yet, despite its importance, the threshold phenomenon is not completely understood. This thesis consists of six sections. In the first section, we will look at two early mathematical models of excitable systems. First, the pioneering Hodgkin and Huxley model will be covered in detail [1], and this will be followed by the more mathematically simple Fitzhugh-Nagumo model [2,3] which is of the form: ππ£ ∂2 π£ = − π£(πΌ − π£)(1 − π£), ππ‘ ∂π₯ 2 1 where πΌ is the threshold parameter. Even using the simplified Fitzhugh-Nagumo system, the non-linearity of this equation makes analysis of the threshold phenomenon particularly difficult. In the second section, we will examine two significant approaches to characterizing the excitation threshold using simplified versions of the Fitzhugh-Nagumo system for a linear chain of cells. The first of these was carried out by H.P. McKean and V. Moll who showed that the excitation threshold can be visualized as a surface in phase space [4]. Their approach was later used as a starting point by J.C. Neu et al. who developed a method called projected dynamics. The projected dynamics made use of a Gaussian-shaped current pulse to project a pair of approximate solutions for the time evolution of the pulse amplitude and pulse width onto a two-dimensional phase plane. The threshold was observed as that set of trajectories on the phase plane which did not lead to either an excitation or a decay back to rest [5]. In the third section, we show that the projected dynamics can be further refined by replacing the Gaussian pulse with one that more closely resembles a square-shaped physiological current pulse. Doing this corrects some of the defects in the Gaussian approach such as the wrong threshold limits being reached in the case of large and small pulse widths. At the same time, we will obtain an approximate analytical expression for the excitation threshold by noting its similarity to the amplitude nullcline of the projected dynamics. In section four, more realism will be added to the model as the applied current pulse is changed from being specified using initial conditions to an externally applied current pulse that is held constant throughout the excitation process. In making this change we will discover that the criteria for the excitation threshold will have to be modified. More 2 importantly, this modification brings the approximate analytical expression for the threshold into very close agreement with the numerical result obtained directly from the partial differential equation. In section five, we ease the restriction on the time duration of the applied current pulse so that the current pulse can be applied for a finite amount of time. In doing so, we will obtain the most general solution for the excitation threshold. It will be shown that this general solution consists of both the solution for the constant square pulse and an additional timedependent term. Finally, in section six, we will introduce a simplification to the model developed in sections four and five. The simple constant square pulse yields a threshold solution that, at the expense of a little accuracy, will make the threshold solution more compact and easier to work with while retaining the correct limiting behavior of the more complicated model that came before. 3 1. BASIC MODELS OF THE ACTION POTENTIAL Establishing the Resting Potential Across the Plasma Membrane for a Cell We will start by describing the electrical properties of a lone cell such as a neuron. The mathematical treatment will follow the now standard treatments of the subject such as that which is found in G.B. Ermentrout and D.H. Terman’s Mathematical Foundations of Neuroscience [6]. A cell in general terms is essentially a collection of organelles, small chemical factories that serve some biological purpose for the cell, which are placed in a mostly aqueous environment that is separated from the aqueous environment on the outer side of the cell by a non-polar plasma membrane. The electrical properties of such a cell are localized at this plasma membrane and are brought about by the flow of charged ions across this membrane. The plasma membrane itself is impermeable to charged ions, but small protein channels are inserted into the membrane that allow a specific charged ion to move across the membrane while excluding other ions from passing through. For example, there are channels in the membrane that allow for sodium ions to move across the membrane, but do not allow potassium ions to pass through. And then there are other channels that allow potassium ions to pass through but block sodium. Once an ion channel is opened, the flow or diffusion of that ion across the membrane is determined by two competing processes: one that is based on concentration and another that depends on the electric potential. As an example, consider a cell with a plasma membrane that separates an electrically neutral internal aqueous solution containing potassium and chloride ions from an external aqueous solution that also contains an equivalent amount of potassium and chloride ions. If a channel protein that selectively allowed passage to potassium ions is now inserted into the 4 plasma membrane, no net change would be observed. This is due to the concentration of the potassium ions inside and outside being the same when the channel protein is inserted. On the other hand, suppose that initially the number of potassium ions on the inside of the cell is greater than the number of potassium ions outside the cell. When the channel protein is inserted, there would be a net diffusion of potassium ions from the inside of the cell to the outside as the concentrations tend to spontaneously equalize. This diffusion is described mathematically by Fick’s law of diffusion: π½π· = −π· π[πΆ] . ππ¦ (1.1) In this expression, the concentration [πΆ] is the number of potassium ions per unit volume, π· is Fick’s diffusion constant, and π½π· is the flux of ions flowing across the cross-sectional area of the plasma membrane. The π¦ coordinate is measured from the inside of the channel protein to the outside, and the negative sign implies that the flow tends to be from a high concentration of ions to a low concentration. If the potassium ions were uncharged, this description would be sufficient. However, potassium ions carry an electric charge of +1 along with them. As the potassium ions diffuse across the membrane in an attempt to equalize their concentrations, a charge imbalance begins to appear. The outer side of the plasma membrane becomes more positively charged and the inside of the plasma membrane more negatively charged. Thus a potential difference starts to build up that resists the flow of the diffusing ions. This potential flux can be described by: π½π = −ππ§[πΆ] ππ . ππ¦ 5 (1.2) Here π represents the mobility of the ion, π§ is the valence of the ion (in the case of potassium it would be +1), [πΆ] is the concentration, and π is the electric potential. Here, the negative sign indicates that the flow is in the direction of decreasing potential. The total flux of ions across the membrane is then described as the sum of these two competing effects: π½ = π½π· + π½π = −π· π[πΆ] ππ − ππ§[πΆ] . ππ¦ ππ¦ (1.3) Albert Einstein developed a theory that expresses the diffusion constant π· in terms of its mobility π, the fundamental charge ππ = 1.60 x 10-19 C, Boltzmann’s constant π = 1.38 x 10-23 J/K, and the absolute temperature π: ππ π. ππ π·= (1.4) This can be substituted for π· in equation (1.3) to give: π½= − ππ π[πΆ] ππ π − ππ§[πΆ] . ππ ππ¦ ππ¦ (1.5) After some time has passed, an equilibrium may be established whereby it is observed that the net concentration of potassium ions on either side of the membrane is no longer changing in time. This condition is satisfied when the flux vanishes and π½ = 0. The potential at equilibrium is called the Nernst potential, and can be found as follows: 0= ππ π[πΆ] ππ + π§[πΆ] , ππ ππ¦ ππ¦ ππ π[πΆ] , π§ππ [πΆ] (1.7) ππ [πΆππ] π[πΆ] ∫ , π§ππ [πΆππ’π‘] [πΆ] (1.8) ππ = − ∫ πππ πππ’π‘ ππ = − (1.6) 6 ππ = πππ − πππ’π‘ = ππ [πΆππ’π‘ ] ln ( ). π§ππ [πΆππ ] (1.9) Multiplying and dividing the right side by Avogadro’s number ππ = 6.022 x 1023 mol-1 yields: ππ = ( πππ΄ π [πΆππ’π‘ ] ) ln ( ), ππ ππ΄ π§ [πΆππ ] ππ = π π [πΆππ’π‘ ] ln ( ). π§πΉ [πΆππ ] (1.10) (1.11) In equation (1.11), π is the ideal gas constant given as 8.314 J/(mol K), and πΉ is Faraday’s constant which is approximately 96,352 C/mol. At rest, the concentration differences for sodium and potassium ions between the inside and the outside of the cell are established by the use of a sodium-potassium pump which requires an external energy input to move the ions against their concentration gradients. Typical values of the concentrations for a cell in the squid giant axon at rest are given in J. Keener and J. Sneyd’s Mathematical Physiology [7] and are listed in Table 1 together with the Nernst potential for each ion which was calculated using equation (1.11) at an assumed body temperature of 37β. The squid giant axon was used in the original studies carried out by Hodgkin and Huxley [1]. The resting potential of the plasma membrane as a whole for the squid giant axon is listed as -66 mV. This value is a net result of the contributions made by the various ion channels, and it depends on the relative number of open channels for each ion while the cell is at rest. In particular, the large negative value indicates that sodium channels play a much smaller role in the overall resting potential than the potassium channels do. We will return to this point shortly. 7 [Cin] [Cout] VN (mmol) (mmol) (mV) K+ 397 20 -80 Na+ 50 437 58 Cl- 40 556 -70 Table 1. Concentrations and Nernst Potentials for the Squid Giant Axon [7]. Modeling Using the Equivalent Circuit Approach Once the rest state is established, the next step is to capture the behavior of a nerve cell when the voltage across the membrane is not at the resting potential. A popular approach is to use the concept of the equivalent circuit first developed by L. Lapicque [8]. This approach treats the charge separation between the inside and the outside of the cell as a capacitor with a potential difference across it. A channel protein is treated as both a battery at the Nernst potential and a resistor that impedes current flow. In such a setup, charge will flow from the capacitor through the channel proteins until the Nernst potential is reached. Consider a plasma membrane that contains channels which only allow potassium ions to cross. The capacitance of the membrane as a whole can be described by: πΆ= π , π (1.12) where π is the net ionic charge that is stored on either side of the membrane while V is the membrane potential established for that charge. The current for the capacitor can be obtained as: 8 π = πΆπ, πΌπΆ = (1.13) ππ ππ =πΆ . ππ‘ ππ‘ (1.14) The other potential changes in the circuit occur as the charge moves through an ion channel across the resistor and the battery. This can be expressed as a current following Ohm’s law: πΌπΎ = (π − ππΎ ) . π πΎ (1.15) In (1.15), π is the potential difference due to the capacitor, π πΎ is the resistance in the potassium channel, and ππΎ is the Nernst potential for potassium. Since there is no external input of current, Kirchhoff’s loop rules require that the sum of these two currents vanish to give: 0 = πΌπΆ + πΌπΎ , 0= πΆ ππ (π − ππΎ ) + , ππ‘ π πΎ (1.16) (1.17) ππ (1.18) = −πΊπΎ (π − ππΎ ). ππ‘ Here, πΊπΎ is the conductance which is equal to 1⁄π πΎ . Other ion channels, such as sodium πΆ or chloride channels, can be connected in parallel to this circuit which simply extends the result to: πΆ ππ = −πΊπΎ (π − ππΎ ) − πΊππ (π − πππ ) − πΊπΆπ (π − ππΆπ ). ππ‘ (1.19) If an external current is now applied to the cell, as would be the case when an electrode is applied to inject current into the cell as an example, equation (1.16) becomes: πΌ = πΌπΆ + πΌπΎ . (1.20) Carrying the steps through as before, the expression for the equivalent circuit is then: 9 πΆ ππ = − πΊπΎ (π − ππΎ ) − πΊππ (π − πππ ) − πΊπΆπ (π − ππΆπ ) + πΌ. ππ‘ (1.21) Hodgkin and Huxley’s Expressions for the Conductance of the Various Ion Channels Equation (1.21) can be used as a basis to understand the equivalent circuit of a cell with various ion channels, but such an equation in and of itself does not clearly describe how an action potential may come about. It was Hodgkin and Huxley who first developed a theory that described how the conductance for an ion channel may change with time depending on changes in the potential difference across the plasma membrane, and they explained how these changes gave rise to an action potential [1]. They distinguished between two types of channels present in the membrane that could conduct ions; there were non-gated channels, which were channels that were always open regardless of the electrical state of the membrane, and there were voltage-gated channels which could change from being open to closed or vice versa depending on the present value of the potential difference across the membrane in the vicinity of the channel. All of the non-gated channels were grouped together into one equivalent resistor, and the current that ran through it was termed the “leak” current. This equivalent resistor included many channels that played a role in establishing the resting potential such as the chloride channels. Introducing this leak current, equation (1.19) became: πΆ ππ = − πΊπΎ (π − ππΎ ) − πΊππ (π − πππ ) − πΊπΏ (π − ππΏ ). ππ‘ (1.22) Since the channel proteins that comprised the leak current were always open, the conductance πΊπΏ was found to be a constant independent of the potential difference across 10 the membrane, and it could be found experimentally [1]. On the other hand, the sodium channels and some of the potassium channels were discovered to have conductance values that changed as a function of time. Hodgkin and Huxley proposed the following two equations to describe the conductance for the voltage-gated potassium channels: πΊπΎ = πΊΜ πΎ π4 , (1.23) ππ = πΌπ (1 − π) − π½π π. ππ‘ (1.24) In equation (1.23), π is a parameter that can take any value between 0 and 1 while πΊΜ πΎ is a constant equal to the maximum value of the potassium conductance. Equation (1.24) describes the manner by which π changes with time. Suppose that there are two states for the potassium channel, a state π that is open (π) and a state (1 − π) that is closed (πΆ). Further suppose that a chemical equilibrium is established between the two states as shown below [6]: πΆ πΌπ β π½π (1.25) π. In this context, πΌπ determines the rate at which a closed state is converted into an open state and π½π determines the rate at which an open state is converted into a closed state. Both πΌπ and π½π depend on the potential difference of the membrane but are otherwise independent of time. The rate of conversion from the closed state to the open state (the forward process) is given by: πππ (1.26) = πΌπ (1 − π), ππ‘ and the rate of the reverse process, the conversion of the open state back to a closed state, is given by: 11 πππ = − π½π π. ππ‘ (1.27) The net change in π is the sum of (1.26) and (1.27) which is exactly (1.24). Equation (1.24) can be cast in a slightly different form. First, the steady state solution of π can be found by setting ππ/ππ‘ = 0 and solving for π∞ : 0 = πΌπ (1 − π∞ ) − π½π π∞ , π∞ = πΌπ . πΌπ + π½π (1.28) (1.29) π∞ can then be reintroduced into equation (1.24): ππ = πΌπ − (πΌπ + π½π )π, ππ‘ (1.30) ππ = (πΌπ + π½π )π∞ − (πΌπ + π½π )π, ππ‘ (1.31) ππ (π∞ − π) = , ππ‘ ππ (1.32) where ππ is a time constant given by: ππ = 1 . (πΌπ + π½π ) (1.33) The sodium conductance is similar, but contains two types of voltage dependent parameters, π and β: πΊππ = πΊΜ ππ π3 β. (1.34) The expressions for the time dependence of the parameters π and β are obtained in the same way as before: ππ (π∞ − π) = , ππ‘ ππ πΌπ π∞ = , πΌπ + π½π 12 (1.35) (1.36) ππ = 1 , (πΌπ + π½π ) (1.37) πβ (β∞ − β) = , ππ‘ πβ πΌβ β∞ = , πΌβ + π½β (1.38) 1 . (πΌβ + π½β ) (1.40) πβ = (1.39) At this point the various equations can be solved if each of the πΌ and π½ expressions are given along with numerical values for the capacitance, the maximum values of each conductance, and values of the Nernst potential for each channel type. Hodgkin and Huxley obtained the required expressions and numerical values from their experimental data [1,6]. A plot of the action potential and the associated time dependent sodium and potassium conductance is provided in Figure 1. Hodgkin and Huxley explained how the action potential arose from the time-dependent changes in the conductivities. An initial external stimulus applied to the cell raised the membrane potential to a certain level until suddenly the πΊππ conductance began to rapidly grow. This growth was determined by the parameter π which increased rapidly at first to signify that the sodium channels were opening and that sodium ions were rushing into the cell. The membrane voltage then rose towards the Nernst potential for sodium of +58 mV. However, as the membrane voltage climbed upwards, the parameter β increased and tended to close the sodium channels by blocking the pathway for the sodium ions. The sodium conductance overall quickly peaked and then dropped back down to a non-conducting state. 13 Figure 1. The solution of the Hodgkin and Huxley equations displaying an action potential. The lower graph shows the time dependent sodium and potassium conductivities [6]. Meanwhile, the potassium channels governed by the parameter π tended to open more at higher values of the potential. With the sodium channels now inactivated by the parameter β, the potassium conductivity became the dominant effect, and the membrane potential returned to the resting potential of -66 mV to complete the action potential. After an action potential is fired, there is a period of time, called the refractory period, where the ion concentrations are re-established at their original resting levels by the sodium-potassium pump. Once the refractory period is complete, the cell is ready to fire an action potential once again. 14 The Diffusion of the Potential Along a Chain of Cells: The Cable Equation So a question might naturally arise at this point: Why would a cell transmit electric signals by using an action potential? The procedure is quite complex, and a cell can surely receive an electric signal and return to rest just as well without undergoing such a cycle. The advantage lies not with the single cell, but rather it comes into play when an electric signal is propagated from one cell to another across a long chain of cells. To see this, consider a linear chain of cells that are attached end to end to one another in the shape of a cylindrical wire. If we were to stimulate one end of the chain with a current pulse, the plasma membrane of the cells near the stimulus would be at a higher potential than the cells much further down the chain. Assuming that charge can flow between cells, a potential difference would be set up along the chain and charge would flow from the end near the stimulus towards the distant cells at a lower potential. This longitudinal motion of charge can be described as a current flowing along the π₯-axis in the direction of the wire [6]. Following Ohm’s law, we have: π(π₯ + βπ₯, π‘) − π(π₯, π‘) = πΌπ· π π· . (1.41) Since we have assumed a cylindrical shape, we can use an expression for the resistance of a cylindrical wire with radius π, resistivity π, and a small change in length βπ₯: π π· = π βπ₯ , ππ2 (π(π₯ + βπ₯, π‘) − π(π₯, π‘)) π = πΌπ· 2 . βπ₯ ππ (1.42) (1.43) Taking the limit of (1.43) as βπ₯ → 0 gives an expression for the longitudinal diffusion current: 15 ππ2 ππ(π₯, π‘) πΌπ· = . π ππ₯ (1.44) Since this is a current that can leave one cell and enter another, Kirchhoff’s loop rules demand for each individual cell that we have: πΌπ· = πΌπΆ + πΌππ + πΌπΎ + πΌπΏ . (1.45) Now, before going any further, it must be said at this point that both the total capacitance and the total conductance depend on the total surface area of the wire which increases as the number of cells in the chain increase. In order to write an expression for the voltage change per small change in the longitudinal distance, we must make use of the specific capacitance (π) defined as the capacitance per unit area and the specific conductivity (π) which is defined as the conductivity per unit area. The total capacitance for the chain as a whole is then the surface area of the cylinder, 2πππ₯, times the specific capacitance and similarly for the total conductance: πΆ = 2πππ₯π, (1.46) πΊ = 2πππ₯π. (1.47) For a small change in π₯, these become πΆ = 2ππβπ₯π and πΊ = 2ππβπ₯π. Using (1.22), (1.44) and (1.45) gives: ππ2 ππ(π₯ + βπ₯, π‘) ππ(π₯, π‘) ( − )= π ππ₯ ππ₯ 2ππβπ₯ (π ππ + ππΎ (π − ππΎ ) + πππ (π − πππ ) + ππΏ (π − ππΏ )), ππ‘ π ππ(π₯ + βπ₯, π‘) ππ(π₯, π‘) ( − )= 2πβπ₯ ππ₯ ππ₯ π (1.48) ππ + ππΎ (π − ππΎ ) + πππ (π − πππ ) + ππΏ (π − ππΏ ). ππ‘ Taking the limit βπ₯ → 0 and rearranging yields: 16 (1.49) ππ π π 2 π(π₯, π‘) π = − ππΎ (π − ππΎ ) − πππ (π − πππ ) − ππΏ (π − ππΏ ). ππ‘ 2π ππ₯ 2 (1.50) The spatial derivative in equation (1.50) accounts for the longitudinal diffusion of the membrane potential from one point along the wire to another. In order to see the advantage of the action potential, let us consider for a moment that voltage-gated ion channels are not present in the membrane. Then we can group all of the non-gated ion channels into a single equivalent channel as was done for the leak current before. Taking this channel to have a constant conductivity π and a Nernst potential of zero for simplicity, equation (1.50) would then become: ππ π π 2 π(π₯, π‘) π = − ππ. ππ‘ 2π ππ₯ 2 (1.51) Solving for the steady state solution with the boundary conditions π = π0 at π₯ = 0, and π = 0 at π₯ = ∞ gives: 0= π 2 π(π₯) 2ππ − π, ππ₯ 2 π π = π0 π −π₯⁄√π⁄2ππ . (1.52) (1.53) At the steady state, the amplitude of the voltage decreases exponentially from the point of stimulation towards the end of the chain. However, such a decrease in the amplitude with distance can be largely avoided by the use of action potentials to propagate the signal. As each new cell along the chain fires an action potential, the next cell reaches the same voltage peak as the one before it (this is provided that each cell can acquire enough charge from the prior cell to fire its own action potential). The effect is much like dominoes in that if the first cell is provided a sufficient kick, each cell will act in turn along the line for great distances with no additional energy input. The action potential then acts as a signal amplifier for the electric wave as it travels along a chain of cells. At each cell, the peak 17 voltage is renewed as an action potential fires, and the electric signal can therefore arrive undiminished at its final destination even if it is initiated at a distant location. The General Form of an Equation Describing an Action Potential In the years that followed the introduction of the Hodgkin and Huxley model, several other models describing action potentials made an appearance. These models varied greatly in both their complexity and purpose, but all of the models shared the same general form. In order to obtain this more general form, the expression for the membrane potential in the Hodgkin and Huxley model can be derived using a different starting point: the conservation of electric charge. Consider an extended linear chain of cells. If a small portion of these cells is electrically stimulated by an external current source, the membrane potential in this region rises to a voltage π as a charge π builds up on the surface of the membrane. The capacitance per unit area in this region can be described by: π (1.54) . π Since the potential is higher in this region of the membrane than in the surrounding regions, π= the charge π that has built up in the vicinity of the stimulus will begin to drift longitudinally down the chain towards the region of lower potential. The law of conservation of electric charge states that the decrease in the amount of charge in the small region of the membrane near the stimulus must equal the increase in the amount of charge that has left the region and is now drifting down the chain. This law can be represented mathematically by a continuity equation: 18 ππ(π₯, π‘) + ∇π½(π₯, π‘) = π. ππ‘ (1.55) In (1.55), π is interpreted as the amount of charge within the small region of the membrane, π½ is the flux of the charge leaving or entering the boundary of the small region, and π is a term representing any additional sources or sinks of electric charge in the small region such as that provided by an external stimulus or by ion channels. Instead of working with charge, it is also possible to write an equation for the continuity of the electric potential π by using (1.54) and substituting for π in (1.55): ππ = π, (1.56) ππ(π₯, π‘) + ∇π½(π₯, π‘) = π. ππ‘ π (1.57) The flux π½(π₯, π‘) in (1.57) can now be interpreted as a potential flux, and it can be described using Fick’s law: π½(π₯, π‘) = −π· ππ(π₯, π‘) . ππ₯ (1.58) Substituting this expression into (1.57) gives: π ππ(π₯, π‘) ∂ ππ(π₯, π‘) −π· ( ) = π, ππ‘ ∂π₯ ππ₯ (1.59) (1.60) ππ(π₯, π‘) ∂2 π(π₯, π‘) =π· + π. 2 ππ‘ ∂π₯ If we were to switch off the external stimulus at some point, and also assume for the π moment that the membrane does not contain any ion channels, π can be taken to be zero, and (1.60) reduces to: π ππ(π₯, π‘) ∂2 π(π₯, π‘) =π· . ππ‘ ∂π₯ 2 (1.61) Equation (1.61) is equivalent in form to the well-known heat equation. It describes the spread of voltage with position and time. 19 On the other hand, when an external current or ion channels are added to the plasma membrane they can be treated as a source or sink of charge as these sources allow for charged ions to move across the plasma membrane in a direction perpendicular to the direction of the longitudinal diffusion. In such a case, the main equation in the Hodgkin and Huxley model can be recovered if the source term π in equation (1.60) is taken to be: π = −ππΎ (π − ππΎ ) − πππ (π − πππ ) − ππΏ (π − ππΏ ) + πΌ. (1.62) A Simplified Model of the Action Potential: The Fitzhugh-Nagumo Model Equation (1.60) may be taken as the general form for an action potential model. Other expressions for the source term π may be introduced into equation (1.60) to create a new model. However, not just any function π that is inserted into (1.60) can generate an action potential; successful functions must possess at least some properties that are similar to the Hodgkin and Huxley model in order to reproduce the same basic behavior. In the years that followed the introduction of the Hodgkin and Huxley model, the focus shifted to finding simpler and more mathematically tractable equations to work with that could describe an action potential without having the complexity of a 4-variable system. It was Richard Fitzhugh that first succeeded in creating a model that only required two variables and could still generate an action potential [2,7,9]. The original motivation behind the reduction was to make the Hodgkin and Huxley model amenable to phase space analysis on a two-dimensional plane. Fitzhugh noticed that the four variables in the Hodgkin and Huxley model can be paired into two groups. The π and π variables were considered to be “fast variables” in that they were the dominant terms during the initial 20 membrane excitation. On the other hand, the π and β variables were considered to be “slow variables” as they only appreciably changed in value after the excitation process was well underway and the system was just beginning to return to rest [7]. Fitzhugh first studied the behavior of the fast variables for a single cell by holding π and β constant, as π0 and β0 , while studying the time dependent changes in π and π. This reduced the Hodgkin and Huxley system to: π ππ = −πΜ πΎ π04 (π − ππΎ ) − πΜ ππ π3 β0 (π − πππ ) − πΜ πΏ (π − ππΏ ), ππ‘ ππ π∞ − π = . ππ‘ ππ (1.63) (1.64) The nullclines for such a system can be immediately found by setting ππ/ππ‘ and ππ/ππ‘ equal to zero and solving for π and π. With these in hand, and by using the appropriate expressions and numerical constants from the Hodgkin and Huxley model, Fitzhugh was able to obtain the fast phase plane shown in Figure 2. In Figure 2, the π and π nullclines intersect at three points labeled A, B, and C. Points A and C are stable fixed points that correspond to the resting state and the excited state respectively. Point B is on a stable manifold that acts as a threshold which must be overcome in order for the cell to become excited. If the initial condition is such that the initial value of the potential does not rise above this threshold, the trajectory will decay down to the resting state at A. If the initial condition for the potential is above the stable manifold then the trajectory will proceed to the excited state at C. A few sample trajectories are provided in Figure 2 to demonstrate this behavior. Although the fast phase plane was useful for visualizing the threshold behavior for exciting a cell, without the time dependence of the recovery variables a full action potential could 21 not be visualized. Systems that were excited would stay excited at point C and never return to rest as required for a true action potential. To visualize a full action potential on a twodimensional phase plane, it became necessary to work with a system that contained one fast variable and one slow variable while holding the other variables constant. To do this, Fitzhugh first treated π as if it was always in an instantaneous equilibrium so that π = π∞ . This reduced the number of the fast variables to one. Figure 2. The fast phase plane of the reduced Hodgkin and Huxley system as plotted by Fitzhugh [9]. The V and m nullclines are visualized as dashed lines while a few sample trajectories are shown with solid lines. The stable manifold at B acts as a threshold for the excitation of the cell. 22 In order to reduce the two slow variables down to one, Fitzhugh exploited a near symmetry in the gating variables π and β. From Figure 3 it appears that β + π ≈ 0.8, and β can therefore be eliminated by replacing it with β = 0.8 − π. A reduced Hodgkin and Huxley system for the variables π and π can then be written as: π ππ 3 (0.8 = −πΜ πΎ π4 (π − ππΎ ) − πΜ ππ π∞ − π)(π − πππ ) − πΜ πΏ (π − ππΏ ), ππ‘ ππ π∞ − π = . ππ‘ ππ (1.65) (1.66) Figure 3. Plots of the time course of the various gating variables during an action potential [7]. 23 The nullclines for this system can be obtained, and the fast-slow phase plane can be plotted as in Figure 4. In the fast-slow phase plane, the nullclines intersect at only one stable fixed point which corresponds to the resting potential. Initial conditions close to this point tend to return to the resting potential directly. However, when the initial condition for the potential is sufficiently large, something else happens entirely. Instead of returning to the stable fixed point directly, the trajectory first moves out to larger and larger values of the potential before cycling back around and returning to the fixed point as shown in Figure 4A. This detour on the phase plane corresponds to the rise and fall of an action potential. Figure 4. A: The fast-slow phase plane for the Hodgkin and Huxley model. The two nullclines are shown together with a trajectory for an action potential. B: A plot of the time course of the action potential that is shown in A [7]. Now, if the shapes of the nullclines in Figure 4 are examined closely, it would appear that the π nullcline that represents the excitation is a simple cubic shape, and the π nullcline that represents the recovery is approximately linear for much of the range of π. Using what he termed the Bonhoeffer-van der Pol (BVP) model, Fitzhugh was able to obtain a much 24 simpler set of equations that captured these phase plane features and could generate an action potential [2]. This set of equations eventually became known as the FitzhughNagumo model [7]: π ππ£ = π(π£) − π€ + πΌ, ππ‘ ππ€ = π£ − πΎπ€. ππ‘ (1.67) (1.68) In this model, π(π£) is taken to be the cubic function: π(π£) = π£(1 − π£)(π£ − πΌ), πππ 0 < πΌ < 1, π βͺ 1. (1.69) In this system the excitation variable is π£, the relaxation variable is π€, and πΌ corresponds to an externally applied current. The rest of the parameters are constants, and typical values as given in Keener and Sneyd’s Mathematical Physiology are reproduced in Table 2. If the nullclines are obtained for this system, it is seen that the π£ nullcline is cubic, and the π€ nullcline is a linear function of π£ in agreement with the approximate form of the nullclines in the Hodgkin and Huxley fast-slow phase plane. Parameter Value πΌ 0.1 πΎ 0.5 π 0.01 Table 2. Typical values for the parameters used in the Fitzhugh-Nagumo equations [7]. 25 A few years after Fitzhugh’s original discovery, a team of Japanese scientists led by J. Nagumo built the equivalent circuit for Fitzhugh’s set of equations [3]. This circuit is diagrammed in Figure 5. The three different parallel paths represent the various currents. The fast excitation current is represented by πΉ(π) on the left path where Nagumo used a tunnel diode to capture the non-linear behavior of πΉ(π). The middle path represents the recovery current where a battery, a resistor, and an inductor are connected in series with one another. The far right path is the capacitor representing the potential across the plasma membrane. Figure 5. The equivalent circuit for the Fitzhugh-Nagumo system [7]. Both of Fitzhugh’s equations can be derived directly from the equivalent circuit approach [7]. To see this, Kirchhoff’s loop rules are used to write expressions for the total current and the voltage drop across the middle path: 26 ππ + πΉ(π) + πΌπ = πΌ0 , ππ‘ ππΌπ πΏ + π πΌπ + π0 = π. ππ‘ π (1.70) (1.71) In (1.70) and (1.71), π represents the potential difference across the membrane given by π = ππ − ππ where ππ is the potential inside the membrane and ππ is the potential outside the membrane. πΌπ represents the current that flows across the middle path that contains a battery with resting potential π0, an inductor πΏ, and a resistor π while πΌ0 is an externally applied current. Now, if πΉ(π) is a cubic function in the form assumed by Fitzhugh in equation (1.69), it should have three fixed points. Two of them are stable at π = 0 and π = 1 and correspond to the rest state π0 and excited state π1 respectively while the middle fixed point represents the value of the threshold for excitation at π = πΌ. Also, a passive resistance for the tunnel diode can be taken as π 1 which is defined as π 1 = − 1⁄πΉ′(0) where πΉ′(0) is the derivative of πΉ(π) evaluated at π = 0. Equations (1.70) and (1.71) can now be expressed using dimensionless variables by making the substitutions: π£ = π ⁄π1, π€ = π 1 πΌπ ⁄π1, π(π£) = − π 1 πΉ(π1 π£)⁄π1 , and π = π 1 π‘⁄πΏ. Equation (1.70) becomes: ππ1 π 1 ππ£ π1 π1 − π(π£) + π€ = πΌ0 , πΏ ππ π 1 π 1 (1.72) π1 ππ 1 2 ππ£ ( − π(π£) + π€) = πΌ0 , π 1 πΏ ππ (1.73) π ππ£ = π(π£) − π€ + πΌ. ππ (1.74) Equation (1.74) is the same as equation (1.67) with π = π 12 π⁄πΏ and πΌ = πΌ0 π 1 ⁄π1. Making similar substitutions in equation (1.71) gives: 27 π1 π 1 πΏ ππ€ π π1 + π€ + π0 = π£π1 , π 1 πΏ ππ π 1 (1.75) ππ€ π π0 + π€ + = π£, ππ π 1 π1 (1.76) ππ€ π0 = π£ − πΎπ€ − . ππ π1 (1.77) If πΎ = π ⁄π 1 , and the resting potential is given as π0 = 0, equation (1.77) reduces to equation (1.68) as desired [7]. The Fitzhugh-Nagumo model was a major step forward in the mathematical analysis of excitation waves, and it will be used as the foundation for the analysis in the sections that follow. 28 2. ANALYTICAL APPROACHES TO STUDYING THE EXCITATION THRESHOLD Studying Threshold Phenomena for a Linear Chain of Cells: The McKean and Moll Approach The threshold phenomena described by Fitzhugh is of practical interest for cardiologists. For example, when defibrillators are employed to try and revive a patient from sudden cardiac arrest there are several questions that a doctor or a medical engineer might have which relate to threshold phenomena. What minimum value of the potential is required to propagate an electric signal that starts at one end of a tissue and travels across it? When an external current of a given magnitude is applied, what minimum number of cells must be excited in order to propagate the signal across the entire tissue? To begin to address such questions, the description of the threshold must be extended from one cell to many connected cells. One of the first attempts at understanding the spatial characteristics of threshold phenomena for a linear chain of cells was carried out by H.P. McKean and V. Moll [4]. They started with the following form of the Fitzhugh-Nagumo system: ππ£(π₯, π‘) ∂2 π£(π₯, π‘) = + π£(1 − π£)(π£ − πΌ) + π¦, ππ‘ ∂π₯ 2 ππ¦ = π½π£ − πΎπ¦. ππ‘ (2.1) (2.2) The solutions to equation (2.1) are propagating waveforms. In order to study such waveforms analytically, two major simplifications were made to these equations. First, the recovery of the system was neglected by setting π¦ = 0, π½ = 0, and πΎ = 0. The effect of neglecting recovery on a propagating waveform is demonstrated in Figure 6. 29 Figure 6 shows that when the recovery variables are present, a current pulse above Figure 6. A: When recovery is neglected, a propagating wave “front” rises to a set value as the wave propagates to the right at speed c. B: When recovery is present, a propagating wave “pulse” travels to the right [7]. threshold will cause cells to excite and then recover back down to rest in sequence as the wave pulse propagates down the chain at speed c. When recovery is removed, a wave front propagates down the chain where cells rise to a stable excited state and stay there. The general expression for a wave front travelling down a linear chain of cells is then: (2.3) ππ£(π₯, π‘) ∂2 π£(π₯, π‘) = + π(π£). 2 ππ‘ ∂π₯ Such an equation that describes wavefronts is referred to as the bistable equation [7]. This is because there are two stable fixed points at π£ = 0 and π£ = 1 in this system. Solutions that exceed the threshold requirement form an expanding wavefront where the cells rise up to π£ = 1 and stay there. On the other hand, solutions that do not meet the threshold requirement eventually return to rest at π£ = 0. For other systems that do contain recovery, 30 the only stable fixed point is at π£ = 0 since each cell will eventually return back to the resting state whether it had been excited at one point or not. The second major simplification was made to the excitation term π(π£). In the original Fitzhugh-Nagumo system this term was the cubic polynomial shown in Figure 7. Figure 7. The cubic function used for excitation in the Fitzhugh-Nagumo system [4]. McKean and Moll suggested that this excitation term could be replaced by a piecewiselinear function whose shape was similar to the cubic [4]. This function is plotted in Figure 8. The piecewise-linear approach introduced a Heaviside function that represented the transition from zero to one that took place at πΌ. This system was expressed as: ππ£(π₯, π‘) ∂2 π£(π₯, π‘) = − π£ + Θ(π£ − πΌ). ππ‘ ∂π₯ 2 (2.4) In equation (2.4) the Heaviside term is Θ(π£ − πΌ), and it has a value of 1 if π£ > πΌ and 0 if π£ < πΌ. Despite some mathematical difficulties that come with the discontinuity in the Heaviside term, this model was considered to be more mathematically tractable for the 31 analysis of waveforms. Figure 8. The piecewise-linear function used by McKean [4]. Next, McKean and Moll specified boundary conditions that determined the initial shape, or voltage profile, of the wave. In particular they chose a symmetric pulse shape where π£(π₯, π‘) went to zero at ±∞ while the peak at π₯ = 0 was a maximum such that π£ ′ (0) = 0. There were two points on such a graph where π£(π₯, π‘) could cross πΌ and these two points were labeled π(π‘) corresponding to the median of π£(π₯, π‘). In addition, the symmetry allowed them to study the time development of the wave in the positive domain only (0 < π₯ < ∞) which further simplified the mathematics. Such a waveform is shown in Figure 9 below. Studying the time dependence of the median allowed McKean and Moll to classify the various possible waves based on the asymptotic behavior of the solutions as π‘ → ∞. In addition to solutions where the waveform collapsed to zero or expanded to one as described 32 previously, there was a unique solution (provided that πΌ < 1/2) where the value of the median remained fixed over time. This solution was a standing wave, and it acted as a threshold between the resting and excited states. Figure 9. The initial waveform studied by McKean and Moll [4]. An analytical expression for the standing wave π€(π₯) was obtained by setting ππ£(π₯, π‘)⁄ππ‘ = 0 and solving the resulting differential equation by matching the boundary conditions at the median: π₯ = ±π. The result was: (2.5) ππ π₯ + π πππ π₯ ≤ −π −π πππ π₯ < π }, π€(π₯) = {1 − π cosh π₯ π₯−π ππ πππ π₯ ≥ π 1 (2.6) π = (1 − π −2π ). 2 In order to classify the various standing wave solutions based on the initial conditions, a parameter was introduced that was proportional to the amplitude of the pulse. Given some arbitrary initial condition for the pulse amplitude, as π‘ → ∞ the solution would collapse to zero if the amplitude was less than πΌ, and it would expand to one if the amplitude was greater than πΌ. In between where the solution neither collapsed nor expanded, the parameter was termed the critical multiplier, and the critical value of the pulse amplitude 33 that corresponded to it appeared on a threshold surface of codimension 1 which represented the standing wave solution. This threshold surface is shown in Figure 10. Figure 10. The threshold surface visualized by McKean and Moll. Initial conditions below the critical surface tended towards zero, those on the critical surface tended towards a saddle point, and those above the surface tended towards one [4]. While the existence of the critical surface was established analytically by McKean and Moll, techniques to calculate the critical multiplier were introduced several years later by Moll and Rosencrans [10]. First, Moll and Rosencrans used standard numerical methods to solve the partial differential equation (see Appendix: Numerical Methods). The critical multiplier could be found by varying the initial condition for the pulse amplitude. They would start at some value of the multiplier where the amplitude was below threshold so that the resulting solution would decay to zero in time. They would then increase the value 34 of the amplitude by some small numerical amount and solve the system again to see what happened. At some point, at a certain critical value of the pulse amplitude, the solution would expand to one indicating that the threshold had been crossed. This value was then taken to be the critical amplitude or critical height. Figure 11. Numerical plots of the critical amplitude (height) vs. the pulse width (base). The result for the McKean and Moll system is on the left while the result for the FitzhughNagumo system appears on the right [10]. The process of finding the critical amplitude was repeated for various initial pulse widths which were referred to as the “base” of the waveform depicted in Figure 9. The results for both the McKean and Moll and Fitzhugh-Nagumo systems were plotted in Figure 11. It was noted that, for sufficiently large pulse widths, the threshold requirement in either approach was simply that the critical amplitude needed to exceed the value of πΌ in order to excite either system. The threshold was found to increase substantially in the small pulse width limit. 35 Moll and Rosencrans also developed an analytical method to calculate the critical multiplier for the McKean approach. Unfortunately, their solution proved to be very nonlinear and complex. To evaluate it, numerical methods had to be employed. The characteristics of the threshold surface had been established in Figure 11, but an analytical description of how these curves came about was still missing. Specifically, the rise in the requirement for the critical amplitude at small widths had not been accounted for in the earlier analysis by McKean and Moll. Their description of the threshold requirement only held in the limit of large pulse widths. Also, the scope of their analysis was limited to simply a discussion of the amplitude vs. the pulse width; the effect of conductance terms on the critical amplitude was never addressed. Moving in the Direction of an Analytical Theory for the Threshold Surface: Projected Dynamics The next step was to provide an analytical theory that could reproduce the threshold curves obtained by Moll and Rosencrans that were shown previously in Figure 11. Moll himself developed a polygonal approximation method that could reproduce curves resembling the critical surface [11]. It did not yield a concise analytical formula for the threshold however. Another approach to this problem was carried out a few years later by J. Neu, R.S. Preissig, Jr., and W. Krassowska [5]. They called their method “projected dynamics”, and it plays such a central role in what follows that their approach will be described below in great detail. The starting point is the Fitzhugh-Nagumo model which originally will be taken to be: ππ£ ∂2 π£ π = π· 2 − ππ ′ (π£) − π¦, ππ‘ ∂π₯ 36 (2.7) ππ¦ = π½π£ − πΎπ¦, ππ‘ (2.8) 1 πππ 0 < πΌ < . 2 π ′ (π£) = π£(πΌ − π£)(1 − π£) (2.9) In equations (2.7)-(2.9) the excitation variable is π£, the source term is π ′ (π£) (which when written this way refers to the derivative of some function π(π£) with respect to π£), and the inactivation variable is π¦. π, π·, π, π½, πΎ, and πΌ are taken to be constants. The symbols π, π·, and π are chosen to be in analogy with the coefficients that modify the corresponding terms in the Hodgkin and Huxley equations. This analogy should not be taken too literally as it must be remembered that these are not exactly the same quantities in the Fitzhugh system which uses non-dimensional variables. However, owing to their positions in the equation, π acts as an effective capacitance term, π· and π act as effective conductance terms for the longitudinal and ion channel currents respectively. The constants π½ and πΎ control the recovery process, and the threshold parameter is πΌ. πΌ is limited to the range of values between zero and one half in order to permit a standing wave solution for the threshold as required by McKean and Moll [4]. To simplify the analysis, the excitation of wavefronts can be studied by setting π¦, π½, and πΎ all equal to zero which removes the recovery process from the system. Since the solution we seek is time independent we can also set π = 1 (or alternatively we can think of this as dividing by π and absorbing it into the constants π· and π). Unlike in Neu’s original analysis, the constants π· and π will be kept as free parameters throughout in order to examine their effects on the threshold at a later point. Taking all of this into account, we have reduced the original system of equations to: 37 ππ£ ∂2 π£ = π· 2 − ππ ′ (π£), ππ‘ ∂π₯ (2.10) 1 πππ 0 < πΌ < . 2 π ′ (π£) = π£(πΌ − π£)(1 − π£) (2.11) For equation (2.10), we can write a functional for the total energy πΈ: ∞ (2.12) π· ππ£ 2 πΈ = ∫ ( ( ) + ππ(π£)) ππ₯. 2 ππ₯ −∞ Given (2.12), we can rewrite (2.10) as the gradient flow of the energy: ππ£ πΏπΈ =− . ππ‘ πΏπ£ (2.13) The Formal Derivation of Projected Dynamics To prove (2.13), we may perturb the potential in (2.12) by some small amount π(π₯, π‘) to get: 2 π· π(π£(π₯, π‘)) π(π(π₯, π‘)) ( + ) + πΈ(π£(π₯, π‘) + π(π₯, π‘)) = ∫ ( 2 ). ππ₯ ππ₯ ∞ (2.14) ππ(π£(π₯, π‘) + π(π₯, π‘))ππ₯ −∞ The first term in the integral can be expanded and simplified by taking the product of the derivative of the infinitesimal term, π(π₯, π‘), to be zero: 2 π(π£(π₯, π‘)) ( ) + ππ₯ 2 π· π(π£(π₯, π‘)) π(π(π₯, π‘)) π· ( + ) ≅ 2 ππ₯ ππ₯ 2 (2.15) . π(π£(π₯, π‘)) π(π(π₯, π‘)) ππ₯ ππ₯ ( ) The second term can be Taylor expanded where again all terms containing the products of 2 infinitesimals are discarded: 38 ππ(π£(π₯, π‘) + π(π₯, π‘)) ≅ π (π(π£(π₯, π‘)) + π ′ (π£(π₯, π‘))π(π₯, π‘)). (2.16) Inserting (2.15) and (2.16) into (2.14) gives: πΈ(π£(π₯, π‘) + π(π₯, π‘)) (2.17) ∞ 2 π(π£(π₯, π‘)) π(π(π₯, π‘)) π· π(π£(π₯, π‘)) = ∫ ( (( ) + 2 ) 2 ππ₯ ππ₯ ππ₯ −∞ + π(π(π£(π₯, π‘)) + π′(π£(π₯, π‘))π(π₯, π‘))) ππ₯. Now, to get the change in the energy, we subtract (2.12) from (2.17): βπΈ = πΈ(π£(π₯, π‘) + π(π₯, π‘)) − πΈ(π£(π₯, π‘)). (2.18) Cancelling terms, we are left with: ∞ π(π£(π₯, π‘)) π(π(π₯, π‘)) βπΈ = ∫ (π· + ππ′(π£(π₯, π‘))π(π₯, π‘)) ππ₯. ππ₯ ππ₯ (2.19) −∞ Integration by parts can be done on the first term in the integral, and the boundary term vanishes to yield: ∞ π 2 (π£(π₯, π‘)) βπΈ = ∫ − (π· − ππ′(π£(π₯, π‘))) π(π₯, π‘)ππ₯. ππ₯ 2 (2.20) −∞ The term in parenthesis is the same as ππ£(π₯, π‘)/ππ‘ as given by (2.10), and when the integral in (2.20) is equal to an extremum this term is also equal to the variational derivative so that: ππ£ πΏπΈ =− . ππ‘ πΏπ£ 39 (2.21) To tackle equation (2.21), Neu chose a parametric representation for the potential π£(π₯, π‘) where it was represented as a vector of linearly independent time-dependent parameters, π(π‘): π£(π₯, π‘) = π(π(π‘), π₯). (2.22) The motivation for this approach is that the amplitude and width can now be set as independent parameters, and the solutions for each can be obtained from (2.21). Using (2.22) to rewrite the left side of (2.21): π ππ£ ππ πππ ππ =∑ = πΜ . ππ‘ πππ ππ‘ πππ π (2.23) π=1 The summation is carried out for 1,2,...π parameters, ππΜ represents the time derivative of the ππ‘β parameter, and the summation notation has been replaced with repeated index notation in the far right representation. Repeated index notation will be used from this point forward. This makes (2.21): ππ πΏπΈ ππΜ = − . πππ πΏπ (2.24) To extract the ordinary differential equations for the various ππΜ , we multiply both sides by ππ⁄πππ and take the functional inner product: ∞ ∞ −∞ −∞ ππ ππ πΏπΈ ππ (∫ ππ₯) ππΜ = − ∫ ππ₯. πππ πππ πΏπ πππ (2.25) Note that ππΜ is independent of π₯ and can be taken out of the integral on the left side. Now the claim is made that: 40 ∞ (2.26) ππΈ πΏπΈ ππ = ∫ ππ₯. πππ πΏπ πππ −∞ The steps to show this are very similar to those that were used to show (2.13). To start with, we’ll express the energy as a function of the parametric representation of the potential: ∞ (2.27) 2 π· π(π(π(π‘), π₯) πΈ(π(π(π‘), π₯)) = ∫ ( ( ) + ππ(π(π(π‘), π₯))) ππ₯ . 2 ππ₯ −∞ The time-dependent parameters are perturbed by a small amount βπ, and the potential is Taylor expanded to the first order in both terms to yield: πΈ(π(π(π‘) + βπ(π‘), π₯)) = ∞ ∫ π· ππ(π(π‘), π₯) π ππ(π(π‘), π₯) ( + ( ) βπ(π‘)) 2 ππ₯ ππ₯ ππ(π‘) (2.28) 2 ππ(π(π‘), π₯) −∞ + (π(π(π(π‘), π₯)) + π′(π(π(π‘), π₯)) βπ(π‘)) ππ(π‘) ( ) ππ₯ . Expanding the first term in the integral and discarding the product of the infinitesimals gives: πΈ(π(π(π‘) + βπ, π₯)) = (2.29) 2 π· ππ(π(π‘), π₯) ππ(π(π‘), π₯) π ππ(π(π‘), π₯) (( ) + 2 ( ) βπ(π‘)) 2 ππ₯ ππ₯ ππ₯ ππ(π‘) ∞ ∫ −∞ ( + π (π(π(π(π‘), π₯)) + π′(π(π(π‘), π₯)) ππ(π(π‘), π₯) βπ(π‘)) ππ(π‘) ππ₯. ) The difference in the energy is obtained by subtracting (2.27) from (2.29): βπΈ = πΈ(π(π(π‘) + βπ(π‘), π₯)) − πΈ(π(π(π‘), π₯)). After cancelling terms and simplifying, we are left with: 41 (2.30) (2.31) π(π(π(π‘), π₯)) π ππ(π(π‘), π₯) ( )+ ππ₯ ππ₯ ππ(π‘) βπΈ = ∫ βπ(π‘)ππ₯. ππ(π(π‘), π₯) −∞ ππ′(π(π(π‘), π₯)) ππ(π‘) ( ) Performing an integration by parts on the first term in the integral and discarding the ∞ π· boundary term yields: ∞ π 2 (π(π(π‘), π₯)) ππ(π(π‘), π₯) βπΈ = ∫ − (π· − ππ′(π(π(π‘), π₯))) βπ(π‘)ππ₯, 2 ππ₯ ππ(π‘) (2.32) −∞ and from here the desired result follows readily: π 2 (π(π(π‘), π₯)) ππ(π(π‘), π₯) πΏπΈ =− =π· − ππ′(π(π(π‘), π₯)), ππ‘ πΏπ ππ₯ 2 ∞ βπΈ πΏπΈ ππ(π(π‘), π₯) = ∫ − (− ) ππ₯ , βπ(π‘) πΏπ ππ(π‘) (2.33) (2.34) −∞ ∞ ππΈ πΏπΈ ππ = ∫ ππ₯. πππ πΏπ πππ (2.35) −∞ With (2.26) now established, we can substitute it into (2.25) to obtain the central equation for projected dynamics: ∞ (∫ −∞ ππ ππ ππΈ ππ₯) ππΜ = − , πππ πππ πππ (2.36) or writing this result in matrix form: ππΜ = −ππΈ. (2.37) Here π is a π x π symmetric matrix whose components are given by: ∞ ππ ππ πππ = ∫ ππ₯. πππ πππ −∞ 42 (2.38) In replacing the original partial differential equation with equation (2.37), the difficult problem of solving the partial differential equation has been reduced to finding the motion for specified parameters in a finite dimensional projected phase space. By analyzing the dynamics of the parameters in phase space, we can acquire information about which conditions lead to trajectories that either expand or collapse. Applying Projected Dynamics Using a Gaussian Pulse Now that the projected dynamics has been formally derived, it is time to apply the theory to the Fitzhugh-Nagumo system and reproduce the plot of the critical amplitude vs. pulse width that was obtained by Moll and Rosencrans. Step one of this procedure involves choosing a shape for the applied pulse that will serve as an initial voltage profile for equation (2.10). To simplify the calculations for the projected dynamics, Neu chose to make π£(π₯, π‘) a Gaussian pulse with time dependent parameters for the amplitude π1 (π‘) and the pulse width π2 (π‘): π(π₯, π‘) = π1 (π‘)π −π₯ 2 /(2π (π‘)2 ) 2 . (2.39) Here the pulse width is interpreted as only the width of the applied pulse to the right of its initial point of stimulation (at π₯ = 0). The reason for this is that we will be examining only the portion of the wavefront that propagates to the right down the cell chain from the initial point of stimulation (from symmetry an identical wavefront will also travel down the cell chain to the left in the negative π₯ direction). This is a different measure for the pulse width than the one originally used by Neu, but it is in agreement with the pulse width as defined in McKean and Moll. We will follow the latter convention for ease of comparison between the different approaches. 43 Forming a vector for the derivatives of the two time-dependent parameters gives us: πΜ = [ π1Μ ]. π2Μ (2.40) The goal is to solve the system: ππΜ = −ππΈ. (2.41) To start with, we will obtain the matrix π where the matrix elements are given by: ∞ πππ = ∫ −∞ ππ ππ ππ₯. πππ πππ (2.42) For the Gaussian pulse, we have: ππ 2 2 = π −π₯ /(2π2 ) , ππ1 (2.43) ππ π1 π₯ 2 −π₯ 2 /(2π 2 ) 2 . = π ππ2 π2 3 (2.44) Using standard Gaussian integration techniques, the πππ become: π11 = π2 √π, π12 = π21 = π22 = π1 √π , 2 3π1 2 √π , 4π2 (2.45) (2.46) (2.47) and the matrix π is: π2 π = √π [ π 1 2 π1 2 3π1 2 ]. 4π2 Now we want to find the energy gradient −ππΈ. The energy is given by: 44 (2.48) ∞ π· ππ 2 πΈ = ∫ ( ( ) + ππ(π)) ππ₯. 2 ππ₯ (2.49) −∞ The derivative for the left term in the integral is: ππ π1 π₯ 2 2 = − 2 π −π₯ /(2π2 ) . ππ₯ π2 (2.50) For the term on the right, Neu argued that it was sufficiently accurate to obtain a more compact solution by applying the condition that πΌ βͺ 1 in (2.11). Since πΌ is related to the median on the voltage profile, a small value of πΌ implies that the height of the voltage profile is also small (π£ βͺ 1). With this approximation, we then have (1 − π£) ~1 and π′(π£) can be expressed as a quadratic: π ′ (π£) = π£(πΌ − π£)(1 − π£) ~ πΌπ£ − π£ 2 . (2.51) Integrating (2.51) gives: (2.52) πΌ 2 π£3 π£ − . 2 3 Although this approximation eventually has a rather large effect on the waveform (instead π(π£) = of going to a value of one after an action potential, the solution blows up to infinity), it was confirmed through numerical simulation that this approach had very little effect on the threshold for excitation. This is because the threshold is reached long before the potential reaches its maximum value following an action potential. Both (2.11) and (2.51) lead to acceptable solutions for projected dynamics, but the solution using (2.52) is more compact than using the full cubic, and so we will adopt this approximation as well. Substitution of (2.50) and (2.52) into (2.49) and carrying out the integration yields: 45 π·π1 2 πΌπ1 2 π2 2 π1 3 πΈ = √π ( +π( −√ π )), 4π2 2 3 3 2 (2.53) and then the energy gradient becomes (in matrix form): (2.54) π·π1 2 + π (πΌπ1 π2 − √ π1 2 π2 ) 2π2 3 −ππΈ = −√π . [ π·π1 2 πΌπ1 2 2 π1 3 √ − +π( − ) 4π2 2 2 3 3 ] Using the inverse matrix for π, the system can finally be solved for πΜ : πΜ = −π−1 ππΈ, − πΜ = [ π1Μ ]= π2Μ [ (2.55) π·π1 7 − π (πΌπ1 − π1 2 ) 2 π2 3√6 π· 2 − π√ π1 π2 π2 27 (2.56) . ] Taking the amplitude to be π1 and the pulse width to be π2 , we can plot the phase portrait for π1 vs. π2 using the solutions from equation (2.56). The phase portrait for the projected dynamics is given in Figure 12 below with π· = π = 1.0. Several representative trajectories and the direction fields are shown to indicate the time evolution for various initial conditions. From the trajectories, it can be seen that the unstable manifold near the center of the plane represents the threshold for excitation; initial conditions below this curve decay to zero in time (as can be seen by following the vector field on the phase plane for any trajectory starting below the unstable manifold) while those trajectories starting above the unstable manifold rise to a greater amplitude. The nullclines for the phase plane can be obtained in the usual way: The first nullcline is plotted in Figure 12 in yellow by setting π1Μ = 0 and solving for π1 as a function of π2 . The second nullcline is plotted in red by 46 setting π2Μ = 0 and solving for π2 as a function of π1 . Figure 12. The phase plane for the projected dynamics with a Gaussian pulse. The amplitude nullcine is yellow, the pulse width nullcline is red, the unstable manifold represents the excitation threshold separatrix, and the stable manifold forms a second separatrix. The green arrows give the direction of the vector field while sample trajectories are in blue. This phase portrait was created using PPLANE2005.10 by John Polking [12]. 47 Analysis of the Projected Dynamics Using a Gaussian Pulse Now that we have obtained the phase plane for the projected dynamics in Figure 12, our focus shifts to the task of trying to understand this result as completely as possible. From examining the general features of the phase plane, we can see that the approach that Neu took does indeed reproduce key features reported previously by McKean and Moll in Figure 10 and Moll and Rosencrans in Figure 11. Figures 10-12 are collected in Figure 13 for a side by side comparison. Figure 13. Comparison of Figures 10-12. On the left is the depiction of the threshold surface given by McKean and Moll, in the center is the numerical threshold surface provided by Moll and Rosencrans, and on the right is the projected dynamics. Upon observing the phase plane for the projected dynamics in Figure 12, it appears that the nullcline for the amplitude is in good qualitative agreement with the actual threshold curve which is the unstable manifold. This is not by coincidence. Regardless of the initial condition or the time course for the pulse width, if an initial condition for the amplitude is above the amplitude nullcline it is guaranteed to fire an action potential. The reason for this is that the nullcline represents a change in sign for the rate of change of the amplitude. 48 Above the nullcline, the amplitude is always changing in a positive direction, and so the amplitude only climbs upward in this region. At small pulse widths, there is a region of initial conditions where the nullcline doesn’t agree with the actual threshold curve. This represents the set of initial conditions where the amplitude initially decreases before eventually turning upward and firing an action potential. Taking this into account, the amplitude nullcline represents a sort of upper bound estimation of the threshold. It overestimates the requirement placed on the amplitude for very small pulse widths, but approximates the requirement well for pulses with a sufficiently large initial pulse width. Owing to this similarity between the amplitude nullcline and the actual threshold curve, the equation for the nullcline may be used to extract an approximate analytical expression for the threshold curve. To do this, we start with the solution for π1Μ from (2.56). Since the initial conditions play the dominant role when it comes to excitation, we may set the pulse width π2 to a constant: the initial pulse width π€. Then we set π1Μ = 0 and solve for the timeindependent critical amplitude π1 → ππ . Doing so, we have: ππ = 3√6 π· ( + πΌ). 7 ππ€ 2 (2.57) To see how this result compares with the true threshold curve, equation (2.10) was solved for the threshold numerically (see Appendix: Numerical Methods) for a linear chain of 50 cells with π· = π = 1 and πΌ = 0.139. This result was plotted in Figure 14 together with Equation (2.57). It is seen in Figure 14 that (2.57) agrees well with the numerical result for the true threshold. Aside from the two-variable projected dynamics that yielded (2.57), it is also possible to do a one-variable projected dynamics where the width is set to a constant π€ at the start of the procedure, and the only free parameter is π1 . In a later section, the 49 derivation of the one-variable projected dynamics will be carried out explicitly. In the meantime, the amplitude nullcline for the one-variable projected dynamics is provided in Figure 14 to compare the two approaches for the case of the Gaussian pulse. Figure 14. Comparison of the amplitude nullclines obtained using projected dynamics vs. the numerical threshold solution. Aside from describing how the amplitude varies with the initial pulse width, equation (2.57) also makes predictions as to how the critical amplitude should depend on the effective ion channel and longitudinal conductivities given by π and π· respectively. Interestingly, equation (2.57) predicts that increasing the conductivity for the ion channel current decreases the requirement on the amplitude to fire an action potential while 50 increasing the conductivity for the longitudinal current has the opposite effect. To see if the predictions made by (2.57) accurately describe the threshold behavior for (2.10), we first solved equation (2.10) numerically as before to find the threshold curve while varying the parameter π and using the following constant parameters: π· = 1, π€ = 5, and πΌ = 0.139. The result is plotted together with the corresponding solution to (2.57) in Figure 15 below. Figure 15. Plot of the critical amplitude (ππ ) vs. the ion channel conductivity (π) for the projected dynamics with a Gaussian pulse. The agreement between the two is quite good, especially for very large and very small values of π. On the other hand, (2.57) predicts that ππ should increase with an increase in the longitudinal conductivity π·. Once more we solved (2.10) numerically (for π· ≥ 0.1) 51 with parameters : π = 1, π€ = 5, and πΌ = 0.139, and (2.57) was also solved with these same parameters. The result is plotted in Figure 16 below. Figure 16. Plot of the critical amplitude (ππ ) vs. the longitudinal conductivity (π·) for the projected dynamics with a Gaussian pulse. The general upward trend for the threshold requirement with increasing longitudinal conductivity carries through in (2.57) when compared to the numerical simulation. However, it is also clear that the actual relationship is not so simple and linear like equation (2.57) implies. For qualitative purposes the agreement is still quite good provided that π· is not too large. 52 Discussion of the Results of the Projected Dynamics Using a Gaussian Pulse For a one-dimensional chain of cells, we can get a qualitative picture of the excitation process from the above analysis. The potential is initially raised for a certain portion of the cell chain, and this is represented by an initial amplitude and an initial width for the current pulse, and then the system is allowed to evolve in time. There are two possible outcomes for this system: either the cells fire an action potential or the cells decay to zero in time. Equation (2.57) predicts that there is a minimum amplitude that is necessary to get the chain to fire. This minimum amplitude (ππ ) depends on various parameters like the initial pulse width (π€), the ion channel conductivity (π), and the longitudinal conductivity (π·). The effect of increasing the initial pulse width is to decrease the requirement on the initial amplitude in order to get the action potential to fire. This corresponds to raising a larger portion of the available cell chain when the initial condition for the potential is applied. For example, this would simulate an electrode with a larger surface area being used to stimulate the cell chain. Likewise, increasing the ion channel conductivity also decreases the initial amplitude that is needed to fire an action potential. Since this parameter relates to the ion channel that is producing excitation in the membrane model, an increase in this conductivity would correspond to additional ion channels opening. This would lead to a faster excitation. Conversely, the effect of increasing the longitudinal conductivity is to increase the requirement on the initial amplitude. Since the longitudinal conductivity refers to a diffusion process, larger values of π· drain current out of a stimulated cell and into neighboring cells that are at rest. If the value of this parameter is too high then current may drain out of a particular cell before that cell can reach the necessary threshold for an action potential to fire. Lowering the longitudinal conductivity has the opposite effect, it keeps 53 the applied current localized for a longer time. This allows the charge to build up easier in a small region of the membrane which may assist the cell in reaching the threshold requirement. Although it provides a nice qualitative description of the threshold curve, equation (2.57) has a few shortcomings. For instance, the numerical results predict that for very large widths the threshold should be precisely πΌ. Equation (2.57) gives a result close to this, but the factor 3√6⁄7 is a little larger than one, and it overestimates the threshold by about 5%: ππ = 3√6 π· 3√6 ( 2 + πΌ) ~ πΌ ~ 1.05πΌ 7 ππ€ 7 (πππ π€ → ∞). (2.58) Also, as mentioned before, there is a noticeable discrepancy between the actual threshold curve and (2.57) for small pulse widths. Finally, the Gaussian shape of the initial voltage profile is not as physiologically accurate as the shape of the voltage profile chosen by McKean and Moll. The voltage profile they chose is closer in shape to one generated by a real electrode. Many of these issues can be improved by choosing a different shape for the initial voltage profile and reformulating the projected dynamics. This task will be undertaken next. 54 3. FORMULATION OF THE THRESHOLD EQUATION FOR A SQUARE PULSE USING INITIAL CONDITIONS A New Formulation: Changing the Initial Voltage Profile to a Square Pulse Now that the projected dynamics has been established for a one-dimensional cell chain using the Fitzhugh-Nagumo relations with a Gaussian pulse for an initial voltage profile, it is natural to try and apply the method towards systems with more realistic physiological characteristics. The most immediate thing to change would be the initial Gaussian pulse. It is unphysiological in at least two ways. First off, the shape of the Gaussian pulse is not a great fit to the more physiological square pulse which is generated by an externally applied current given by: πππ − π ⁄2 < π₯ < π ⁄2, (3.1) πΌ(π₯, π‘) = 0 πππ π₯ > π ⁄2 ππ π₯ < − π ⁄2. (3.2) πΌ(π₯, π‘) = πΌ(π‘) An example of a square current pulse is shown below in Figure 17. Figure 17. A sample plot of a square pulse. The amplitude of the applied pulse is πΌ, and the width extends from π ⁄2 to − π ⁄2. 55 The Gaussian voltage profile was chosen for its desirable mathematical properties, but presumably a more accurate solution can be obtained if we go with a shape generated by a square pulse. Nevertheless, a square pulse is a discontinuous function and will have to be approximated to some extent when modeling it analytically. Notice in Figure 17 that the full width of the square pulse is π but the width that we will use is that portion of the solution travelling to the right. This width will be given by π ⁄2 which corresponds to the median of the pulse width from the McKean and Moll treatment. The second aspect is a bit more subtle. In the original projected dynamics, Neu approximated an initial pulse from an external electrode using initial conditions for the potential specified at the initial time π‘ = 0. The significance of this idealization is that by specifying this initial pulse using initial conditions for the potential directly, one need not have to worry about introducing an external current πΌ(π₯, π‘) into the Fitzhugh-Nagumo relation. Introducing such a current would make the system that was used previously a bit more complicated: ππ£ ∂2 π£ = π· 2 − ππ ′ (π£) + πΌ(π₯, π‘), ππ‘ ∂π₯ ′ (π£) π = π£(πΌ − π£) πππ 0 < πΌ < 1⁄2 , πΌ βͺ 1. (3.3) (3.4) In contrast to the initial conditions approach, for the actual physiological case, the applied pulse πΌ(π₯, π‘) will be applied for some finite time that is typically longer than the time for an action potential to fire. Eventually, we would like to replace the initial conditions approach with an applied current πΌ(π₯, π‘), as in the general expression (3.3), so that a timedependent applied current can be used to drive the system as would be the case with a real electrode or other external current source. Moving in that general direction, we will start with (3.3) and attempt to find an acceptable shape for the voltage profile that will fit the 56 criteria for an external square pulse πΌ(π₯, π‘). If we consider that πΌ(π₯, π‘) is applied for a time that is much greater than the time needed to excite the system, then πΌ(π₯, π‘) can be treated as being constant in time and can be labeled as πΌ(π₯). In doing this, we assume that the current is still being applied while the threshold is reached. Substitution of equation (3.4) into (3.3) and expanding the source term gives: ππ£ ∂2 π£ = π· 2 − ππΌπ£ + ππ£ 2 + πΌ(π₯). ππ‘ ∂π₯ (3.5) Working in analogy to a real physiological model, we will now label the coefficients for the excitation terms as: ππΌ → ππΎ , (3.6) π → πππ , (3.7) to make (3.5): (3.8) ππ£ ∂2 π£ = π· 2 − ππΎ π£ + πππ π£ 2 + πΌ(π₯). ππ‘ ∂π₯ Here the negative term that tends to return the system to zero as the potential increases is labeled with the coefficient ππ in analogy to the potassium channels in the Hodgkin and Huxley model. Likewise, the positive term that tends to raise the voltage will be labeled with the coefficient πππ in analogy with the sodium channels. Again, caution must be taken not to take this analogy too literally here. These terms are mainly used to help us interpret our results as (3.6) and (3.7) play similar roles in (3.8) to the corresponding terms in the more complicated Hodgkin and Huxley model. In particular, it must be kept in mind that equation (3.8) is too simple to fire an action potential as written. The sodium term is not at least a cubic, which was determined to be a minimum requirement for firing an action potential by Fitzhugh, and it also lacks an 57 inactivation term to keep the solution from blowing up at high values of the potential. However, for the study of excitation we can get around this difficulty by working as we did before under the assumption that πΌ βͺ 1 which implies that the system will reach threshold long before these inaccuracies can appreciably influence the threshold result. Since π£ βͺ 1 at the time of excitation in this case, we can assume the πππ π£ 2 term is small and neglect it while we search for the shape of the threshold surface. This gives us: (3.9) ππ£ ∂2 π£ = π· 2 − ππΎ π£ + πΌ(π₯). ππ‘ ∂π₯ This equation is now similar to the linear piecewise expression used by McKean and Moll, and we can solve for the time-independent solution using familiar methods. The system is given by: 0=π· π 2 π£1 − ππΎ π£1 + πΌ ππ₯ 2 (3.10) πππ − π ⁄2 < π₯ < π ⁄2, (3.11) π 2 π£2 − ππΎ π£2 πππ π₯ > π ⁄2 ππ π₯ < − π ⁄2. 2 ππ₯ Due to the symmetry of the desired solution, we need only solve for the solutions in the 0=π· region where π₯ > 0. The solution for (3.11) is solved in the standard way for a linear second order differential equation. Making the identification: π2 π£2 0=π· − ππΎ π£2 → ππ₯ 2 π 2 π£2 − πΎ 2 π£2 = 0 ππ₯ 2 ππΎ , π· π€βπππ πΎ = √ (3.12) we assume a solution of: π£2 = πΆπ ππ₯ , (3.13) where πΆ and π are constants. Inserting (3.13) into (3.12) and simplifying yields: π = ± πΎ. (3.14) The general solution is then given by: 58 π£2 = πΆ1 π πΎπ₯ + πΆ2 π −πΎπ₯ . (3.15) To satisfy the specified requirement for the voltage profile we must have π£2 → 0 at π₯ → ∞ so that πΆ1 = 0. Also, π£2 must originate at π ⁄2 and then exponentially decay from there as π₯ increases. Taking these features into account, (3.15) reduces to: π£2 = πΆπ −πΎ(π₯−π⁄2) πππ π₯ ≥ π/2. (3.16) For the first solution, we have a non-homogeneous second order differential equation: π 2 π£1 πΌ − πΎ 2 π£1 = − . 2 ππ₯ π· (3.17) The general solution is given as the sum of the homogeneous solution π£β and a particular solution π£π : π£ = π£β + π£π , (3.18) π£β = πΈ1 π πΎπ₯ + πΈ2 π −πΎπ₯ , (3.19) π£π = π1 (π₯)π¦1 (π₯) + π2 (π₯)π¦2 (π₯), (3.20) π¦1 (π₯) = π πΎπ₯ , (3.21) π¦2 (π₯) = π −πΎπ₯ . (3.22) The solutions for π1 (π₯) and π2 (π₯) are given by [13]: π1 (π₯) = − ∫ π2 (π₯) = ∫ π¦2 (π₯)π(π₯) ππ₯, π(π¦1 , π¦2 )(π₯) π¦1 (π₯)π(π₯) ππ₯, π(π¦1 , π¦2 )(π₯) π(π₯) = − πΌ ⁄π· , ππ¦2 (π₯) ππ¦1 (π₯) − π¦2 (π₯) ππ₯ ππ₯ = π πΎπ₯ (−πΎ)π −πΎπ₯ − (πΎ)π πΎπ₯ π −πΎπ₯ = −2πΎ. π(π¦1 , π¦2 )(π₯) = π¦1 (π₯) (3.23) (3.24) (3.25) (3.26) Inserting (3.21), (3.22), (3.25), and (3.26) into (3.23) and (3.24), and performing the integration gives: 59 π1 (π₯) = πΌ π −πΎπ₯ , 2π·πΎ 2 (3.30) π2 (π₯) = πΌ π πΎπ₯ . 2π·πΎ 2 (3.31) Placing (3.21), (3.22), (3.30), and (3.31) in (3.20), we obtain the particular solution π£π : π£π = πΌ πΌ πΌ πΌ −πΎπ₯ πΎπ₯ πΎπ₯ −πΎπ₯ π π + π π = = . 2π·πΎ 2 2π·πΎ 2 π·πΎ 2 ππΎ (3.32) The general solution for equation (3.18) becomes: π£1 = π£β + π£π = πΈ1 π πΎπ₯ + πΈ2 π −πΎπ₯ + πΌ . ππΎ (3.33) We can simplify (3.33) by noting that the maximum of the curve must appear at π₯ = 0, and the derivative at this point must be continuous: ππ£1 (0) = 0, ππ₯ (3.34) ππ£1 (0) = πΎπΈ1 − πΎπΈ2 = 0 → πΈ1 = πΈ2 . ππ₯ (3.35) This makes (3.33): π£1 = πΈ(π πΎπ₯ + π −πΎπ₯ ) + πΌ . ππΎ (3.36) Now to link the solutions for π£1 and π£2 together, we use boundary conditions that require continuity at π₯ = π/2 of the two solutions and their first derivatives: π£1 = π£2 ππ‘ π₯ = π ⁄2, (3.37) ππ£1 ππ£2 = ππ₯ ππ₯ ππ‘ π₯ = π ⁄2. (3.38) πΌ = πΆ, ππΎ (3.39) Inserting (3.16) and (3.36), these become: πΈ(π πΎπ/2 + π −πΎπ/2 ) + πΈ(π πΎπ/2 − π −πΎπ/2 ) = −πΆ. 60 (3.40) Adding (3.39) to (3.40) gives an expression for πΈ while subtracting the two equations gives an expression for πΆ. These are: πΈ=− πΆ= πΌ −πΎπ⁄2 π , 2ππΎ (3.41) πΌ (1 − π −πΎπ ). 2ππΎ (3.42) The solutions to the differential equation are then: (π πΎπ₯ + π −πΎπ₯ ) πΌ −πΎπ⁄2 π£1 = (1 − π ) ππΎ 2 = πΌ (1 − π −πΎπ⁄2 cosh(πΎπ₯)) ππΎ π£2 = (3.43) πππ 0 < π₯ < π ⁄2, πΌ (1 − π −πΎπ )π −πΎ(π₯−π⁄2) 2ππΎ πππ π₯ > π/2. (3.44) Now, before starting into the projected dynamics, we wish to express the potential as a vector of parameters: π£(π₯, π‘) = π(π(π‘), π₯). (3.45) Replacing πΌ/ππΎ with the time dependent amplitude π1 (π‘) and replacing π with the time dependent pulse width π2 (π‘), we obtain our voltage profile: π1 (π(π‘), π₯) = π1 (π‘)(1 − π −πΎπ2 (π‘)⁄2 cosh(πΎπ₯)) πππ 0 < π₯ < π2 (π‘)⁄2, (3.46) π1 (π‘) (3.47) (1 − π −πΎπ2 (π‘) )π −πΎ(π₯−π2 (π‘)⁄2) πππ π₯ > π2 (π‘)/2. 2 These equations should be compared with those originally provided by McKean and Moll π2 (π(π‘), π₯) = (equations (2.5) and (2.6)). The basic form is nearly identical. The only differences are the variable amplitude due to the presence of the external current input πΌ and the appearance of πΎ which contains conductivity parameters within it. This form will now be introduced into the projected dynamics. 61 Applying Projected Dynamics for a Square Pulse Using Initial Conditions As before for the Gaussian, the goal of projected dynamics is to solve: ππΜ = −ππΈ. (3.48) where π is given by: ∞ ππ ππ πππ = ∫ ππ₯. πππ πππ (3.49) −∞ When the equation for the threshold surface was worked out before, an approximation was made where the width was set to a constant either at the start of the procedure (for onevariable projected dynamics) or at the end after obtaining the projected dynamics (for twovariable projected dynamics). In the Gaussian section, we carried out the analysis explicitly for the two-variable projected dynamics. In this section, we will carry out the analysis explicitly for the one-variable projected dynamics. Doing so allows us to neglect the π matrix entirely in the calculation; with π2 set to a constant π€ at the start, π is reduced from a matrix to a single term as the only surviving term in (3.49) will be π11 . This then makes (3.48): π1Μ = − ∂πΈ/ππ1 . π11 (3.50) Here π1Μ = 0 is satisfied simply when the numerator − ∂πΈ/ππ1 is equal to zero, so the denominator can be neglected in this case. This allows us to do far less calculation to obtain the solution, and we can skip straight to solving the energy integral: ∞ π· ππ 2 π2 π3 πΈ = ∫ ( ( ) + ππΎ − πππ ) ππ₯ 2 ππ₯ 2 3 −∞ 62 (3.51) ∞ π· ππ 2 π2 π3 = 2 ∫ ( ( ) + ππΎ − πππ ) ππ₯. 2 ππ₯ 2 3 0 For the integral on the right, we have exploited the symmetry in π to rewrite the integral over only the positive domain of π₯. Since we have two domains for π, we can further split up this integral to get: π€ 2 2 3 (3.52) ππ1 2π1 πΈ = ∫ (π· ( ) + ππΎ π1 2 − πππ ) ππ₯ ππ₯ 3 0 ∞ ππ2 2 2π2 3 2 + ∫ (π· ( ) + ππΎ π2 − πππ ) ππ₯. ππ₯ 3 π€/2 The two derivatives are obtained by differentiating (3.46) and (3.47): ππ1 = −π1 πΎπ −πΎπ€⁄2 sinh(πΎπ₯), ππ₯ π€ ππ2 1 = − π1 πΎπ −πΎ(π₯− 2 ) (1 − π −πΎπ€ ). ππ₯ 2 (3.53) (3.54) We now Insert (3.46), (3.47), (3.53), and (3.54) into (3.52), perform the integration, and simplify to yield: πΈ= 1 π1 2 π −2πΎπ€ (−π1 πππ + π 2πΎπ€ (3πΎ 2 + ππΎ (−9 + 6πΎπ€) + 9πΎπ1 −) 12πΎ (3.55) 4πΎπππ π1 π€ + π πΎπ€ (3ππΎ (3 + πΎπ€) − 8πππ π1 − 3πΎ(πΎ + πΎ 2 π€ + 2πππ π1 π€)). Calculating the negative of the energy gradient for the time derivative of the amplitude yields: π1Μ = − − ππΈ , ππ1 ππΈ 1 = π π −2πΎπ€ (π1 πππ + 2π πΎπ€ (3πΎ 2 + πΎ 3 π€ − ππΎ (3 + πΎπ€)) ππ1 4πΎ 1 +4π1 πππ + 3πΎπππ π1 π€ + π 2πΎπ€ (−2πΎ 2 + π(6 − 4πΎπ€) + π1 (−9 + 4πΎπ€)πππ )). 63 (3.56) (3.57) Finally, we set (3.57) equal to zero and solve for π1 = ππ . After simplifying, we arrive at our approximate solution for the threshold: ππ = 4π πΎπ€ ππΎ (1 + π πΎπ€ (πΎπ€ − 1)) . πππ (1 + 2π πΎπ€ (4 + 3πΎπ€) + π 2πΎπ€ (−9 + 4πΎπ€)) (3.58) Analysis of the Threshold Equation for a Square Pulse Using Initial Conditions Both the one-variable projected dynamics, with the width set to a constant at the start of the procedure, and the two-variable projected dynamics, with the width set to a constant after the procedure, were plotted together with the numerical solution in Figure 18 below. Figure 18. Plot of the critical amplitude (ππ ) vs. the pulse width (π€) for the projected dynamics with a square pulse. 64 The numerical solution was provided using: ππ£ ∂2 π£ = π· 2 − ππΎ π£ + πππ π£ 2 , ππ‘ ∂π₯ (3.59) and the parameters chosen were ππΎ = 0.139, πππ = 1.0, π· = 1.0. In this case, the initial pulse was specified using initial conditions for the potential as before in the Gaussian case (see Appendix: Numerical Methods). The more realistic case, where a term containing the constant πΌ is applied as an external pulse instead of initial conditions, will be treated separately in the next section. While examining Figure 18, we see right away that the solution obtained by setting the pulse width to a constant at the start of the projected dynamics is drastically improved over the solution obtained under these same conditions for the Gaussian pulse (see Figure 14). In fact, despite the fact that the curve requires larger amplitudes for a given pulse width, the shape of the curve is arguably a much better fit for the full range of widths than the curve obtained by setting the width to a constant after the projected dynamics. From this point forward, we shall focus our attention on the one-variable projected dynamics where the width is set to a constant at the start of the procedure. The equation for this curve is: ππ = 4π πΎπ€ ππΎ (1 + π πΎπ€ (πΎπ€ − 1)) . πππ (1 + 2π πΎπ€ (4 + 3πΎπ€) + π 2πΎπ€ (−9 + 4πΎπ€)) (3.60) The first thing we would like to do with equation (3.60) is check the large width limit to see if the threshold value is improved over the result for the Gaussian pulse. For π€ → ∞, the dominating terms in (3.60) will be those containing both π 2πΎπ€ and π€. In this limit, (3.60) reduces to: 65 4πΎπ€π 2πΎπ€ ππΎ ππΎ ππ = → π = . π 4πΎπ€π 2πΎπ€ πππ πππ (3.61) If we would like to compare this result to the older result, we need to briefly reintroduce Fitzhugh’s notation by substituting (3.6) and (3.7) into (3.61): ππ = ππΎ ππΌ = = πΌ. πππ π (3.62) We see that, as π€ → ∞, the threshold requirement reduces precisely to the threshold πΌ in complete agreement with the same limit obtained from the numerical simulation. Thus, the square pulse approach has fixed one of the issues from the Gaussian projected dynamics where the threshold value at π€ → ∞ was 5% larger than πΌ. We may also check equation (3.60) in the other extreme limit of small pulse widths. Performing a series expansion on (3.60), and keeping terms to the second order, gives: ππ ≅ 6ππΎ 2ππΎ 2ππΎ πΎπ€ 3(ππΎ πΎ 2 )π€ 2 − + − . πΎπππ π€ πππ πππ 2πππ (3.63) In the small width limit, as π€ → 0, the dominant term in the expansion is: ππ ≅ 6ππΎ . πΎπππ π€ (3.64) The scaling goes as 1⁄π€ as opposed to the 1/π€ 2 scaling in the Gaussian approach (2.57). It appears that using one-variable projected dynamics qualitatively improves upon the small pulse limit in comparison to this limit in the two-variable approach (see Figure 18). This qualitative improvement in the shape of the nullcline compared to the true threshold is the main reason that we will work on refining the approximate solution obtained from the one-variable projected dynamics. Returning to the large pulse width limit, equation (3.61) implies that, in our square pulse model, the critical amplitude at large widths will not be specified by a single parameter as 66 in the Fitzhugh approach. Rather, the threshold will arise from a competition between the two independent parameters ππΎ and πππ . Increasing the parameter ππΎ increases the strength for the term that drives the system back to rest. To overcome this, a greater initial amplitude must be applied to get the cells to excite. On the other hand, increasing πππ has the opposite effect of causing the potential to rise faster and subsequently lowers the threshold requirement. The above interpretation can be visualized by first plotting the critical amplitude as a function of ππΎ . This was done in Figure 19 where (3.60) was plotted with π€ = 5, π· = 1.0, πππ = 1.0, and πΎ = √ππΎ for the theory curve. Equation (3.59) was solved using the appropriate parameters for the numerical case. Figure 19. Plot of the critical amplitude (ππ ) vs. the potassium ion channel conductivity (ππΎ ) for the projected dynamics with a square pulse. 67 Likewise, the critical amplitude as a function of πππ with π€ = 5, π· = 1.0, ππΎ = 0.139, and πΎ = √0.139 ~ 0.373 was plotted alongside the numerical solution in Figure 20. Figure 20. Plot of the critical amplitude (ππ ) vs. the sodium ion channel conductivity (πππ ) for the projected dynamics with a square pulse. It is worthwhile to note that equation (3.60) reduces to an especially simple form for the present case: ππ ≅ 0.3895 . πππ (3.65) Comparing equation (3.65) to the Gaussian result in equation (2.57) reveals that both approaches have the same ~1/π dependence for the excitation parameter. 68 For the longitudinal conductivity π·, (3.60) was plotted in Figure 21 using π€ = 5, πππ = 1.0, ππΎ = 0.139, and πΎ = √0.139⁄π· ~ 0.373⁄√π· together with the numerical solution. In comparing Figure 21 to the prior result from the Gaussian projected dynamics (Figure 16 and equation (2.57)), it is seen that the theory equation for the square pulse no longer predicts a linear relationship for the amplitude vs. the longitudinal conductivity as it did in the Gaussian case. Figure 21. Plot of the critical amplitude (ππ ) vs. the longitudinal conductivity (π·) for the projected dynamics with a square pulse. While observing Figures 18-21, the greatest defect in each plot seems to be that the theory equation (3.60) consistently predicts larger values for the critical amplitude when compared to the numerical results. This issue presumably has to do in part with the choice 69 of setting the pulse width equal to a constant at the start of the procedure, but another effect is that our voltage profile for the square pulse approach was derived from (3.9) with a nonzero applied current πΌ. This is not reflected in the present analysis where πΌ was set to zero in both the projected dynamics and the numerical simulations. Therefore, the next step in the procedure is to introduce the applied current πΌ and redo the analysis. Although one might expect this addition to have a trivial effect on the outcome that turns out not to be the case. 70 4. FORMULATION OF THE THRESHOLD EQUATION USING A CONSTANT SQUARE PULSE From Using the Initial Conditions for a Square Pulse to a Constant Square Pulse In the prior treatments, it had always been assumed that treating the applied pulse using initial conditions for the potential would be a sufficient approximation to the physiological case. The physiological case of course requires that the pulse be applied to the cells by an external current source for some finite period of time. In fact, in many cases a pulse of approximately constant magnitude is applied to a chain of cells via an electrode for a time that is much larger than the rather short time that it takes for a chain of cells to excite. To account for this applied current pulse, a term πΌ(π₯) is added to the governing equation: ππ£ ∂2 π£ = π· 2 − ππΎ π£ + πππ π£ 2 + πΌ(π₯). ππ‘ ∂π₯ (4.1) In the prior section, we took πΌ(π₯) to be zero for all π₯ as we treated the external pulse by specifying the initial condition for the potential π£ over a certain number of cells given by the initial pulse width π€. Now we will take πΌ(π₯) to be non-zero for some initial pulse width. The initial condition for the potential is taken to be zero, as the external applied current πΌ(π₯) takes the place of specifying non-zero initial conditions for the potential. The effect this change has on the critical amplitude turns out to be rather dramatic. The difference between the two approaches is shown below in Figure 22. Although the shapes are similar, the actual values for the critical amplitude can be seen to be significantly lower across all pulse widths when a constant pulse πΌ(π₯) is applied in place of using initial conditions for the potential. 71 Figure 22. A comparison of the numerical results for the critical amplitude vs. pulse width from two different approaches to applying a square pulse. In resolving this discrepancy between the two curves, it was found that the criteria for exciting the cells is totally different for each of the two cases. To see this, let’s start with equation (4.1) and look for the “space clamped solution”. This is the solution where a constant applied pulse πΌ is applied to the entire chain in the π€ → ∞ limit. In this limit we can take ∂2 π£⁄∂π₯ 2 = 0 as all cells are stepped up to the same potential by the applied current simultaneously so that no gradient forms between adjacent cells (alternatively, this limit can be thought of as the limit for a chain that contains only a single cell within it; in this case, the spatial derivative is zero as well). 72 Searching for the time-independent threshold, we reduce equation (4.1) to: 0 = −ππΎ π£ + πππ π£ 2 + πΌ. (4.2) This is just a quadratic equation, whose positive root π£ = ππ is given by: ππ = ππΎ + √ππΎ 2 − 4(πππ )πΌ . 2πππ (4.3) If we take πΌ = 0, the critical amplitude is given by: ππΎ + √ππΎ 2 ππΎ ππ = → ππ = . 2πππ πππ (4.4) This is just the result that we had earlier from the projected dynamics for the square pulse using initial conditions for the potential (see equation (3.61)). It shows that the critical amplitude in the limit of a large pulse width is just the positive root of equation (4.2). If we take ππΎ = 0.139 and πππ = 1.0 then we have ππ = 0.139 which was verified by solving (4.1) numerically for the threshold in the large pulse width limit. Now, if we take πΌ ≠ 0, the criteria for excitation is completely different. Instead of being a root of (4.2), the new criteria for excitation is merely that the solution for ππ be a real number. In order to ensure this, we must have: √ππΎ 2 − 4(πππ )πΌ ≥ 0. (4.5) For this equation to hold true, the minimum requirement is met if the term inside the square root sign is equal to zero. Thus the critical amplitude (πΌπ ) in this case turns out to be: ππΎ 2 − 4(πππ )πΌ = 0, πΌ= πΌπ = ππΎ 2 , 4πππ πΌ ππΎ = . ππΎ 4πππ 73 (4.6) (4.7) (4.8) We see that the new threshold requirement πΌπ is reduced by a factor of four in comparison to the threshold for the instantaneous square pulse. Using ππΎ = 0.139 and πππ = 1.0, we now have that πΌπ ≅ 0.035 which agrees with the result obtained numerically in the limit of a large pulse width. Reformulating the Projected Dynamics Using a Constant Square Pulse Now that we have some idea of how the nature of the threshold curve changes as we apply a constant square pulse, we have to modify the method used to obtain the projected dynamics in two ways. First, we need to introduce the external current pulse πΌ(π₯) to the expression for the total energy. Second, after we carry through the procedure and obtain the expression for ππ , we will need to extract the true criteria for excitation from underneath the square root sign, and then solve for πΌπ . We will follow the same basic approach from before, but we will make the appropriate changes as we go along. Starting with: ππΜ = −ππΈ, (4.9) where π is given by: ∞ πππ = ∫ −∞ ππ ππ ππ₯, πππ πππ (4.10) we again choose to do one-variable projected dynamics and set the pulse width equal to a constant π€ at the start of the procedure. This allows us to write (4.9) as: π1Μ = − ∂πΈ/ππ1 . π11 74 (4.11) The condition π1Μ = 0 is met when the numerator − ∂πΈ/ππ1 is equal to zero, and so we can ignore π11 and focus on obtaining the energy integral. Now the applied current pulse πΌ(π₯) is: πΌ(π₯) = πΌ πππ π₯ < π€ ⁄2, (4.12) πΌ(π₯) = 0 πππ π₯ > π€ ⁄2. (4.13) Introducing this current pulse to the energy integral, we will arrive at: π€ 2 ππ1 2 2π1 3 πΈ = ∫ (π· ( ) + ππΎ π1 2 − πππ − 2πΌπ1 ) ππ₯ ππ₯ 3 (4.14) 0 ∞ ππ2 2 2π2 3 2 + ∫ (π· ( ) + ππΎ π2 − πππ ) ππ₯. ππ₯ 3 π€/2 To evaluate (4.14), we need the functions π1 and π2 and their two derivatives which are unchanged from before: π1 = π1 (1 − π −πΎπ€⁄2 cosh(πΎπ₯)) π2 = πππ π₯ < π€ ⁄2, π1 (1 − π −πΎπ€ )π −πΎ(π₯−π€⁄2) πππ π₯ > π€/2, 2 ππ1 = −π1 πΎπ −πΎπ€⁄2 sinh(πΎπ₯), ππ₯ π€ ππ2 1 = − π1 πΎπ −πΎ(π₯− 2 ) (1 − π −πΎπ€ ). ππ₯ 2 (4.15) (4.16) (4.17) (4.18) Inserting (4.15)-(4.18) into (4.14), performing the integration, and simplifying the result leads to: πΈ= 1 −2πΎπ€ π (−π1 3 πππ − π1 π 2πΎπ€ (12πΌ(−1 + πΎπ€) + π1 2 πππ (−9 + 4πΎπ€)) 12πΎ −3π1 (π·πΎ 2 + ππΎ − (3 + 2πΎπ€)) − π1 π πΎπ€ (12πΌ + π1 (3π·πΎ 2 (1 + πΎπ€) −3ππΎ (3 + πΎπ€) + 2π1 πππ (4 + 3πΎπ€)))), and then: 75 (4.19) π1Μ = − ππΈ 1 −2πΎπ€ 2 = π (π1 πππ + π 2πΎπ€ (6π1 ππΎ − 9π1 2 πππ − 4πΌ + ππ1 4πΎ (4.20) 4(−π1ππΎ + π1 2 πππ + πΌ)πΎπ€ − 2π1 π·πΎ 2 ) + 2π πΎπ€ (2πΌ + π1 (π·πΎ 2 (1 + πΎπ€) − ππΎ (3 + πΎπ€) + π1 πππ (4 + 3πΎπ€)))). Setting π1Μ = 0 and solving for π1 gives us the following quadratic equation: π πΎπ€ (−π·πΎ 2 (1 + πΎπ€) + ππΎ (3 + πΎπ€)) + π 2πΎπ€ (π·πΎ 2 + ππΎ (−3 + 2πΎπ€)) ± √π π1 = , πππ + 2π πΎπ€ πππ (4 + 3πΎπ€) + π 2πΎπ€ πππ (−9 + 4πΎπ€) (4.21) where the term under the square root sign, π, is given by: π = π πΎπ€ (π πΎπ€ (π·πΎ 2 (1 − π πΎπ€ + πΎπ€) − ππΎ (3 + πΎπ€ + π πΎπ€ (−3 + 2πΎπ€)))2 (4.22) −4πππ πΌ(1 + π πΎπ€ (−1 + πΎπ€))(1 + π πΎπ€ (8 + 6πΎπ€ + π πΎπ€ (−9 + 4πΎπ€)))). The criteria for equation (4.22) to be real is: π ≥ 0, (4.23) and the minimum criteria for excitation is then: π = 0 = π πΎπ€ (π πΎπ€ (π·πΎ 2 (1 − π πΎπ€ + πΎπ€) − ππΎ (3 + πΎπ€ + π πΎπ€ (−3 + 2πΎπ€)))2 (4.24) −4πππ πΌ(1 + π πΎπ€ (−1 + πΎπ€))(1 + π πΎπ€ (8 + 6πΎπ€ + π πΎπ€ (−9 + 4πΎπ€)))). This can be solved for πΌ: πΌ= π πΎπ€ (π·πΎ 2 (1 − π πΎπ€ + πΎπ€) − ππΎ (3 + πΎπ€ + π πΎπ€ (−3 + 2πΎπ€)))2 , 4πππ (1 + π πΎπ€ (−1 + πΎπ€))(1 + π πΎπ€ (8 + 6πΎπ€ + π πΎπ€ (−9 + 4πΎπ€))) (4.25) and then further simplified to: ππΎ 2 π πΎπ€ (1 + π πΎπ€ (πΎπ€ − 1)) πΌ= . πππ (1 + 2π πΎπ€ (4 + 3πΎπ€) + π 2πΎπ€ (4πΎπ€ − 9)) (4.26) Finally, we can divide (4.26) by ππΎ to arrive at our desired expression for the critical amplitude πΌπ : πΌπ = ππΎ π πΎπ€ (1 + π πΎπ€ (πΎπ€ − 1)) . πππ (1 + 2π πΎπ€ (4 + 3πΎπ€) + π 2πΎπ€ (4πΎπ€ − 9)) 76 (4.27) A comparison with equation (3.58) shows that the two expressions only differ from one another by a factor of four. While this indicates that the form of the projected dynamics remains largely the same, the agreement between (4.27) and the numerical threshold solution has now substantially improved as discussed in the next section. Analysis of the Threshold Equation for a Constant Square Pulse Equation (4.27) is plotted against the numerical solution of (4.1) in Figure 23. For the numerical solution (see Appendix: Numerical Methods) the parameters used were ππΎ = 0.139, πππ = 1.0, π· = 1.0, and the amplitude was given by πΌπ = πΌ ⁄ππΎ . Figure 23. A plot of the critical amplitude vs. pulse width for a constant applied square pulse. 77 The agreement between theory and the numerical solution has been drastically improved. Since we know that the basis for our approximation is the amplitude nullcline, and since we took the width as constant at the start of the projected dynamics, it seems reasonable to attribute the improved accuracy for the threshold curve to a diminished dependence on the time evolution of the pulse width for the case of a constant square pulse. At this point, it is important to check and see if equation (4.27) reduces properly to the new threshold criteria that was established in equation (4.8). The expression for the critical amplitude πΌπ was given in (4.27) as: ππΎ π πΎπ€ (1 + π πΎπ€ (πΎπ€ − 1)) πΌπ = . πππ (1 + 2π πΎπ€ (4 + 3πΎπ€) + π 2πΎπ€ (4πΎπ€ − 9)) (4.28) In the limit where π€ → ∞, the largest terms will be those that include both π 2πΎπ€ and π€ as a product. This reduces (4.28) to: πΌπ = πΎπ€ππΎ π 2πΎπ€ ππΎ → πΌπ = . 2πΎπ€ 4πΎπ€πππ π 4πππ (4.29) Indeed, equation (4.29) reduces to the proper threshold in the limit of a large pulse width. The small width limit also checks out where the leading term in the series expansion is: πΌπ ~ 3ππΎ , 2πΎπππ π€ (4.30) which differs from the result in equation (3.64) only by the constant factor of 1/4. The strong agreement in Figure 23 provides evidence that the correct scaling at small widths is indeed 1/π€ . Now it is time to examine threshold behavior for the ion channel and longitudinal conductivities. Figure 24 shows the result for the potassium ion channel conductivity while Figure 25 shows the result for the sodium ion channel conductivity. 78 Figure 24. A plot of the critical amplitude vs. potassium ion channel conductivity for a constant applied square pulse. Figure 25. A plot of the critical amplitude vs. sodium ion channel conductivity for a constant applied square pulse. 79 Both plots were generated using the same parameter values as in the prior section: πππ = 1.0, π· = 1.0, and π€ = 5 for the potassium ion channel plot, and ππΎ = 0.139, π· = 1.0, and π€ = 5 for the sodium ion channel plot. For the case of sodium, the solution reduces to a simple relationship (πΌπ ~ 0.0973769/πππ ) as it did before in the instantaneous case with the only difference being that the value of the constant has now been reduced by a factor of four as in the prior cases. Although the potassium result begins to diverge from theory after a certain point, the agreement is still quite good for small values of ππΎ . In general, the same can be said for the longitudinal conductivity, which is plotted below in Figure 26 where the numerical parameters were ππΎ = 0.139, πππ = 1.0, and π€ = 5. Figure 26. A plot of the critical amplitude vs. longitudinal conductivity for a constant applied square pulse. 80 In Figure 26, the trend for the theoretical approach is greatly improved over the prior case, where initial conditions for the potential were used, but it still starts to become more and more inaccurate as the longitudinal conductivity increases in value. Perhaps this is to be expected as the longitudinal conductivity affects the width of the pulse more than any other parameter, and we have neglected the time dependence of the width in the foregoing analysis. However, this issue is minimized for small values of the conductivity, and the constant width approximation is certainly less of an issue in the present case than it was in the prior case. In comparing Figures 23-26, we see that, in general, equation (4.28) provides a far more accurate description of the threshold for the constant square pulse in comparison to the numerical result than what came before. 81 5. FORMULATION OF THE THRESHOLD EQUATION FOR A TIME-DEPENDENT SQUARE PULSE Generalizing the Approach to find the Threshold for a Time-Dependent Square Pulse In the previous section, an equation for the excitation threshold was obtained for the case where an external square current pulse was applied for some time that was considered to be much longer than the time needed to excite the cells. In such a situation the applied current pulse was treated as being a constant since, by the time the current was switched off, the cells had either long since been excited or they would not be excited by the current no matter how long it was applied. But what about the more general case where the external current can be applied for some finite amount of time π? Can the projected dynamics be used to obtain not only the solution for when π is very large, as was done previously, but also the case where π can be specified to be any amount of time we wish? Indeed, such a general solution can be found. To better understand the approach, we will first start with the simplest case where the entire cell chain is excited to the same level all at once. In doing so, the system becomes spatially homogeneous, and we can neglect the spatial derivative in equation (4.1). This reduces equation (4.1) to: ππ£ = −ππΎ π£ + πππ π£ 2 + πΌ, ππ‘ which can be rearranged to solve for the differential of time: ππ£ = ππ‘. −ππΎ π£ + πππ π£ 2 + πΌ 82 (5.1) (5.2) At the point when the current is switched off, at the time π‘ = π, the system instantly reverts back to the πΌ = 0 dynamics, where the threshold requirement is found from: 0 = −ππΎ π£ ∗ + πππ π£ ∗ 2 , π£∗ = (5.3) ππΎ . πππ (5.4) Equation (5.4) is in agreement with equation (3.61) that was obtained for the threshold in the case π€ → ∞ using projected dynamics with πΌ = 0. The potential must rise from its initial value of zero to the value given in equation (5.4) by the time the current is switched off in order for the system to fire an action potential. The integral can now be expressed as: ∫ ππΎ ⁄πππ 0 π ππ£ = ∫ ππ‘ = π, −ππΎ π£ + πππ π£ 2 + πΌ 0 (5.5) and, after performing the integration, (5.5) becomes: ππΎ 4 tan−1 ( ) √−ππΎ 2 + 4πππ πΌ π= . √−ππΎ 2 + 4πππ πΌ (5.6) In equation (5.6), we can see that the threshold condition for non-zero values of πΌ, which is that the solution merely exist, is present inside the square root sign and is identical in form to the threshold condition worked out previously in equations (4.5) to (4.8) for this case. At this point, our goal is to invert equation (5.6) to obtain an expression for the critical amplitude as a function of π. Letting π¦ = √−ππΎ 2 + 4πππ πΌ, equation (5.6) becomes: π= π 4 tan−1 ( π¦πΎ ) π¦ 83 (5.7) , tan ( π¦π ππΎ )= . 4 π¦ (5.8) If we assume that π¦π⁄4 βͺ 1, we can make the following approximation: π¦π π¦π )≅ , 4 4 (5.9) 4ππΎ , π (5.10) tan ( which leads to: π¦2 = −ππΎ 2 + 4πππ πΌ = πΌ= 4ππΎ , π (5.11) ππΎ 2 ππΎ + . 4πππ ππππ (5.12) ππΎ 1 + . 4πππ ππππ (5.13) The critical amplitude is then: πΌπ = Equation (5.13) is plotted in Figure 27 against the numerical result (see Appendix: Numerical Methods) for a time-dependent current pulse πΌ(π‘) applied uniformly to a chain of cells using π· = 1.0, πππ = 1.0, and ππΎ = 0.139 as parameters. Let us now pause for a moment to examine the features of equation (5.13). First, we note that, as π → ∞, equation (5.13) reduces to: πΌπ = ππΎ , 4πππ (5.14) which is in complete agreement with the threshold for a homogeneous constant square pulse given by equation (4.8). At the other extreme where π → 0, the requirement to excite becomes infinitely large. Qualitatively this result makes sense as, given less time to excite the cell, the required potential difference to excite must be greater so that the necessary 84 charge is driven across the membrane in a shorter period of time. This is reflected in the increase that is required for the critical amplitude πΌπ when π is small. Figure 27. A plot of the critical amplitude vs. the time that the pulse was applied for a uniformly applied square pulse. Also of physiological interest here is the inverse dependence of the second term in equation (5.13) on the sodium ion channel conductivity, πππ . Other studies and models have indicated that the most important factor in reducing excitability is a reduced availability of sodium channels associated with a smaller value of πππ [15]. In equation (5.13), we see that lowering the sodium conductivity raises the critical amplitude requirement in both terms; not only does it raise the critical amplitude for the case when π → ∞, but it also 85 serves to increase the second term for the case of finite π. Hence, equation (5.13) proves to be remarkably sensitive to a decrease in the sodium conductance which is in qualitative agreement with those earlier observations. It also suggests that, to decrease the risk for excitation failure, the pulse application time π for the externally applied current must not be too small. Although π does not affect the threshold requirement from the first term in (5.13) at all, a small enough value of π will cause the second term in equation (5.13) to completely dominate the threshold requirement. Obtaining the Threshold Equation for a Time-Dependent Square Pulse We have obtained a time-dependent threshold solution for the case where the pulse is uniformly applied to all cells in the chain, but we also wish to obtain the solution where the applied current pulse may only be applied to a finite number of cells (of pulse width π€) in the cell chain. For this we take the same approach as above, but here we start with the equation for the time evolution of the amplitude that we get by applying the method of projected dynamics. Much of the procedure to obtain such an equation is exactly the same as it was when projected dynamics was performed earlier for the constant square pulse. However, one step at the beginning of the procedure must be modified; for this, let us take a closer look at equation (4.11): π1Μ = − ∂πΈ/ππ1 . π11 (5.15) Earlier, when we used projected dynamics for the square pulse, we were interested only in the time-independent solution for the amplitude. Since π1Μ = 0 in that case, we were able to ignore the M matrix, and we did not have to evaluate π11 explicitly. Now, we now seek 86 an expression for π1Μ itself, and so, we must evaluate π11 for the present case. The matrix element π11 is given by: π€/2 π11 ∞ ππ1 2 ππ2 2 =∫ ( ) ππ₯ + ∫ ( ) ππ₯. ππ1 ππ1 0 (5.16) π€/2 To evaluate equation (5.16), we differentiate π1 and π2 given in equations (4.15) and (4.16) with respect to π1 : π1 = π1 (1 − π −πΎπ€⁄2 cosh(πΎπ₯)) π2 = πππ π₯ < π€ ⁄2, π1 (1 − π −πΎπ€ )π −πΎ(π₯−π€⁄2) πππ π₯ > π€/2, 2 ππ1 = (1 − π −πΎπ€⁄2 cosh(πΎπ₯)), ππ1 ππ2 1 = (1 − π −πΎπ€ )π −πΎ(π₯−π€⁄2) . ππ1 2 (5.17) (5.18) (5.19) (5.20) After placing (5.19) and (5.20) into (5.16) and performing the integration, we arrive at: π11 π −2πΎπ€ (π πΎπ€ − 1)2 2πΎπ€ − 4 + π −πΎπ€ (4 + πΎπ€ + sinh(πΎπ€)) = + . 4πΎ 2πΎ (5.21) The negative of the energy gradient, − ∂πΈ/ππ1, is given exactly as it was before in equation (4.20), namely: − ππΈ 1 −2πΎπ€ 2 = π (π1 πππ + π 2πΎπ€ (6π1 ππΎ − 9π1 2 πππ − 4πΌ + ππ1 4πΎ (5.22) 4(−π1 ππΎ + π1 2 πππ + πΌ)πΎπ€ − 2π1 π·πΎ 2 ) + 2π πΎπ€ (2πΌ + π1 (π·πΎ 2 (1 + πΎπ€) − ππΎ (3 + πΎπ€) + π1 πππ (4 + 3πΎπ€)))). Substituting (5.21) and (5.22) into equation (5.15) gives us our desired expression for the time derivative of the amplitude: 87 1 −2πΎπ€ 2 (π1 πππ + π 2πΎπ€ (6π1 ππΎ − 9π1 2 πππ − 4πΌ + 4(−π1 ππΎ 4πΎ π +π1 2 πππ + πΌ)πΎπ€ − 2π1 π·πΎ 2 ) + 2π πΎπ€ (2πΌ + π1 (π·πΎ 2 (1 + πΎπ€) − ππ1 ππΎ (3 + πΎπ€) + π1 πππ (4 + 3πΎπ€)))) = , −2πΎπ€ πΎπ€ π (π − 1)2 2πΎπ€ − 4 + π −πΎπ€ (4 + πΎπ€ + sinh(πΎπ€)) ππ‘ + 4πΎ 2πΎ (5.23) while some further simplification of equation (5.23) leads to: (π −πΎπ€ (π1 2 πππ + π 2πΎπ€ (6π1 ππΎ − 9π1 2 πππ − 4πΌ + 4(−π1 ππΎ + π1 2 πππ + πΌ)πΎπ€ − 2π1 ππΎ ) + 2π πΎπ€ (2πΌ + π1 (ππΎ (1 + πΎπ€) ππ1 −ππΎ (3 + πΎπ€) + π1 πππ (4 + 3πΎπ€))))) = . ππ‘ 2(3 + πΎπ€ + π πΎπ€ (2πΎπ€ − 3)) (5.24) Now, we want to group all the terms in (5.24) containing ππΎ , all the terms containing πππ , and all the terms containing πΌ separately; then we can write the right side of the equation in a way the resembles how we wrote equation (5.1). After doing this, we end up with: ππ1 = −π΄π1 + π΅π1 2 + πΆπΌ, ππ‘ (5.25) where: 2(1 + π πΎπ€ (πΎπ€ − 1)) , 3 + πΎπ€ + π πΎπ€ (2πΎπ€ − 3) (5.26) 8 + π −πΎπ€ + 6πΎπ€ + π πΎπ€ (4πΎπ€ − 9) , 2(3 + πΎπ€ + π πΎπ€ (2πΎπ€ − 3)) (5.27) π΄ = ππΎ π΅ = πππ πΆ= 2(1 + π πΎπ€ (πΎπ€ − 1)) . 3 + πΎπ€ + π πΎπ€ (2πΎπ€ − 3) (5.28) Equation (5.25) can be rearranged to solve for the differential of time: ππ1 = ππ‘. −π΄π1 + π΅π1 2 + πΆπΌ (5.29) Following in the same manner as we did before for the case of an infinite pulse width, we wish to integrate (5.29) to find an expression for the applied current amplitude as a function of the time duration of the applied current, π. At the time the current is switched off, the 88 threshold requirement becomes the same as for the πΌ = 0 case, and we find that the threshold requirement for the potential (π1 ∗ ) can be written as: 0 = −π΄π1 ∗ + π΅π1 ∗ 2 , π1 ∗ = (5.30) π΄ . π΅ (5.31) The integral of (5.29) is then: π΄/π΅ ∫ 0 π ππ1 = ∫ ππ‘, −π΄π1 + π΅π1 2 + πΆπΌ 0 π= (5.32) π΄ 4 tan−1 ( π¦ ) (5.33) , π¦ with π¦ given by: (5.34) 2 π¦= −4ππΎ 2 (1 + π πΎπ€ (πΎπ€ − 1)) + √4πππ πΌ(1 + π πΎπ€ (πΎπ€ − 1))(8 + π −πΎπ€ + 6πΎπ€ + π πΎπ€ (4πΎπ€ − 9)) (3 + πΎπ€ + π πΎπ€ (2πΎπ€ − 3)) 2 . The goal once again is to solve for the critical amplitude as a function of π. Proceeding exactly as before: π¦π π΄ (5.35) )= , 4 π¦ π¦π π¦π (5.36) tan ( ) ≅ , 4 4 4π΄ (5.37) π¦2 = . π Inserting equations (5.26) and (5.34) into (5.37), simplifying, and rearranging leads to: tan ( πΌ= (1 + π πΎπ€ (πΎπ€ − 1)) ππΎ 2 ( )+ πππ (8 + π −πΎπ€ + 6πΎπ€ + π πΎπ€ (4πΎπ€ − 9)) 2(3 + πΎπ€ + π πΎπ€ (2πΎπ€ − 3)) ππΎ ( ). ππππ (8 + π −πΎπ€ + 6πΎπ€ + π πΎπ€ (4πΎπ€ − 9)) 89 (5.38) Dividing (5.38) by ππΎ gives the expression for the critical amplitude: (1 + π πΎπ€ (πΎπ€ − 1)) ππΎ πΌπ = ( )+ πππ (8 + π −πΎπ€ + 6πΎπ€ + π πΎπ€ (4πΎπ€ − 9)) (5.39) 2(3 + πΎπ€ + π πΎπ€ (2πΎπ€ − 3)) 1 ( ). ππππ (8 + π −πΎπ€ + 6πΎπ€ + π πΎπ€ (4πΎπ€ − 9)) Using (5.26) and (5.27), this result can also be written more compactly as: πΌπ = π΄ 1 + . 4π΅ π΅π (5.40) In Figure 28, equation (5.39) is plotted against the numerical threshold solution with π· = 1.0, πππ = 1.0, and ππΎ = 0.139. The case where the pulse width is π€ = 100 cells (essentially the homogeneous case) and the case where the pulse width is π€ = 5 cells are included for comparison. Figure 28. A plot of the critical amplitude vs. the time that the square pulse was applied is provided for both a 100 cell pulse width and a 5 cell pulse width. 90 We now wish to check equation (5.39) in both the large pulse width limit and the large pulse time limit. As π€ → ∞, the only surviving terms contain the product π πΎπ€ πΎπ€, and the equation for the critical amplitude reduces to: πΌπ = ππΎ π πΎπ€ πΎπ€ 1 2(π πΎπ€ 2πΎπ€) ππΎ 1 ( πΎπ€ )+ ( πΎπ€ )= + , πππ (π 4πΎπ€) ππππ (π 4πΎπ€) 4πππ ππππ (5.41) which is exactly what we found before in equation (5.13) for the large pulse width limit. As π → ∞, we find that the second term in equation (5.39) vanishes to give: πΌπ = (1 + π πΎπ€ (πΎπ€ − 1)) ππΎ ( ). πππ (8 + π −πΎπ€ + 6πΎπ€ + π πΎπ€ (4πΎπ€ − 9)) (5.42) Multiplying the top and bottom of this equation by π πΎπ€ and rewriting 8 + 6πΎπ€ in the denominator as 2(4 + 3πΎπ€) yields: πΌπ = ππΎ π πΎπ€ (1 + π πΎπ€ (πΎπ€ − 1)) . πππ (1 + 2π πΎπ€ (4 + 3πΎπ€) + π 2πΎπ€ (4πΎπ€ − 9)) (5.43) which is the same as equation (4.27) which was derived earlier for the case of a constant square pulse. Equation (5.39) then emerges as our most general approximate solution to the one-dimensional square pulse threshold problem. (5.39) answers the question of whether or not an action potential will fire given the width and time of application of an applied square pulse using a wide range of conductivity parameters. 91 6. A SIMPLIFIED APPROXIMATE SOLUTION FOR THE SQUARE PULSE Introducing a Simplified Initial Voltage Profile for the Constant Square Pulse From the foregoing analysis, we have obtained a rather accurate approximate analytical expression of the excitation threshold for the case of a continuously applied square pulse: πΌπ = ππΎ π πΎπ€ (1 + π πΎπ€ (πΎπ€ − 1)) . πππ (1 + 2π πΎπ€ (4 + 3πΎπ€) + π 2πΎπ€ (4πΎπ€ − 9)) (6.1) We also have obtained a less accurate, but easier to understand, model of the excitation threshold for the case of an applied Gaussian pulse specified by using initial conditions for the potential: ππ = 3√6 π· ( + πΌ). 7 ππ€ 2 (6.2) Both results have their strengths and weaknesses. The strength of equation (6.1) lies in how accurately it is able to predict the results of the numerical solution over a wide range of the given parameters. This gives us confidence in the method we are using and also sheds some light on the sources of error in our less accurate earlier attempts. The downside to equation (6.1) is that it is relatively difficult to analyze, and only a limited amount of useful information can be extracted from it. Equation (6.2) suffers from several defects in that it doesn’t give the correct value for the critical amplitude at the large pulse width limit and differs significantly from the numerical solution at the small pulse width limit. Despite these drawbacks, the qualitative predictions made by equation (6.2), concerning the behavior of the system as the various parameters are changed, are still in complete agreement with the trends predicted by the more accurate threshold equation (6.1). In both cases, increasing the longitudinal conductivity increases the requirement for a cell to reach 92 the excitation threshold whereas increasing the “sodium” ion channel conductivity and increasing the width of the initial pulse tend to have the opposite effect. In contrast to equation (6.1), all of these trends for each parameter are very easily interpreted from the analysis of equation (6.2). Ideally, what we would like to have is a threshold solution that provides a compromise between the two extremes outlined above. Such a solution should incorporate some of the quantitative improvements gained in going from the Gaussian pulse and equation (6.2) to the constant square pulse and equation (6.1) while decreasing the mathematical complexity of equation (6.1). To accomplish this task, we need to introduce a new initial voltage profile that will then be used with the projected dynamics. Let us start by examining the initial voltage profile that was used previously for the constant square pulse. This was given for the one-variable projected dynamics as: π1 = π(1 − π −πΎπ€⁄2 cosh(πΎπ₯)) π2 = π (1 − π −πΎπ€ )π −πΎ(π₯−π€⁄2) 2 πππ π₯ < π€ ⁄2, πππ π₯ > π€/2. (6.3) (6.4) This initial voltage profile features a relatively flat shape for the central region when π₯ < π€/2, and an exponential decrease outside of that region. The overall shape of this voltage profile then approximates the shape of an applied square pulse. Now, we can introduce those two key features in a simpler way by introducing the following initial voltage profile: π1 = π πππ π₯ < π€ ⁄2, π2 = ππ −πΎ(π₯−π€⁄2) πππ π₯ > π€/2. (6.5) (6.6) It should be noted that the simplified initial voltage profile given above still matches the first boundary condition from solving the partial differential equation, since π1 (π€/2) = π2 (π€/2), but it does not match the condition for the first derivatives to agree at the π€/2 93 boundary. This discontinuity arises from the removal of the curvature in the π1 solution that was due to the hyperbolic cosine term. This is the price that is paid for simplifying equations (6.3) and (6.4) to (6.5) and (6.6) respectively, and we certainly expect it to affect the accuracy of the threshold solution to some extent. Still, let us carry through the projected dynamics to find the threshold solution using equations (6.5) and (6.6). We will then compare the threshold solution we obtain to the numerical result as we have done before, and, in doing so, we will get a qualitative assessment of the error that is introduced. The Projected Dynamics and the Threshold Equation for the Simple Constant Square Pulse The procedure for obtaining the threshold equation is exactly the same as the procedure described previously for a constant square pulse. Briefly, for the one-variable projected dynamics, we need to solve: − ∂πΈ/ππ , π11 (6.7) πΜ = − ∂πΈ/ππ, (6.8) πΜ = or simply: since we can neglect the π11 term when solving the projected dynamics for a constant square pulse. The energy integral is of the same form as equation (4.14), that is: π€ 2 2 3 (6.9) ππ1 2π1 πΈ = ∫ (π· ( ) + ππΎ π1 2 − πππ − 2πΌπ1 ) ππ₯ + ππ₯ 3 0 ∞ ππ2 2 2π2 3 ∫ (π· ( ) + ππΎ π2 2 − πππ ) ππ₯. ππ₯ 3 π€/2 The derivatives needed for (6.9) are obtained by differentiating (6.5) and (6.6) to get: 94 ππ1 = 0, ππ₯ (6.10) ππ2 = −ππΎπ −πΎ(π₯−π€⁄2) . ππ₯ (6.11) With (6.5), (6.6), (6.10), and (6.11) in hand, the integral of equation (6.9) becomes: πΈ= π(−18πΌπΎπ€ − 2π2 πππ (2 + 3πΎπ€) + 9πππΎ (2 + πΎπ€)) , 18πΎ (6.12) and the time derivative for the amplitude is then: πΜ = − ∂πΈ/ππ = 3πΌπΎπ€ + π2 πππ (2 + 3πΎπ€) − 3πππΎ (2 + πΎπ€) . 3πΎ (6.13) Setting πΜ = 0 and solving for π gives us: π= ππΎ 3(2 + πΎπ€) + ±√π , 2πππ (2 + 3πΎπ€) (6.14) where π is: π = −12πΌπππ πΎπ€(2 + 3πΎπ€) + 9ππΎ 2 (2 + πΎπ€)2 . (6.15) The minimum criteria for excitation is met when: π = 0. (6.16) Applying this condition and solving (6.15) for πΌ yields: ππΎ 2 3(2 + πΎπ€)2 πΌ= . 4πππ πΎπ€(2 + 3πΎπ€) (6.17) After dividing (6.17) by ππΎ , we arrive at our expression for the critical amplitude πΌπ : πΌπ = ππΎ 3(2 + πΎπ€)2 . 4πππ πΎπ€(2 + 3πΎπ€) (6.18) Figure 29 shows how equation (6.18) for the simple constant square pulse compares with our previous result for the constant square pulse that was given in Figure 23. 95 Figure 29. A plot of the critical amplitude vs. pulse width for the simple constant applied square pulse. The numerical and constant square pulse results from Figure 23 are provided for comparison. It is also instructive to check the behavior of equation (6.18) in the large and small pulse width limits. At the large pulse width limit, where π€ → ∞, equation (6.18) becomes approximately: πΌπ = 3ππΎ (πΎπ€)2 ππΎ = , 12πππ (πΎπ€)2 4πππ (6.19) and at the small width limit, where π€ → 0, the leading term in the series expansion of (6.18) gives: 96 πΌπ ~ 3ππΎ . 2πΎπππ π€ (6.20) Both of these results are exactly the same as those obtained for the constant square pulse earlier in equations (4.29) and (4.30). Taking that into consideration, the simple constant square pulse does indeed appear to be an acceptable compromise between the Gaussian and constant square pulse approaches. Equation (6.18) is considerably more compact than equation (6.1), and, as Figure 29 reveals, the agreement between the two is quite good (although, from Figure 29, it is also evident that equation (6.18) does not provide as accurate a description of the numerical result as equation (6.1)). At the same time, in the limit of large and small pulse widths, equation (6.18) gives precisely the same results as the more complicated equation (6.1). This feature still stands as a marked improvement over the Gaussian approach which is incorrect in both the large and small pulse width limits. The Threshold Equation for the Simple Time-Dependent Square Pulse For the case of the time-dependent square pulse, we must return to equation (6.7) and evaluate the π11 term. π11 was given previously as: π€/2 π11 ∞ ππ1 2 ππ2 2 =∫ ( ) ππ₯ + ∫ ( ) ππ₯ . ππ ππ 0 (6.21) π€/2 The derivatives with respect to π are: ππ1 = 1, ππ (6.22) ππ2 = π −πΎ(π₯−π€⁄2) , ππ (6.23) which, when placed in (6.21) and integrated, yields: 97 1 . πΎ π11 = π€ + (6.24) Dividing (6.13) by (6.24) we have: ∂πΈ/ππ 3πΌπΎπ€ + π2 πππ (2 + 3πΎπ€) − 3πππΎ (2 + πΎπ€) πΜ = − = , π11 3(1 + πΎπ€) (6.25) which can be rewritten as: ππ = −π΄π + π΅π2 + πΆπΌ, ππ‘ (6.26) where: π΄ = ππΎ π΅ = πππ πΆ= (2 + πΎπ€) , (1 + πΎπ€) (2 + 3πΎπ€) (3(1 + πΎπ€)) (6.27) , πΎπ€ . (1 + πΎπ€) (6.28) (6.29) Equation (6.26) is then integrated as before: π΄/π΅ ∫ 0 π ππ = ∫ ππ‘, −π΄π + π΅π2 + πΆπΌ 0 π= π΄ 4 tan−1 ( π¦ ) (6.30) (6.31) , π¦ π¦π π΄ )= , 4 π¦ π¦π π¦π tan ( ) ≅ , 4 4 4π΄ π¦2 = , π (6.32) 4πππ πΌπΎπ€(2 + 3πΎπ€) ππΎ 2 (2 + πΎπ€)2 √ π¦= − . (1 + πΎπ€)2 3(1 + πΎπ€)2 (6.35) tan ( (6.33) (6.34) with π¦ given as: 98 Inserting (6.35) into (6.34), simplifying, and rearranging brings us to: ππΎ 2 3(2 + πΎπ€)2 ππΎ 3(1 + πΎπ€)(2 + πΎπ€) πΌ= + , 4πππ πΎπ€(2 + 3πΎπ€) ππππ πΎπ€(2 + 3πΎπ€) (6.36) which can be divided by ππΎ to give the desired expression for the critical amplitude as a function of π: πΌπ = ππΎ 3(2 + πΎπ€)2 3(1 + πΎπ€)(2 + πΎπ€) + . 4πππ πΎπ€(2 + 3πΎπ€) ππππ πΎπ€(2 + 3πΎπ€) (6.37) Equation (6.37) is plotted in Figure 30 against the numerical threshold solution with π· = 1.0, πππ = 1.0, and ππΎ = 0.139. The case where the pulse width is π€ = 100 cells and the case where the pulse width is π€ = 5 cells are given to facilitate comparison with the prior treatment in Figure 28. Figure 30. A plot of the critical amplitude vs. the time that the square pulse was applied is provided for both a 100 cell pulse width and a 5 cell pulse width. Here, the simple square pulse approach is compared with the numerical result. 99 An additional discrepancy between theory and the numerical solution once again shows up for the simple square pulse in Figure 30 as compared to the square pulse that was used to produce Figure 28, but the qualitative agreement still seems to be quite good for the simple square pulse. Finally, we have to check equation (6.37) in the large pulse width and large pulse time limits to see if the theory reduces properly. As π€ → ∞, the surviving terms contain the product 3(πΎπ€)2, and the equation for the critical amplitude reduces to: πΌπ = ππΎ 3(πΎπ€)2 1 3(πΎπ€)2 ππΎ 1 ( ) + ( )= + , 2 2 4πππ 3(πΎπ€) ππππ 3(πΎπ€) 4πππ ππππ (6.38) while in the limit that π → ∞, equation (6.37) reduces to: πΌπ = ππΎ 3(2 + πΎπ€)2 , 4πππ πΎπ€(2 + 3πΎπ€) (6.39) which is in complete agreement with equation (6.18) for the simple constant square pulse as expected. In comparing the general equation (6.37) for the simple square pulse with the general equation for the square pulse derived previously in equation (5.39): (1 + π πΎπ€ (πΎπ€ − 1)) ππΎ πΌπ = ( )+ πππ (8 + π −πΎπ€ + 6πΎπ€ + π πΎπ€ (4πΎπ€ − 9)) (6.40) 2(3 + πΎπ€ + π πΎπ€ (2πΎπ€ − 3)) 1 ( ), ππππ (8 + π −πΎπ€ + 6πΎπ€ + π πΎπ€ (4πΎπ€ − 9)) we see immediately that the simple square pulse completely removes all of the exponential terms that were present in the prior result. In place of the exponential terms in (6.40), the leading terms in (6.37), describing the dependence of the critical amplitude on the pulse width, are now quadratic in πΎπ€. This difference has its root in our choice to replace the hyperbolic cosine term in equation (6.3) with a constant term for π1 in equation (6.5) at the 100 beginning of this section. The final result in equation (6.37) is somewhat less complicated than equation (6.40), and it indeed appears to capture the important qualitative features of the latter equation as desired. 101 CONCLUSION From the prior considerations, we have emerged with a methodology that can generate an approximate analytical threshold solution for a one-dimensional Fitzhugh-Nagumo system. The method can be modified by choosing a different pulse shape as we did in going from a Gaussian pulse shape to a square pulse shape; the method can be modified by choosing to initiate the pulse via initial conditions on the potential or by applying an external current where the threshold criteria is different for each case; and finally, the method can be modified to account for the amount of time that the external current is applied. What has been done to this point represents only the first few steps that need to be taken towards the ultimate goal of producing a threshold equation that can accurately model cardiac systems. The next step involves generalizing the method to two- and threedimensions. After that, the source terms in the Fitzhugh-Nagumo equation will have to be replaced with more physiologically accurate expressions that better represent those found in real cardiac systems. In particular, the dependence on calcium will have to be introduced into the model at some point as calcium plays a crucial role in the onset of cardiac action potentials and the associated arrhythmias [15]. Nevertheless, the threshold solution that we have obtained at present offers an accurate analyzable solution to the question of whether or not an action potential will fire and propagate as a wave along a one-dimensional chain of cells in the Fitzhugh-Nagumo system. That is, given the initial size of the pulse width, the time that the pulse is applied, and the values of the various conduction parameters relevant to the problem, we can predict with considerable accuracy whether or not an excitation wave will appear when we observe the numerical solution to the full partial differential equation in the Fitzhugh-Nagumo 102 system. Although the Fitzhugh-Nagumo system is indeed much simpler than a real physiological system, these results at least provide us with the hope that an analytical solution can be found which describes the threshold behavior of the more complicated systems found in nature. 103 REFERENCES 1. Hodgkin AL and Huxley AF. A quantitative description of membrane current and its application to excitation and conduction in nerve. J Physiol 117: 500-544, 1952. 2. Fitzhugh R. Impulses and physiological states in theoretical models of the nerve membrane. Biophysical Journal 1: 445-466, 1961. 3. Nagumo J, Arimoto S, and Yoshizawa S. An active pulse transmission line simulating nerve axon. Proceedings of the IRE 50: 2061-2070, 1962. 4. McKean HP and Moll V. Stabilization to the standing wave in a simple caricature of the nerve equation. Comm Pure and Appl Math 39: 485-529, 1986. 5. Neu JC, Preissig Jr. RS, and Krassowska, W. Initiation of propagation in a onedimensional excitable medium. Physica D 102: 285-299, 1997. 6. Ermentrout GB and Terman DH. Mathematical Foundations of Neuroscience. Springer, 2010. 7. Keener J and Sneyd J. Mathematical Physiology I: Cellular Physiology. Second Edition. Springer, 2009. 8. Lapicque L, Brunel N, and van Rossum MCW, Trans. Quantitative investigations of electrical nerve excitation treated as polarization. Biol Cybern 2007. 9. Fitzhugh R. Thresholds and Plateaus in the Hodgkin-Huxley Nerve Equations. J Gen Physiol 43: 867-896, 1960. 10. Moll V and Rosencrans SI. Calculation of the Threshold Surface for Nerve Equations. SIAM J Appl Math 50: 1419-1441, 1990. 11. Moll V. Polygonal Approximation to the flow on the critical surface for the bistable equation. Computers Math Appl 25: 45-51, 1993. 104 12. Polking J. dfield and pplane: the java versions. math.rice.edu/~dfield/dfpp.html, 2002. 13. Boyce WE and DiPrima RC. Elementary Differential Equations. Sixth Edition. John Wiley and Sons Inc., 1997. 14. Strang G. Mathematical Methods for Engineers II: Heat Equation. http://www.cosmolearning.com/video-lectures/heat-equation/, Cosmo Learning, 2013. 15. Kleber AG and Rudy Y. Basic Mechanisms of Cardiac Impulse Propagation and Associated Arrhythmias. Physiol Rev 84: 431-488, 2004. 105 APPENDIX: NUMERICAL METHODS The method of projected dynamics produces an approximate analytical solution for the threshold of excitation that must be compared with the threshold solution obtained by directly solving the appropriate partial differential equation. This direct solution is generated using numerical methods, and here we use a finite difference approximation to solve the following partial differential equation: ππ£ ∂2 π£ = π· 2 − ππ ′ (π£), ππ‘ ∂π₯ π ′ (π£) = π£(πΌ − π£) πππ 0 < πΌ < (A.1) 1 , πΌ βͺ 1. 2 (A.2) The simplest finite difference approach uses a forward difference approximation for ππ£⁄ππ‘ and a centered difference approximation for π 2 π£/ππ₯ 2 [14]. Applying these approximations to (A.1), we get: π£π,π+1 − π£π,π βπ‘ βπ‘ π£ βπ₯ 2 π£ = π· − ππΌπ£ + ππ£ 2 , βπ‘ βπ₯ 2 π£π+1,π − 2π£π,π + π£π−1,π 2 = π· − ππΌπ£π,π + ππ£π,π . βπ₯ 2 (A.3) (A.4) In (A.4), π indexes the time, and π indexes the position coordinate which we usually refer to as the “cell number” to maintain some connection to the biology. (A.4) can be rearranged to: π£π,π+1 = π£π,π + βπ‘ (π· π£π+1,π − 2π£π,π + π£π−1,π 2 − ππΌπ£π,π + ππ£π,π ). βπ₯ 2 (A.5) The crucial ratio in (A.5) is βπ‘/βπ₯ 2 which must be kept small. Typical values used were βπ‘ = 0.01 , π = 10,000, and βπ₯ = 1.0 for the majority of the numerical simulations. π was usually indexed from 1 to 122, although we typically limited our focus to a 50 cell 106 chain length starting with a current pulse centered at π = 50 and extending to π = 109. The extra cells were added to avoid end effects at the boundaries where we used the following non-flux boundary conditions: π£(1) = π£(3), (A.6) π£(ππππ₯ ) = π£(ππππ₯ − 2), (A.7) where ππππ₯ is the highest indexed number in the range of π (for example, in this case ππππ₯ is 122). We now turn our attention to the different approaches that were used for the applied current pulse. For the Gaussian Pulse, we used the following voltage profile as the initial condition of the potential: 2 π£0 (π) = π0 π −π(πΏπ−50) . (A.8) In (A.8), π0 is the initial amplitude of the pulse, and π is related to the initial pulse width π€ as: π= 1 . 2π€ 2 (A.9) The index π refers to the cell number when the constant πΏ is set to 1, and the 50 indicates that the current pulse has been centered on cell 50. The approach for the square pulse specified using the initial conditions is much the same except the initial voltage profile is replaced with a constant initial amplitude π0 spread over a specified pulse width π€ that is chosen to be within the 50 cell range that we are observing: π£0 (π) = π0 (50: 49 + π€), π£0 (π) = 0 (πππ πππ¦ π ≠ 50: 49 + π€). 107 (A.10) (A.11) In both of the above cases, (A.8), (A.10), and (A.11) specify the initial conditions on the potential for a system in the form of equation (A.4). When the approach is changed to applying a constant square pulse, the governing equation changes to: ππ,π+1 = ππ,π + βπ‘ (π· ππ+1,π − 2ππ,π + ππ−1,π 2 − ππΎ ππ,π + πππ ππ,π + πΌπ ). βπ₯ 2 (A.12) The constant square pulse πΌ is given by: πΌ(π) = πΌ(50: 49 + π€), πΌ(π) = 0 (A.13) (πππ πππ¦ π ≠ 50: 49 + π€). (A.14) Unlike in the prior cases, equations (A.13) and (A.14) are not specified as initial conditions here. Rather the applied pulse is given as a constant that always remains present in the differential equation even as the system evolves in time. This is to simulate an external electrode providing a continuously applied current. For the time-dependent square pulse, πΌ(π‘), we simply add a time condition to equations (A.13) and (A.14): (πππ π‘ ≤ π), (A.15) πππ πππ¦ (π ≠ 50: 49 + π€)π€βππ π‘ ≤ π, ). πππ πππ πππ π π€βππ π‘ > π (A.16) πΌ(π) = πΌ(50: 49 + π€) πΌ(π) = 0 ( Here π represents the time at which the applied square current pulse is switched off. 108