12.07.2015 Views

Dynamical Systems in Neuroscience:

Dynamical Systems in Neuroscience:

Dynamical Systems in Neuroscience:

SHOW MORE
SHOW LESS
  • No tags were found...

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Eugene M. IzhikevichThe <strong>Neuroscience</strong>s Institute<strong>Dynamical</strong> <strong>Systems</strong> <strong>in</strong> <strong>Neuroscience</strong>:The Geometry of Excitability and Burst<strong>in</strong>gDecember 19, 2005The MIT Press


4.3.1 Bistability and attraction doma<strong>in</strong>s . . . . . . . . . . . . . . . . 1134.3.2 Stable/unstable manifolds . . . . . . . . . . . . . . . . . . . . . 1144.3.3 Homocl<strong>in</strong>ic/heterocl<strong>in</strong>ic trajectories . . . . . . . . . . . . . . . . 1164.3.4 Saddle-node bifurcation . . . . . . . . . . . . . . . . . . . . . . 1164.3.5 Andronov-Hopf bifurcation . . . . . . . . . . . . . . . . . . . . . 122Summary and Bibliographical Notes . . . . . . . . . . . . . . . . . . . . . . 124Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1275 Conductance-Based Models and Their Reductions 1335.1 M<strong>in</strong>imal Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1335.1.1 Amplify<strong>in</strong>g and resonant gat<strong>in</strong>g variables . . . . . . . . . . . . . 1355.1.2 I Na,p +I K -model . . . . . . . . . . . . . . . . . . . . . . . . . . . 1385.1.3 I Na,t -model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1395.1.4 I Na,p +I h -model . . . . . . . . . . . . . . . . . . . . . . . . . . . 1425.1.5 I h +I Kir -model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1445.1.6 I K +I Kir -model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1455.1.7 I A -model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1485.1.8 Ca 2+ -gated m<strong>in</strong>imal models . . . . . . . . . . . . . . . . . . . . 1525.2 Reduction of multi-dimensional models . . . . . . . . . . . . . . . . . . 1555.2.1 Hodgk<strong>in</strong>-Huxley model . . . . . . . . . . . . . . . . . . . . . . . 1555.2.2 Equivalent potentials . . . . . . . . . . . . . . . . . . . . . . . . 1585.2.3 Nullcl<strong>in</strong>es and I-V record<strong>in</strong>gs . . . . . . . . . . . . . . . . . . . 1585.2.4 Reduction to simple model . . . . . . . . . . . . . . . . . . . . . 161Summary and Bibliographical Notes . . . . . . . . . . . . . . . . . . . . . . 163Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1646 Bifurcations 1676.1 Equilibrium (Rest State) . . . . . . . . . . . . . . . . . . . . . . . . . . 1676.1.1 Saddle-node (fold) . . . . . . . . . . . . . . . . . . . . . . . . . 1706.1.2 Saddle-node on <strong>in</strong>variant circle . . . . . . . . . . . . . . . . . . 1736.1.3 Supercritical Andronov-Hopf . . . . . . . . . . . . . . . . . . . . 1776.1.4 Subcritical Andronov-Hopf . . . . . . . . . . . . . . . . . . . . . 1816.2 Limit Cycle (Spik<strong>in</strong>g State) . . . . . . . . . . . . . . . . . . . . . . . . 1866.2.1 Saddle-node on <strong>in</strong>variant circle . . . . . . . . . . . . . . . . . . 1886.2.2 Supercritical Andronov-Hopf . . . . . . . . . . . . . . . . . . . . 1896.2.3 Fold limit cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . 1906.2.4 Homocl<strong>in</strong>ic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1946.3 Other Interest<strong>in</strong>g Cases . . . . . . . . . . . . . . . . . . . . . . . . . . . 1996.3.1 Three-dimensional phase space . . . . . . . . . . . . . . . . . . 1996.3.2 Cusp and pitchfork . . . . . . . . . . . . . . . . . . . . . . . . . 2016.3.3 Bogdanov-Takens . . . . . . . . . . . . . . . . . . . . . . . . . . 2026.3.4 Relaxation oscillators and Canards . . . . . . . . . . . . . . . . 2076.3.5 Baut<strong>in</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209


6.3.6 Saddle-node homocl<strong>in</strong>ic orbit . . . . . . . . . . . . . . . . . . . 2106.3.7 Hard and soft loss of stability . . . . . . . . . . . . . . . . . . . 213Summary and Bibliographical Notes . . . . . . . . . . . . . . . . . . . . . . 214Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2207 Neuronal Excitability 2257.1 Excitability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2257.1.1 Bifurcations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2267.1.2 Hodgk<strong>in</strong>’s classification . . . . . . . . . . . . . . . . . . . . . . . 2287.1.3 Classes 1 and 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . 2317.1.4 Class 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2327.1.5 Ramps, steps, and shocks . . . . . . . . . . . . . . . . . . . . . 2347.1.6 Bistability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2367.1.7 Class 1 and 2 spik<strong>in</strong>g . . . . . . . . . . . . . . . . . . . . . . . . 2387.2 Integrators vs. Resonators . . . . . . . . . . . . . . . . . . . . . . . . . 2397.2.1 Fast subthreshold oscillations . . . . . . . . . . . . . . . . . . . 2407.2.2 Frequency preference and resonance . . . . . . . . . . . . . . . . 2427.2.3 Frequency preference <strong>in</strong> vivo . . . . . . . . . . . . . . . . . . . . 2477.2.4 Thresholds and action potentials . . . . . . . . . . . . . . . . . 2497.2.5 Threshold manifolds . . . . . . . . . . . . . . . . . . . . . . . . 2507.2.6 Rheobase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2527.2.7 Post-<strong>in</strong>hibitory spike . . . . . . . . . . . . . . . . . . . . . . . . 2537.2.8 Inhibition-<strong>in</strong>duced spik<strong>in</strong>g . . . . . . . . . . . . . . . . . . . . . 2557.2.9 Spike latency . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2577.2.10 Flipp<strong>in</strong>g from an <strong>in</strong>tegrator to a resonator . . . . . . . . . . . . 2597.2.11 Transition between <strong>in</strong>tegrators and resonators . . . . . . . . . . 2617.3 Slow Modulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2647.3.1 Spike-frequency modulation . . . . . . . . . . . . . . . . . . . . 2657.3.2 I-V relation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2677.3.3 Slow Subthreshold oscillation . . . . . . . . . . . . . . . . . . . 2707.3.4 Rebound response and voltage sag . . . . . . . . . . . . . . . . 2707.3.5 AHP and ADP . . . . . . . . . . . . . . . . . . . . . . . . . . . 272Summary and Bibliographical Notes . . . . . . . . . . . . . . . . . . . . . . 274Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2768 Simple Models 2798.1 Simplest Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2798.1.1 Integrate-and-fire . . . . . . . . . . . . . . . . . . . . . . . . . . 2808.1.2 Resonate-and-fire . . . . . . . . . . . . . . . . . . . . . . . . . . 2818.1.3 Quadratic <strong>in</strong>tegrate-and-fire . . . . . . . . . . . . . . . . . . . . 2828.1.4 Simple model of choice . . . . . . . . . . . . . . . . . . . . . . . 2838.1.5 Canonical models . . . . . . . . . . . . . . . . . . . . . . . . . . 2898.2 Cortex . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 292


8.2.1 Regular spik<strong>in</strong>g (RS) neurons . . . . . . . . . . . . . . . . . . . 2958.2.2 Intr<strong>in</strong>sically burst<strong>in</strong>g (IB) neurons . . . . . . . . . . . . . . . . . 3018.2.3 Multi-compartment dendritic tree . . . . . . . . . . . . . . . . . 3058.2.4 Chatter<strong>in</strong>g (CH) neurons . . . . . . . . . . . . . . . . . . . . . . 3078.2.5 Low-threshold spik<strong>in</strong>g (LTS) <strong>in</strong>terneurons . . . . . . . . . . . . 3088.2.6 Fast spik<strong>in</strong>g (FS) <strong>in</strong>terneurons . . . . . . . . . . . . . . . . . . . 3118.2.7 Late spik<strong>in</strong>g (LS) <strong>in</strong>terneurons . . . . . . . . . . . . . . . . . . . 3138.2.8 Diversity of <strong>in</strong>hibitory <strong>in</strong>terneurons . . . . . . . . . . . . . . . . 3148.3 Thalamus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3158.3.1 Thalamo-cortical (TC) relay neurons . . . . . . . . . . . . . . . 3168.3.2 Reticular thalamic nucleus (RTN) neurons . . . . . . . . . . . . 3178.3.3 Thalamic <strong>in</strong>terneurons . . . . . . . . . . . . . . . . . . . . . . . 3178.4 Other <strong>in</strong>terest<strong>in</strong>g cases . . . . . . . . . . . . . . . . . . . . . . . . . . . 3188.4.1 Hippocampal CA1 pyramidal neurons . . . . . . . . . . . . . . . 3188.4.2 Sp<strong>in</strong>y projection neurons of neostriatum and basal ganglia . . . 3198.4.3 Mesencephalic V neurons of bra<strong>in</strong>stem . . . . . . . . . . . . . . 3208.4.4 Stellate cells of entorh<strong>in</strong>al cortex . . . . . . . . . . . . . . . . . 3208.4.5 Mitral neurons of olfactory bulb . . . . . . . . . . . . . . . . . . 321Summary and Bibliographical Notes . . . . . . . . . . . . . . . . . . . . . . 323Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3269 Burst<strong>in</strong>g 3419.1 Electrophysiology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3419.1.1 Example: The I Na,p +I K +I K(M) -model . . . . . . . . . . . . . . . 3439.1.2 Fast-Slow Dynamics . . . . . . . . . . . . . . . . . . . . . . . . 3459.1.3 M<strong>in</strong>imal models . . . . . . . . . . . . . . . . . . . . . . . . . . . 3479.1.4 Central pattern generators and half-center oscillators . . . . . . 3519.2 Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3519.2.1 Fast-slow bursters . . . . . . . . . . . . . . . . . . . . . . . . . . 3529.2.2 Phase portraits . . . . . . . . . . . . . . . . . . . . . . . . . . . 3529.2.3 Averag<strong>in</strong>g . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3559.2.4 Equivalent voltage . . . . . . . . . . . . . . . . . . . . . . . . . 3579.2.5 Hysteresis loops and slow waves . . . . . . . . . . . . . . . . . . 3589.2.6 Bifurcations “rest<strong>in</strong>g ↔ burst<strong>in</strong>g ↔ spik<strong>in</strong>g” . . . . . . . . . . . 3609.3 Classification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3649.3.1 fold/homocl<strong>in</strong>ic . . . . . . . . . . . . . . . . . . . . . . . . . . . 3659.3.2 circle/circle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3719.3.3 subHopf/fold cycle . . . . . . . . . . . . . . . . . . . . . . . . . 3749.3.4 fold/fold cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3819.3.5 fold/Hopf . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3819.3.6 fold/circle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3819.4 Neuro-Computational Properties . . . . . . . . . . . . . . . . . . . . . 3839.4.1 How to dist<strong>in</strong>guish? . . . . . . . . . . . . . . . . . . . . . . . . . 383


9.4.2 Integrators vs. Resonators . . . . . . . . . . . . . . . . . . . . . 3849.4.3 Bistability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3869.4.4 Bursts as a unit of neuronal <strong>in</strong>formation . . . . . . . . . . . . . 3879.4.5 Synchronization . . . . . . . . . . . . . . . . . . . . . . . . . . . 389Summary and Bibliographical Notes . . . . . . . . . . . . . . . . . . . . . . 392Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39510 Synchronization (see www.izhikevich.com) 403Solutions to Exercises 407References 44110 Synchronization (see www.izhikevich.com) 45710.1 Pulsed Coupl<strong>in</strong>g . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45710.1.1 Phase of oscillation . . . . . . . . . . . . . . . . . . . . . . . . . 45810.1.2 Isochrons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45910.1.3 PRC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46010.1.4 Type 0 and 1 phase response . . . . . . . . . . . . . . . . . . . . 46310.1.5 Po<strong>in</strong>care phase map . . . . . . . . . . . . . . . . . . . . . . . . 46610.1.6 Fixed po<strong>in</strong>ts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46710.1.7 Synchronization . . . . . . . . . . . . . . . . . . . . . . . . . . . 46810.1.8 Phase lock<strong>in</strong>g . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46910.1.9 Arnold tongues . . . . . . . . . . . . . . . . . . . . . . . . . . . 47010.2 Weak Coupl<strong>in</strong>g . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47110.2.1 W<strong>in</strong>free’s approach . . . . . . . . . . . . . . . . . . . . . . . . . 47210.2.2 Kuramoto’s approach . . . . . . . . . . . . . . . . . . . . . . . . 47410.2.3 Malk<strong>in</strong>’s approach . . . . . . . . . . . . . . . . . . . . . . . . . 47510.2.4 Measur<strong>in</strong>g PRCs experimentally . . . . . . . . . . . . . . . . . . 47610.2.5 Phase model for coupled oscillators . . . . . . . . . . . . . . . . 47910.3 Synchronization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48110.3.1 Two oscillators . . . . . . . . . . . . . . . . . . . . . . . . . . . 48310.3.2 Cha<strong>in</strong>s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48510.3.3 Networks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48710.3.4 Mean-field approximations . . . . . . . . . . . . . . . . . . . . . 48810.4 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48910.4.1 Phase oscillators . . . . . . . . . . . . . . . . . . . . . . . . . . 48910.4.2 SNIC oscillators . . . . . . . . . . . . . . . . . . . . . . . . . . . 49010.4.3 Homocl<strong>in</strong>ic oscillators . . . . . . . . . . . . . . . . . . . . . . . 49610.4.4 Relaxation oscillators and FTM . . . . . . . . . . . . . . . . . . 49810.4.5 Burst<strong>in</strong>g oscillators . . . . . . . . . . . . . . . . . . . . . . . . . 500Summary and Bibliographical Notes . . . . . . . . . . . . . . . . . . . . . . 501Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 506Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 511


PrefaceHistorically, much of theoretical neuroscience research concerned neuronal circuits andsynaptic organization. The neurons were divided <strong>in</strong>to excitatory and <strong>in</strong>hibitory types,but their electrophysiological properties were largely neglected or taken to be identicalto those of Hodgk<strong>in</strong>-Huxley’s squid axon. The present awareness of the importance ofthe electrophysiology of <strong>in</strong>dividual neurons is best summarized by David McCormick <strong>in</strong>the fifth edition of Gordon Shepherd’s book “The Synaptic Organization of the Bra<strong>in</strong>”:“Information process<strong>in</strong>g depends not only on the anatomical substrates of synapticcircuits but also on the electrophysiological properties of neurons... . Even iftwo neurons <strong>in</strong> different regions of the nervous system possess identical morphologicalfeatures, they may respond to the same synaptic <strong>in</strong>put <strong>in</strong> very differentmanners because of each cell’s <strong>in</strong>tr<strong>in</strong>sic properties.”David A. McCormick (2004)Much of present neuroscience research concerns voltage- and second-messengergatedcurrents <strong>in</strong> <strong>in</strong>dividual cells with the goal to understand the cell’s <strong>in</strong>tr<strong>in</strong>sic neurocomputationalproperties. It is widely accepted that know<strong>in</strong>g the currents suffices todeterm<strong>in</strong>e what the cell is do<strong>in</strong>g and why. This, however, contradicts a half-century oldobservation that cells hav<strong>in</strong>g similar currents can still exhibit quite different dynamics.Indeed, study<strong>in</strong>g isolated axons hav<strong>in</strong>g presumably similar electrophysiology (all arefrom crustacean Carc<strong>in</strong>us maenas), Hodgk<strong>in</strong> (1948) <strong>in</strong>jected a dc-current of vary<strong>in</strong>gamplitude, and discovered that some preparations could exhibit repetitive spik<strong>in</strong>g witharbitrarily low frequencies, while the others discharged <strong>in</strong> a narrow frequency band.This observation was largely ignored by the neuroscience community until the sem<strong>in</strong>alpaper by R<strong>in</strong>zel and Ermentrout (1989), who showed that the difference <strong>in</strong> behavior isdue to different bifurcation mechanisms of excitability.Let us treat the amplitude of the <strong>in</strong>jected current <strong>in</strong> Hodgk<strong>in</strong>’s experiments as abifurcation parameter: When the amplitude is small, the cell is quiescent; when theamplitude is large, the cell fires repetitive spikes. When we change the amplitude of the<strong>in</strong>jected current, the cell undergoes a transition from quiescence to repetitive spik<strong>in</strong>g.From the dynamical systems po<strong>in</strong>t of view the transition corresponds to a bifurcationfrom equilibrium to a limit cycle attractor. The type of bifurcation determ<strong>in</strong>es the mostfundamental computational properties of neurons, such as the class of excitability, theexistence or non-existence of threshold, all-or-none spikes, subthreshold oscillations,the ability to generate post-<strong>in</strong>hibitory rebound spikes, bistability of rest<strong>in</strong>g and spik<strong>in</strong>gstates, whether the neuron is an <strong>in</strong>tegrator or resonator, etc.This book is devoted to a systematic study of the relationship between electrophysiology,bifurcations, and computational properties of neurons. The reader willlearn why cells hav<strong>in</strong>g nearly identical currents may undergo dist<strong>in</strong>ct bifurcations, andhence they will have fundamentally different neuro-computational properties. (Conix


xPrefaceversely, cells hav<strong>in</strong>g quite different currents may undergo identical bifurcations, andhence they will have similar neuro-computational properties.) The major messageof the book can be summarized as follows (compare with the McCormick statementabove):Information process<strong>in</strong>g depends not only on the electrophysiological propertiesof neurons but also on their dynamical properties. Even if two neurons <strong>in</strong> thesame region of the nervous system possess similar electrophysiological features,they may respond to the same synaptic <strong>in</strong>put <strong>in</strong> very different manners becauseof each cell’s bifurcation dynamics.Non-l<strong>in</strong>ear dynamical system theory is a core of the computational neuroscience research,but it is not a standard part of the graduate neuroscience curriculum. Neitheris it taught <strong>in</strong> most math/physics departments <strong>in</strong> a form suitable for a general biologicalaudience. As a result, many neuroscientists fail to grasp such fundamentalconcepts as equilibrium, stability, limit cycle attractor, and bifurcations, even thoughneuroscientists encounter these non-l<strong>in</strong>ear phenomena constantly.This book <strong>in</strong>troduces dynamical systems start<strong>in</strong>g with simple one- and two-dimensionalspik<strong>in</strong>g models and cont<strong>in</strong>u<strong>in</strong>g all the way to burst<strong>in</strong>g systems. Each chapteris organized “from simple to complex”, so everybody can start read<strong>in</strong>g the book; thereader’s background would only determ<strong>in</strong>e where he or she stops. The book emphasizesthe geometrical approach, so there are few equations but a lot of figures. Half of themare simulations of various neural models, so there are hundreds of possible exercisessuch as “Use MATLAB (GENESIS, NEURON, XPPAUT, etc.) and parameters <strong>in</strong> thecaption of Fig. X to simulate the figure”. Additional homework problems are providedat the end of each chapter; the reader is encouraged to solve at least some of them andlook at the solutions of the others at the end of the book. Problems marked [M.S.] or[Ph.D.] are suggested thesis topics.Acknowledgment. The author thanks all scientists who reviewed the first draftof the book: Pablo Achard, Jose M. Amigo, Brent Doiron, George Bard Ermentrout,Richard FitzHugh, David Golomb, Andrei Iacob, Maciej Lazarewicz, GeorgiMedvedev, John R<strong>in</strong>zel, Anil K. Seth, Gautam C Sethia, Arthur Sherman, Klaus M.Stiefel, Takashi Tateno. The author thanks anonymous referees who peer-reviewed thebook and made quite a few valuable suggestions <strong>in</strong>stead of just reject<strong>in</strong>g it. Specialthanks are to Niraj S. Desai who made most of the <strong>in</strong> vitro record<strong>in</strong>gs used <strong>in</strong> the book(the data are available on the author’s webpage www.izhikevich.com) and to Brunovan Sw<strong>in</strong>deren who drew the caricatures. The author has enjoyed the hospitality ofThe <strong>Neuroscience</strong>s Institute — a monastery of <strong>in</strong>terdiscipl<strong>in</strong>ary science, and he hasbenefited greatly from the expertise and support of its fellows.F<strong>in</strong>ally, the author thanks his wife Tatyana and wonderful daughters Elizabeth andKate for their support and patience dur<strong>in</strong>g the four-year gestation of this book.Eugene M. Izhikevichwww.izhikevich.comSan Diego, California December 19, 2005


Chapter 1IntroductionThis chapter highlights some of the most important concepts developed <strong>in</strong> the book.First, we discuss several common misconceptions regard<strong>in</strong>g the spike-generation mechanismof neurons. Our goal is to motivate the reader <strong>in</strong>to th<strong>in</strong>k<strong>in</strong>g of a neuron notonly <strong>in</strong> terms of ions and channels, as many biologists do, and not only <strong>in</strong> terms of<strong>in</strong>put/output relationship, as many theoreticians do, but also as a nonl<strong>in</strong>ear dynamicalsystem that looks at the <strong>in</strong>put through the prism of its own <strong>in</strong>tr<strong>in</strong>sic dynamics. Weask such questions as “what makes a neuron fire?” or “where is the threshold”, andthen outl<strong>in</strong>e the answers us<strong>in</strong>g geometrical theory of dynamical systems.From a dynamical systems po<strong>in</strong>t of view, neurons are excitable because they arenear a transition, called bifurcation, from rest<strong>in</strong>g to susta<strong>in</strong>ed spik<strong>in</strong>g activity. Whilethere is a huge number of possible ionic mechanisms of excitability and spike-generation,there are only four different bifurcation mechanisms that can result <strong>in</strong> such a transition.Consider<strong>in</strong>g the geometry of phase portraits at these bifurcations, we can understandmany computational properties of neurons, such as the nature of threshold and all-ornonespik<strong>in</strong>g, the co-existence of rest<strong>in</strong>g and spik<strong>in</strong>g states, the orig<strong>in</strong> of spike latencies,post-<strong>in</strong>hibitory spikes, the mechanism of <strong>in</strong>tegration and resonance, etc. Moreover, wecan understand how these properties are <strong>in</strong>terrelated, why some are equivalent andsome are mutually exclusive.1.1 NeuronsIf somebody were to put a gun to the head of the author of this book and ask himto name the s<strong>in</strong>gle most important concept <strong>in</strong> bra<strong>in</strong> science, he would say it is theconcept of a neuron. There are only 10 11 or so neurons <strong>in</strong> the human bra<strong>in</strong>, muchfewer than the number of non-neural cells such as glia. Yet neurons are unique <strong>in</strong> thesense that only they can transmit electrical signals over long distances. From neuronallevel we can go down to cell biophysics, to molecular biology of gene regulation, etc.From neuronal level we can go up to neuronal circuits, to cortical structures, to thewhole bra<strong>in</strong>, and f<strong>in</strong>ally to the behavior of the organism. So, let us see how much weunderstand of what is go<strong>in</strong>g on at the level of <strong>in</strong>dividual neurons.1


2 Introductionapical dendritessomarecord<strong>in</strong>gelectrodebasal dendritesmembrane potential, mV+35 mVspike40 ms-60 mVtime, mssynapse0.1 mmaxonFigure 1.1: Two <strong>in</strong>terconnected cortical pyramidal neurons (hand draw<strong>in</strong>g) and <strong>in</strong> vitrorecorded spike.1.1.1 What is a spike?A typical neuron receives <strong>in</strong>puts from more than 10, 000 other neurons through the contactson its dendritic tree called synapses; see Fig. 1.1. The <strong>in</strong>puts produce electricaltransmembrane currents that change the membrane potential of the neuron. Synapticcurrents produce changes, called post-synaptic potentials (PSPs). Small currents producesmall PSPs; larger currents produce significant PSPs that could be amplified bythe voltage-sensitive channels embedded <strong>in</strong> neuronal membrane and lead to the generationof an action potential or spike – an abrupt and transient change of membranevoltage that propagates to other neurons via a long protrusion called an axon.Such spikes are the ma<strong>in</strong> means of communication between neurons. In general,neurons do not fire on their own, they do it as a result of the <strong>in</strong>com<strong>in</strong>g spikes from otherneurons. One of the most fundamental question of neuroscience is what exactly makesneurons fire? What is it <strong>in</strong> the <strong>in</strong>com<strong>in</strong>g pulses that elicits a response <strong>in</strong> one neuronbut not <strong>in</strong> another one? Why could two neurons have different responses to exactlythe same <strong>in</strong>put and identical responses to completely different <strong>in</strong>puts? To answerthese questions, we need to understand the dynamics of spike-generation mechanismsof neurons.Most <strong>in</strong>troductory neuroscience books describe neurons as <strong>in</strong>tegrators with a threshold:Neurons sum up <strong>in</strong>com<strong>in</strong>g PSPs and “compare” the <strong>in</strong>tegrated PSP with a certa<strong>in</strong>voltage value, called fir<strong>in</strong>g threshold. If it is below the threshold, the neuron rema<strong>in</strong>squiescent; when it is above the threshold, the neuron fires an all-or-none spike, as <strong>in</strong>Fig. 1.3, and resets its membrane potential. To add theoretical plausibility to thisargument, the books refer to the Hodgk<strong>in</strong>-Huxley model of spike-generation <strong>in</strong> squid


Introduction 3Figure 1.2: What makes a neuron fire?giant axons, which we study <strong>in</strong> the next chapter. The irony is that the Hodgk<strong>in</strong>-Huxleymodel does not have a well-def<strong>in</strong>ed threshold, it does not fire all-or-none spikes, andit is not an <strong>in</strong>tegrator, but a resonator, i.e., it prefers <strong>in</strong>puts hav<strong>in</strong>g certa<strong>in</strong> frequenciesthat resonate with the frequency of subthreshold oscillations of the neuron. Weconsider these and other properties <strong>in</strong> detail <strong>in</strong> this book.1.1.2 Where is the threshold?Much effort has been spent try<strong>in</strong>g to determ<strong>in</strong>e experimentally the fir<strong>in</strong>g thresholdsof neurons. Here, we challenge the classical view of a threshold. Let us consider twotypical experiments, depicted <strong>in</strong> Fig. 1.4, that are designed to measure the threshold.On the left, we shock a cortical neuron, i.e., we <strong>in</strong>ject brief but strong pulses of currentof various amplitudes to depolarize the membrane potential to various values. Is therea clear-cut voltage value, as <strong>in</strong> Fig. 1.3, above which the neuron fires but below whichno spikes occur? If you f<strong>in</strong>d one, let the author know! In Fig. 1.4b we <strong>in</strong>ject long butweak pulses of current of various amplitudes, which result <strong>in</strong> slow depolarization anda spike. The fir<strong>in</strong>g threshold, if it exists, must be somewhere <strong>in</strong> the shaded region, butwhere? Where does the slow depolarization end and the spike start? Is it mean<strong>in</strong>gfulto talk about fir<strong>in</strong>g thresholds at all?all-or-nonespikesthresholdrest<strong>in</strong>gno spikeFigure 1.3: The concept of a fir<strong>in</strong>g threshold.


4 Introduction(a)spikes(b)spikes cutthreshold?-40 mVsubthreshold response1 ms<strong>in</strong>jected pulses of current20 mV<strong>in</strong>jected pulses of current15 msFigure 1.4: Where is the fir<strong>in</strong>g threshold? Shown are <strong>in</strong> vitro record<strong>in</strong>gs of two layer 5pyramidal neurons of rat. Notice the difference of voltage and time scales.(a)(b)20 mV20 mV5 ms100 ms-60 mVFigure 1.5: Where is the rheobase, i.e., the m<strong>in</strong>imal current that fires the cell? (a) <strong>in</strong>vitro record<strong>in</strong>gs of pyramidal neuron of layer 2/3 of rat’s visual cortex show <strong>in</strong>creas<strong>in</strong>glatencies as the amplitude of the <strong>in</strong>jected current decreases. (b) Simulation of theI Na,p +I K -model shows spikes of graded amplitude.Perhaps, we should measure current thresholds <strong>in</strong>stead of voltage thresholds? Thecurrent threshold, i.e., the m<strong>in</strong>imal amplitude of <strong>in</strong>jected current of <strong>in</strong>f<strong>in</strong>ite durationneeded to fire a neuron, is called rheobase. In Fig. 1.5 we decrease the amplitudesof <strong>in</strong>jected pulses of current to f<strong>in</strong>d the m<strong>in</strong>imal one that still elicits a spike or themaximal one that does not. In Fig. 1.5a, progressively weaker pulses result <strong>in</strong> longerlatencies to the first spike. Eventually the neuron does not fire because the latency islonger than the duration of the pulse, which is 1 second <strong>in</strong> the figure. Did we reallymeasure the neuronal rheobase? What if we waited a bit longer? How long is longenough? In Fig. 1.5b the latencies do not grow but the spike amplitudes decrease untilthe spikes do not look like spikes at all. To determ<strong>in</strong>e the current threshold, we needto draw the l<strong>in</strong>e and separate spike responses from “subthreshold” ones. How can wedo that if the spikes are not all-or-none? Is the response denoted by the dashed l<strong>in</strong>e aspike?Risk<strong>in</strong>g add<strong>in</strong>g more confusion to the notion of a threshold, consider the follow<strong>in</strong>g:If excitatory <strong>in</strong>puts depolarize the membrane potential, i.e., br<strong>in</strong>g it closer to the “fir<strong>in</strong>gthreshold”, and <strong>in</strong>hibitory <strong>in</strong>puts hyperpolarize the potential and move it away from


Introduction 510 ms10 mV-45 mV0 pA-100 pAFigure 1.6: In vitro record<strong>in</strong>g of rebound spikes ofrats bra<strong>in</strong>stem mesV neuron <strong>in</strong> response to a briefhyperpolariz<strong>in</strong>g pulse of current.10ms5msnon-resonant burst10msresonant burst15msnon-resonant burst<strong>in</strong>hibitory burstFigure 1.7: Resonant response of the mesencephalic V neuron of rat bra<strong>in</strong>stem to pulsesof <strong>in</strong>jected current hav<strong>in</strong>g 10 ms period (<strong>in</strong> vitro).the threshold, then how can the neuron <strong>in</strong> Fig. 1.6 fire <strong>in</strong> response to the <strong>in</strong>hibitory<strong>in</strong>put? This phenomenon is also observed <strong>in</strong> the Hodgk<strong>in</strong>-Huxley model, and it iscalled anodal break excitation, rebound spike, or post-<strong>in</strong>hibitory spike. Many biologistssay that rebound responses are due to the activation and <strong>in</strong>activation of certa<strong>in</strong> slowcurrents, which br<strong>in</strong>g the membrane potential over the threshold, or equivalently, lowerthe threshold upon release from the hyperpolarization – a phenomenon called a lowthresholdspike <strong>in</strong> thalamocortical neurons. The problem with this explanation is thatneither the Hodgk<strong>in</strong>-Huxley model nor the neuron <strong>in</strong> the figure have these currents,and even if they did, the hyperpolarization is too short and too weak to affect thecurrents.Another <strong>in</strong>terest<strong>in</strong>g phenomenon is depicted <strong>in</strong> Fig. 1.7. The neuron is stimulatedwith brief pulses of current mimick<strong>in</strong>g an <strong>in</strong>com<strong>in</strong>g burst of three spikes. When thestimulation frequency is high (5 ms period), presumably reflect<strong>in</strong>g a strong <strong>in</strong>put,the neuron does not fire at all. However, stimulation with a lower frequency (10ms period) that resonates with the frequency of subthreshold oscillation of the neuronevokes a spike response, regardless of whether the stimulation is excitatory or <strong>in</strong>hibitory.Stimulation with even lower frequency (15 ms period) cannot elicit spike response aga<strong>in</strong>.Thus, the neuron is sensitive only to the <strong>in</strong>puts hav<strong>in</strong>g resonant frequency. The samepulses applied to a cortical pyramidal neuron evoke a response only <strong>in</strong> the first case(small period), but not <strong>in</strong> the other cases.


6 Introduction1.1.3 Why are neurons different and why do we care?Why would two neurons respond completely differently to the same <strong>in</strong>put? A biologistwould say that the response of a neuron depends on many factors, such as the typeof voltage- and Ca 2+ -gated channels expressed by the neuron, the morphology of itsdendritic tree, the location of the <strong>in</strong>put, etc. These factors are <strong>in</strong>deed important, butthey do not determ<strong>in</strong>e the neuronal response per se. They rather determ<strong>in</strong>e the rulesthat govern dynamics of the neuron. Different conductances and currents can result <strong>in</strong>the same rules and hence <strong>in</strong> the same responses, and conversely, similar currents canresult <strong>in</strong> different rules and <strong>in</strong> different responses. The currents def<strong>in</strong>e what k<strong>in</strong>d of adynamical system the neuron is.We study ionic transmembrane currents <strong>in</strong> the next chapter. In subsequent chapterswe <strong>in</strong>vestigate how the type of currents determ<strong>in</strong>e neuronal dynamics. We divide allcurrents <strong>in</strong>to two major classes: amplify<strong>in</strong>g and resonant, with persistent Na + currentI Na,p and persistent K + current I K be<strong>in</strong>g the typical examples of the former and thelatter. S<strong>in</strong>ce there are tens of known currents, purely comb<strong>in</strong>atorial argument impliesthat there are millions of different electrophysiological mechanisms of spike generation.We will show later that any such mechanism must have at least one amplify<strong>in</strong>g and oneresonant current. Some mechanisms, called m<strong>in</strong>imal <strong>in</strong> this book, have precisely oneresonant and one amplify<strong>in</strong>g current. They provide an <strong>in</strong>valuable tool <strong>in</strong> classify<strong>in</strong>gand understand<strong>in</strong>g the electrophysiology of spike-generation.Many illustrations <strong>in</strong> this book are based on simulations of the reduced I Na,p + I K -model, which consists of fast persistent Na + (amplify<strong>in</strong>g) current and slower persistentK + (resonant) current. It is equivalent to the famous and widely used Morris-LecarI Ca + I K -model (Morris and Lecar 1981). We show that the model exhibits quite differentdynamics depend<strong>in</strong>g on the values of the parameters, e.g., the half-activationvoltage of the K + current: In one case, it can fire <strong>in</strong> a narrow frequency range, exhibitco-existence of rest<strong>in</strong>g and spik<strong>in</strong>g states, damped subthreshold oscillations of membranepotential, etc. In another case, it can fire <strong>in</strong> a wide frequency range and show noco-existence of rest<strong>in</strong>g and spik<strong>in</strong>g and no subthreshold oscillations. Thus, seem<strong>in</strong>gly<strong>in</strong>essential differences <strong>in</strong> parameter values could result <strong>in</strong> drastically dist<strong>in</strong>ct behaviors.1.1.4 Build<strong>in</strong>g modelsTo build a good model of a neuron, electrophysiologists apply different pharmacologicalblockers to tease out the currents that the neuron has. Then, they apply differentstimulation protocols to measure the k<strong>in</strong>etic parameters of the currents, such as theBoltzmann activation function, time constants, maximal conductances, etc. We considerall these functions <strong>in</strong> the next chapter. Then, they create a Hodgk<strong>in</strong>-Huxley-typemodel and simulate it us<strong>in</strong>g NEURON, GENESIS, XPP environments or just pla<strong>in</strong>MATLAB (the first two are <strong>in</strong>valuable tools for simulat<strong>in</strong>g realistic dendritic structures).The problem is that the parameters are measured <strong>in</strong> different neurons and then puttogether <strong>in</strong>to a s<strong>in</strong>gle model. As an illustration, consider two neurons hav<strong>in</strong>g the same


Introduction 7Figure 1.8: Neurons are dynamical systems.currents, say I Na,p and I K , and exhibit<strong>in</strong>g excitable behavior; that is, both neurons arequiescent but can fire a spike <strong>in</strong> response to a stimulation. Suppose the second neuronhas stronger I Na,p , which is balanced by stronger I K . If we measure Na + conductanceus<strong>in</strong>g the first neuron and K + conductance us<strong>in</strong>g the second neuron, the result<strong>in</strong>gI Na,p + I K -model would have an excess of K + current and probably not be able to firespikes at all. Conversely, if we measure Na + and K + conductances us<strong>in</strong>g the secondand then the first neuron, respectively, the model would have too much Na + currentand probably exhibit susta<strong>in</strong>ed pacemak<strong>in</strong>g activity. In any case, the model fails toreproduce the excitable behavior of the neurons whose parameters we measured.Some of the parameters cannot be measured at all, so many arbitrary choices aremade via a process called “f<strong>in</strong>e-tun<strong>in</strong>g”. Navigat<strong>in</strong>g <strong>in</strong> the dark, possibly with the helpof some biological <strong>in</strong>tuition, the researcher modifies parameters, compares simulationswith experiment, and repeats this trial-and-error procedure until he or she is satisfiedwith the results. S<strong>in</strong>ce seem<strong>in</strong>gly similar values of parameters can result <strong>in</strong> drasticallydifferent behaviors, and quite different parameters can result <strong>in</strong> seem<strong>in</strong>gly similar behaviors,how do we know that the result<strong>in</strong>g model is correct? How do we know that itsbehavior is equivalent to that of the neuron we want to study? And what is equivalent<strong>in</strong> this case? Now, the reader is primed to consider dynamical systems.1.2 <strong>Dynamical</strong> <strong>Systems</strong>In the next chapter we <strong>in</strong>troduce the Hodgk<strong>in</strong>-Huxley formalism to describe neuronaldynamics <strong>in</strong> terms of activation and <strong>in</strong>activation of voltage-gated conductances. Animportant consequence of the Hodgk<strong>in</strong>-Huxley studies is that neurons are dynamicalsystems, so they should be studied as such. Below we mention some of the important


8 Introductionconcepts of dynamical systems theory. The reader does not have to follow all the detailsof this section because the concepts are expla<strong>in</strong>ed <strong>in</strong> a greater detail <strong>in</strong> subsequentchapters.A dynamical system consists of a set of variables that describe its state and alaw that describes the evolution of the state variables with time, i.e., how the stateof the system <strong>in</strong> the next moment of time depends on the <strong>in</strong>put and its state <strong>in</strong> theprevious moment of time. The Hodgk<strong>in</strong>-Huxley model is a four-dimensional dynamicalsystem because its state is determ<strong>in</strong>ed uniquely by the membrane potential, V , and socalled gat<strong>in</strong>g variables n, m and h for persistent K + and transient Na + currents. Theevolution law is given by a four-dimensional system of ord<strong>in</strong>ary differential equations.Typically, all variables describ<strong>in</strong>g neuronal dynamics can be classified <strong>in</strong>to fourclasses, accord<strong>in</strong>g to their function and the time scale:1. Membrane potential.2. Excitation variables, such as activation of Na + current. This variables are responsiblefor the upstroke of the spike.3. Recovery variables, such as <strong>in</strong>activation of Na + current and activation of fast K +current. This variables are responsible for the repolarization (downstroke) of thespike.4. Adaptation variables, such as activation of slow voltage- or Ca 2+ -dependent currents.This variables build up dur<strong>in</strong>g prolonged spik<strong>in</strong>g and can affect excitabilityon the long run.The Hodgk<strong>in</strong>-Huxley model does not have variables of the fourth type, but manyneuronal models do, especially those exhibit<strong>in</strong>g burst<strong>in</strong>g dynamics.1.2.1 Phase portraitsThe power of the dynamical systems approach to neuroscience, as well as to manyother sciences, is that we can tell someth<strong>in</strong>g, or many th<strong>in</strong>gs, about a system withouteven know<strong>in</strong>g all the details that govern the system evolution. We do not even useequations to do that! Some may even wonder why we call it a mathematical theory.As a start, let us consider a quiescent neuron whose membrane potential is rest<strong>in</strong>g.From the dynamical systems po<strong>in</strong>t of view, there are no changes of the statevariables of such a neuron, hence it is at an equilibrium po<strong>in</strong>t. All the <strong>in</strong>ward currentsthat depolarize the neuron are balanced, or equilibrated, by the outward currents thathyperpolarize it. If the neuron rema<strong>in</strong>s quiescent despite small disturbances and membranenoise, as <strong>in</strong> Fig. 1.9a, top, then we conclude that the equilibrium is stable. Isn’tit amaz<strong>in</strong>g that we can make such a conclusion without know<strong>in</strong>g the equations thatdescribe the neuron’s dynamics? We do not even know the number of variables neededto describe the neuron; it could be <strong>in</strong>f<strong>in</strong>ite, for all we care.


Introduction 9K + activation gate, n membrane potential, V(t)(a) rest<strong>in</strong>g (b) excitable (c) periodic spik<strong>in</strong>gPSPstimulusequilibriummembrane potential, Vtime, tPSPA BstimuliPSPA BspikespikeperiodicorbitFigure 1.9: Rest<strong>in</strong>g, excitable, and periodic spik<strong>in</strong>g activity correspond to a stableequilibrium (a and b) or limit cycle (c), respectively.In this book we <strong>in</strong>troduce the notions of equilibria, stability, threshold, and attractiondoma<strong>in</strong>s us<strong>in</strong>g one- and two-dimensional dynamical systems, e.g., the I Na,p +I K -model with <strong>in</strong>stantaneous Na + k<strong>in</strong>etics. Its state is described by the membrane potential,V , and the activation variable, n, of the persistent K + current, so it is atwo-dimensional vector (V, n). Instantaneous activation of Na + current is a functionof V , so it does not result <strong>in</strong> a separate variable of the model. The evolution of themodel is a trajectory (V (t), n(t)) on the V × n-plane. Depend<strong>in</strong>g on the <strong>in</strong>itial po<strong>in</strong>t,the system can have many trajectories, such as those depicted <strong>in</strong> Fig. 1.9a, bottom.Time is not present explicitly <strong>in</strong> the figure, but units of time may be thought of asplotted along each trajectory. All of the trajectories <strong>in</strong> the figure are attracted to thestable equilibrium denoted by the black dot, called an attractor. The overall qualitativedescription of dynamics can be obta<strong>in</strong>ed through the study of the phase portrait of thesystem, which depicts certa<strong>in</strong> special trajectories (equilibria, separatrices, limit cycles)that determ<strong>in</strong>e the topological behavior of all the other trajectories <strong>in</strong> the phase space.Probably 50 % of illustrations <strong>in</strong> this book are phase portraits.A fundamental property of neurons is excitability, illustrated <strong>in</strong> Fig. 1.9b. Theneuron is rest<strong>in</strong>g, i.e., its phase portrait has a stable equilibrium. Small perturbations,such as A, result <strong>in</strong> small excursions from the equilibrium, denoted as PSP (postsynapticpotential). In contrast, larger perturbations, such as B, are amplified bythe neuronal <strong>in</strong>tr<strong>in</strong>sic dynamics and result <strong>in</strong> the spike response. To understand thedynamic mechanism of such amplification, we need to consider the geometry of thephase portrait near the equilibrium, i.e., <strong>in</strong> the region where the decision to fire or notto fire is made.If we <strong>in</strong>ject a sufficiently strong current <strong>in</strong>to the neuron, we br<strong>in</strong>g it to a pacemak<strong>in</strong>gmode, so that it exhibits periodic spik<strong>in</strong>g activity, as <strong>in</strong> Fig. 1.9c. From the dynamical


10 Introductionspik<strong>in</strong>gmoderest<strong>in</strong>g modeFigure 1.10: Rhythmic transitions between rest<strong>in</strong>g and spik<strong>in</strong>g modes result <strong>in</strong> burst<strong>in</strong>gbehavior.layer 5 pyramidal cellbra<strong>in</strong>stem mesV celltransition20 mVtransition-60 mV-50 mV200 pA3000 pA0 pA500 ms0 pA500 msFigure 1.11: As the magnitude of the <strong>in</strong>jected current slowly <strong>in</strong>creases, the neuronsbifurcate from rest<strong>in</strong>g (equilibrium) to tonic spik<strong>in</strong>g (limit cycle) modes.systems po<strong>in</strong>t of view, the state of such a neuron has a stable limit cycle, also knownas a periodic orbit. The electrophysiological details of the neuron, i.e., the numberand the type of currents it has, their k<strong>in</strong>etics, etc., determ<strong>in</strong>e only the location, theshape and the period of the limit cycle. As long as the limit cycle exists, the neuroncan have periodic spik<strong>in</strong>g activity. Of course, equilibria and limit cycles can co-exist,so a neuron can be switched from one mode to another one by a transient <strong>in</strong>put. Thefamous example is the permanent ext<strong>in</strong>guish<strong>in</strong>g of ongo<strong>in</strong>g spik<strong>in</strong>g activity <strong>in</strong> the squidgiant axon by a brief transient depolariz<strong>in</strong>g pulse of current applied at a proper phase(Guttman et al. 1980) — a phenomenon predicted by John R<strong>in</strong>zel (1978) purely onthe basis of theoretical analysis of the Hodgk<strong>in</strong>-Huxley model. The transition betweenrest<strong>in</strong>g and spik<strong>in</strong>g modes could be triggered by <strong>in</strong>tr<strong>in</strong>sic slow conductances, result<strong>in</strong>g<strong>in</strong> the burst<strong>in</strong>g behavior <strong>in</strong> Fig. 1.10.1.2.2 BifurcationsNow suppose that the magnitude of the <strong>in</strong>jected current is a parameter that we cancontrol, e.g., we can ramp it up as <strong>in</strong> Fig. 1.11. Each cell <strong>in</strong> the figure is quiescentat the beg<strong>in</strong>n<strong>in</strong>g of the ramps, so its phase portrait has a stable equilibrium and it


Introduction 11may look like the one <strong>in</strong> Fig. 1.9a or b. Then it starts to fire tonic spikes, so its phaseportrait has a limit cycle attractor and it may look like the one <strong>in</strong> Fig. 1.9c, with whitecircle denot<strong>in</strong>g an unstable rest<strong>in</strong>g equilibrium. Apparently, there is some <strong>in</strong>termediatelevel of <strong>in</strong>jected current that corresponds to the transition from rest<strong>in</strong>g to susta<strong>in</strong>edspik<strong>in</strong>g, i.e., from the phase portrait <strong>in</strong> Fig. 1.9b to Fig. 1.9c. What does the transitionlook like?From dynamical systems po<strong>in</strong>t of view, the transition corresponds to a bifurcationof neuron dynamics, i.e., a qualitative change of phase portrait of the system. Forexample, there is no bifurcation go<strong>in</strong>g from phase portrait <strong>in</strong> Fig. 1.9a to that <strong>in</strong>Fig. 1.9b, s<strong>in</strong>ce both have one globally stable equilibrium; the difference <strong>in</strong> behavior isquantitative but not qualitative. In contrast, there is a bifurcation go<strong>in</strong>g from Fig. 1.9bto Fig. 1.9c s<strong>in</strong>ce the equilibrium is no longer stable and another attractor, limit cycle,appeared. The neuron is not excitable <strong>in</strong> Fig. 1.9a but it is <strong>in</strong> Fig. 1.9b simply becausethe former phase portrait is far away from the bifurcation and the latter is near.In general, neurons are excitable because they are near bifurcations from rest<strong>in</strong>g tospik<strong>in</strong>g activity, so the type of the bifurcation determ<strong>in</strong>es the excitable properties of theneuron. Of course, the type depends on the neuron’s electrophysiology. An amaz<strong>in</strong>gobservation is that there could be millions of different electrophysiological mechanismsof excitability and spik<strong>in</strong>g, but there are only 4, yes four, different types of bifurcationsof equilibrium that a system can undergo without any additional constra<strong>in</strong>ts, suchas symmetry. Thus, consider<strong>in</strong>g these four bifurcations <strong>in</strong> a general setup we canunderstand excitable properties of many models, even those that have not been <strong>in</strong>ventedyet. What is even more amaz<strong>in</strong>g, we can understand excitable properties of neuronswhose currents are not measured and whose models are not known, provided thatwe can identify experimentally which of the four bifurcations the rest<strong>in</strong>g state of theneuron undergoes.The four bifurcations are summarized <strong>in</strong> Fig. 1.12, which plots the phase portraitbefore (left), at (center), and after (right) a particular bifurcation occurs. Mathematiciansrefer to these bifurcations as be<strong>in</strong>g of co-dimension-1 because we need to vary onlyone parameter, e.g., the magnitude of the <strong>in</strong>jected dc-current I, to observe the bifurcationsreliably <strong>in</strong> simulations or experiments. There are many more co-dimension-2,3, etc., bifurcation, but they need special conditions to be observed. We discuss theselater <strong>in</strong> Chap. 6.Let us consider the four bifurcation and their phase portraits <strong>in</strong> the figure. Thehorizontal and vertical axes are the membrane potential with <strong>in</strong>stantaneous activationvariable and a recovery variable, respectively. At this stage, the reader is not requiredto fully understand the <strong>in</strong>tricacies of the phase portraits <strong>in</strong> the figure, s<strong>in</strong>ce they willbe expla<strong>in</strong>ed systematically <strong>in</strong> later chapters.• Saddle-node bifurcation. As the magnitude of the <strong>in</strong>jected current or any otherbifurcation parameter changes, a stable equilibrium correspond<strong>in</strong>g to the rest<strong>in</strong>gstate (black circle marked “node” <strong>in</strong> Fig. 1.12a) is approached by an unstableequilibrium (white circle marked “saddle”), they coalesce and annihilate eachother, as <strong>in</strong> Fig. 1.12a, middle. S<strong>in</strong>ce the rest<strong>in</strong>g state no longer exists, the


12 Introduction(a)spik<strong>in</strong>glimitcyclerecoverypotentialnodesaddlesaddle-nodesaddle-node bifurcation(b)<strong>in</strong>variantcirclenode saddle saddle-nodesaddle-node on <strong>in</strong>variant circle (SNIC) bifurcation(c)spik<strong>in</strong>glimitcycle attractorunstablesubcritical Andronov-Hopf bifurcation(d)supercritical Andronov-Hopf bifurcationFigure 1.12: Four generic (co-dimension-1) bifurcations of an equilibrium state lead<strong>in</strong>gto the transition from rest<strong>in</strong>g to periodic spik<strong>in</strong>g behavior <strong>in</strong> neurons.


Introduction 13trajectory describ<strong>in</strong>g the evolution of the system jumps to the limit cycle attractor<strong>in</strong>dicat<strong>in</strong>g that the neuron starts to fire tonic spikes. Notice that the limit cycle,or some other attractor, must co-exist with the rest<strong>in</strong>g state <strong>in</strong> order for thetransition rest<strong>in</strong>g → spik<strong>in</strong>g to occur.• Saddle-node on <strong>in</strong>variant circle bifurcation is similar to the saddle-node bifurcationabove with the exception that there is an <strong>in</strong>variant circle at the moment ofbifurcation, which then becomes a limit cycle attractor, as <strong>in</strong> Fig. 1.12b.• Subcritical Andronov-Hopf bifurcation. A small unstable limit cycle shr<strong>in</strong>ks toa stable equilibrium and makes it lose stability, as <strong>in</strong> Fig. 1.12c. Because of<strong>in</strong>stabilities, the trajectory diverges from the equilibrium and approaches a largeamplitudespik<strong>in</strong>g limit cycle or some other attractor.• Supercritical Andronov-Hopf bifurcation. The stable equilibrium loses stabilityand gives birth to a small-amplitude limit cycle attractor, as <strong>in</strong> Fig. 1.12d. Asthe magnitude of the <strong>in</strong>jected current <strong>in</strong>creases, the amplitude of the limit cycle<strong>in</strong>creases and it becomes full-size spik<strong>in</strong>g limit cycle.Notice that there is a co-existence of rest<strong>in</strong>g and spik<strong>in</strong>g states <strong>in</strong> the case of saddle-nodeand subcritical Andronov-Hopf bifurcations, whereas there is not <strong>in</strong> the other two cases.Such a co-existence reveals itself via a hysteresis behavior when the <strong>in</strong>jected current<strong>in</strong>creases and then decreases past the bifurcation value, because the transitions “rest<strong>in</strong>g→ spik<strong>in</strong>g” and “spik<strong>in</strong>g → rest<strong>in</strong>g” occur at different values of the current. In addition,brief stimuli applied at the appropriate times can switch the activity from spik<strong>in</strong>g torest<strong>in</strong>g and back. There are also spontaneous noise-<strong>in</strong>duced transitions between thetwo modes result<strong>in</strong>g <strong>in</strong> the stutter<strong>in</strong>g spik<strong>in</strong>g, as e.g. exhibited by the so called fastspik<strong>in</strong>g (FS) cortical <strong>in</strong>terneurons when they are kept close to the bifurcation (Tatenoet al. 2004). Some bistable neurons have a slow adaptation current that activatesdur<strong>in</strong>g the spik<strong>in</strong>g mode and impedes spik<strong>in</strong>g, often result<strong>in</strong>g <strong>in</strong> burst<strong>in</strong>g activity.<strong>Systems</strong> undergo<strong>in</strong>g Andronov-Hopf bifurcations, whether subcritical or supercritical,exhibit damped oscillations of membrane potential, whereas systems near saddlenodebifurcations, whether on or off an <strong>in</strong>variant circle, do not. The existence ofsmall amplitude oscillations creates the possibility of resonance to the frequency of the<strong>in</strong>com<strong>in</strong>g pulses, as <strong>in</strong> Fig. 1.7, and other <strong>in</strong>terest<strong>in</strong>g features.We refer to neurons with damped subthreshold oscillations as resonators and tothose that do not have this property as <strong>in</strong>tegrators. We refer to the neurons that exhibitthe co-existence of rest<strong>in</strong>g and spik<strong>in</strong>g states, at least near the transition from rest<strong>in</strong>g tospik<strong>in</strong>g, as bistable, and to those that do not exhibit the bistability as monostable. Thefour bifurcations <strong>in</strong> Fig. 1.12 are uniquely def<strong>in</strong>ed by these two features. For example,a bistable resonator is a neuron undergo<strong>in</strong>g subcritical Andronov-Hopf bifurcation,and a monostable <strong>in</strong>tegrator is a neuron undergo<strong>in</strong>g saddle-node on <strong>in</strong>variant circlebifurcation; see table <strong>in</strong> Fig. 1.13. Cortical fast spik<strong>in</strong>g (FS) and regular spik<strong>in</strong>g(RS) neurons, studied <strong>in</strong> Chap. 8, are typical examples of the former and the latter,respectively.


14 Introductionco-existence of rest<strong>in</strong>g and spik<strong>in</strong>g statesYES(bistable)NO(monostable)subthreshold oscillationsNO(<strong>in</strong>tegrator)YES(resonator)saddle-nodesubcriticalAndronov-Hopfsaddle-node on<strong>in</strong>variant circlesupercriticalAndronov-HopfFigure 1.13: Classification of neurons <strong>in</strong>tomonostable/bistable <strong>in</strong>tegrators/resonatorsaccord<strong>in</strong>g to the bifurcation of the rest<strong>in</strong>gstate <strong>in</strong> Fig. 1.12.asymptotic fir<strong>in</strong>g frequency, Hz40302010Class 1 excitabilityF-I curve00 100 200 300<strong>in</strong>jected dc-current, I (pA)asymptotic fir<strong>in</strong>g frequency, Hz25020015010050Class 2 excitabilityF-I curve00 500 1000 1500<strong>in</strong>jected dc-current, I (pA)Figure 1.14: Frequency-current (F-I) curves of cortical pyramidal neuron and bra<strong>in</strong>stemmesV neuron from Fig. 7.3. These are the same neurons used <strong>in</strong> the ramp experiment<strong>in</strong> Fig. 1.11.1.2.3 Hodgk<strong>in</strong> classificationHodgk<strong>in</strong> (1948) was the first to study bifurcations <strong>in</strong> neuronal dynamics, years beforethe mathematical theory of bifurcations was developed. He stimulated squid axonswith pulses of various amplitudes and identified three classes of responses:• Class 1 neural excitability. Action potentials can be generated with arbitrarilylow frequency, depend<strong>in</strong>g on the strength of the applied current.• Class 2 neural excitability. Action potentials are generated <strong>in</strong> a certa<strong>in</strong>frequency band that is relatively <strong>in</strong>sensitive to changes <strong>in</strong> the strength of theapplied current.• Class 3 neural excitability. A s<strong>in</strong>gle action potential is generated <strong>in</strong> responseto a pulse of current. Repetitive (tonic) spik<strong>in</strong>g can be generated only forextremely strong <strong>in</strong>jected currents or not at all.The qualitative dist<strong>in</strong>ction between the classes is that the frequency-current relation(the F-I curve <strong>in</strong> Fig. 1.14) starts from zero and cont<strong>in</strong>uously <strong>in</strong>creases for Class 1


Introduction 15neurons, is discont<strong>in</strong>uous for Class 2 neurons, and is not def<strong>in</strong>ed at all for Class 3neurons.Obviously, neurons belong<strong>in</strong>g to different classes have different neuro-computationalproperties: Class 1 neurons, which <strong>in</strong>clude cortical excitatory pyramidal neurons, cansmoothly encode the strength of the <strong>in</strong>put <strong>in</strong>to the output fir<strong>in</strong>g frequency, as <strong>in</strong>Fig. 1.11, left. In contrast, Class 2 neurons, such as fast-spik<strong>in</strong>g (FS) cortical <strong>in</strong>hibitory<strong>in</strong>terneurons, cannot do that; <strong>in</strong>stead, they fire <strong>in</strong> a relatively narrow frequency band,as <strong>in</strong> Fig. 1.11, right. Class 3 neurons cannot exhibit susta<strong>in</strong>ed spik<strong>in</strong>g activity, soHodgk<strong>in</strong> regarded them as “sick” or “unhealthy”. There are other dist<strong>in</strong>ctions betweenthe classes, which we discuss later.Different classes of excitability occur because neurons have different bifurcations ofrest<strong>in</strong>g and spik<strong>in</strong>g states – a phenomenon first expla<strong>in</strong>ed by R<strong>in</strong>zel and Ermentrout(1989). If ramps of current are <strong>in</strong>jected to measure the F-I curves, then Class 1 excitabilityoccurs when the neuron undergoes the saddle-node bifurcation on <strong>in</strong>variantcircle depicted <strong>in</strong> Fig. 1.12b. Indeed, the period of the limit cycle attractor is <strong>in</strong>f<strong>in</strong>iteat the bifurcation po<strong>in</strong>t, and then it decreases as the bifurcation parameter – say, the<strong>in</strong>jected current – <strong>in</strong>creases. The other three bifurcations result <strong>in</strong> Class 2 excitability.Indeed, the limit cycle attractor exists and has a f<strong>in</strong>ite period when the rest<strong>in</strong>g state<strong>in</strong> Fig. 1.12 undergoes a subcritical Andronov-Hopf bifurcation, so emerg<strong>in</strong>g spik<strong>in</strong>ghas a non-zero frequency. The period of the small limit cycle attractor appear<strong>in</strong>g viasupercritical Andronov-Hopf bifurcation is also f<strong>in</strong>ite, so the frequency of oscillationsis non-zero, but their amplitudes are small. In contrast to the common and erroneousfolklore, the saddle-node bifurcation (off limit cycle) also results <strong>in</strong> Class 2 excitabilitybecause the limit cycle has a f<strong>in</strong>ite period at the bifurcation. There is a considerablelatency (delay) to the first spike <strong>in</strong> this case, but the subsequent spik<strong>in</strong>g has non-zerofrequency. Thus, the simple scheme “Class 1 = saddle-node, Class 2 = Hopf” thatpermeates many publications is <strong>in</strong>correct.When pulses of current are used to measure the F-I curve, as <strong>in</strong> Hodgk<strong>in</strong>’s experiments,the fir<strong>in</strong>g frequency depends on other factors, and not only the type of thebifurcation of the rest<strong>in</strong>g state. In particular, low-frequency fir<strong>in</strong>g can be observed <strong>in</strong>systems near Andronov-Hopf bifurcations, as we show <strong>in</strong> Chap. 7. To avoid possibleconfusion, we def<strong>in</strong>e the class of excitability based only on slow ramp experiments.Hodgk<strong>in</strong>’s classification has an important historical value but it is of little use forthe dynamic description of a neuron, s<strong>in</strong>ce nam<strong>in</strong>g a class of excitability of a neurondoes not tell much about the bifurcations of the rest<strong>in</strong>g state. Indeed, it only says thatsaddle-node on <strong>in</strong>variant circle bifurcation (Class 1) is different from the other threebifurcations (Class 2), and only when ramps are <strong>in</strong>jected. Instead, divid<strong>in</strong>g neurons <strong>in</strong>to<strong>in</strong>tegrators and resonators with bistable or monostable activity is more <strong>in</strong>formative, sowe adopt the classification <strong>in</strong> Fig. 1.13 <strong>in</strong> this book. In this classification, Class 1 neuronis a monostable <strong>in</strong>tegrator, whereas Class 2 neuron could be a bistable <strong>in</strong>tegrator or aresonator.


16 Introduction1.2.4 Neuro-computational propertiesUs<strong>in</strong>g the same arrangement as <strong>in</strong> Fig. 1.13, we depict typical geometry of phaseportraits near the four bifurcations <strong>in</strong> Fig. 1.15. Let us use the portraits to expla<strong>in</strong> whathappens “near the threshold”, i.e., near the place where the decision to fire or not ismade. To simplify our geometrical analysis we assume here that neurons receive shock<strong>in</strong>puts, i.e., brief but strong pulses of current that do not change the phase portraitsbut only push or reset the state of the neuron <strong>in</strong>to various regions of the phase space.We consider these and other cases <strong>in</strong> detail <strong>in</strong> Chap. 7.The horizontal axis <strong>in</strong> each plot <strong>in</strong> Fig. 1.15 corresponds to the membrane potentialV with <strong>in</strong>stantaneous Na + current, and the vertical axis corresponds to a recovery variable,say activation of K + current. Black circles denote stable equilibria correspond<strong>in</strong>gto the neuronal rest<strong>in</strong>g state. Spik<strong>in</strong>g limit cycle attractors correspond to susta<strong>in</strong>edspik<strong>in</strong>g states, which exist <strong>in</strong> the two cases depicted <strong>in</strong> the left half of the figure correspond<strong>in</strong>gto the bistable dynamics. The limit cycles are surrounded by the shadedregions — their attraction doma<strong>in</strong>s. The white region is the attraction doma<strong>in</strong> of theequilibrium. To <strong>in</strong>itiate spik<strong>in</strong>g, the external <strong>in</strong>put should push the state of the system<strong>in</strong>to the shaded region, and to ext<strong>in</strong>guish spik<strong>in</strong>g, the <strong>in</strong>put should push the state back<strong>in</strong>to the white region.There are no limit cycles <strong>in</strong> the two cases depicted <strong>in</strong> the right half of the figure,so the entire phase space is the attraction doma<strong>in</strong> of the stable equilibrium, and thedynamics are monostable. However, if the trajectory starts <strong>in</strong> the shaded region, itmakes a large-amplitude rotation before return<strong>in</strong>g to the equilibrium — a transientspike. Apparently, to elicit such a spike, the <strong>in</strong>put should push the state of the system<strong>in</strong>to the shaded region.Now let us contrast the upper and lower halves of the figure correspond<strong>in</strong>g to<strong>in</strong>tegrators and resonators, respectively. We dist<strong>in</strong>guish these two modes of operationbased on the existence of subthreshold oscillations near the equilibrium.First, let us show that <strong>in</strong>hibition impedes spik<strong>in</strong>g <strong>in</strong> <strong>in</strong>tegrators, but can promote it<strong>in</strong> resonators. In the <strong>in</strong>tegrator case, the shaded region is <strong>in</strong> the depolarized voltagerange, i.e., to the right of the equilibrium. Excitatory <strong>in</strong>puts push the state of thesystem toward the shaded region, while <strong>in</strong>hibitory <strong>in</strong>puts push it away. In the case ofresonators, both excitation and <strong>in</strong>hibition push the state toward the shaded region, becausethe region wraps around the equilibrium and can be reached along any direction.This expla<strong>in</strong>s the rebound spik<strong>in</strong>g phenomenon depicted <strong>in</strong> Fig. 1.6.Integrators have all-or-none spikes while resonators may not. Indeed, any trajectorystart<strong>in</strong>g <strong>in</strong> the shaded region <strong>in</strong> the upper half of the figure has to rotate aroundthe white circle at the top correspond<strong>in</strong>g to an unstable equilibrium. Moreover, thestate of the system is quickly attracted to the spik<strong>in</strong>g trajectory and moves along thetrajectory thereby generat<strong>in</strong>g a stereotypical spike. A resonator neuron can also firelarge-amplitude spikes when its state is pushed to or beyond the trajectory denoted“spike”. Such neurons generate subthreshold responses when the state slides along


Introduction 17saddle-node bifurcationsaddle-node on <strong>in</strong>variant circle bifurcationrecoverypotentialspik<strong>in</strong>gthresholdlimitcycle attractorspik<strong>in</strong>gtrajectorythreshold<strong>in</strong>hexc12subcritical Andronov-Hopf bifurcationsupercritical Andronov-Hopf bifurcationspik<strong>in</strong>g limitcycle attractorhalf-amplitudespikespikehalf-amplitudespikePSPspike1<strong>in</strong>hPSPexc23Figure 1.15: The geometry of phase portraits of excitable systems near 4 bifurcationscan expla<strong>in</strong> many neuro-computational properties (see Sect. 1.2.4 for detail).the smaller trajectory denoted “PSP”; they can also generate spikes of an <strong>in</strong>termediateamplitude when the state is pushed between the “PSP” and “spike” trajectories, whichexpla<strong>in</strong>s the partial-amplitude spik<strong>in</strong>g <strong>in</strong> Fig. 1.5 or <strong>in</strong> the squid axon <strong>in</strong> Fig. 7.26. Theset of <strong>in</strong>itial conditions correspond<strong>in</strong>g to such spik<strong>in</strong>g is quite small, so typical spikeshave large amplitudes and partial spikes are rare.Integrators have well-def<strong>in</strong>ed thresholds while resonators may not. The white circlesnear the rest<strong>in</strong>g states of <strong>in</strong>tegrators <strong>in</strong> Fig. 1.15 are called saddles. They are stablealong the vertical direction and unstable along the horizontal direction. The two trajectoriesthat lead to the saddle along the vertical direction are called separatrices becausethey separate the phase space <strong>in</strong>to two regions, <strong>in</strong> this case <strong>in</strong>to white and shaded. Theseparatrices play the role of thresholds s<strong>in</strong>ce only those perturbations that push thestate of the system beyond the separatrices result <strong>in</strong> a spike. The closer is the state of


18 IntroductionspikeFigure 1.16: Phase portrait of a system near aBogdanov-Takens bifurcation that corresponds tothe transition from <strong>in</strong>tegrator to resonator mode.the system to the separatrices, the longer it takes to converge and then diverge fromthe saddle, result<strong>in</strong>g <strong>in</strong> a long latency to the spike. Notice that the threshold is not apo<strong>in</strong>t but a tilted curve that spans a range of voltage values.Resonators have a well-def<strong>in</strong>ed threshold <strong>in</strong> the case of subcritical Andronov-Hopfbifurcation: it is the small unstable limit cycle that separates the attraction doma<strong>in</strong>sof stable equilibrium and spik<strong>in</strong>g limit cycle. Trajectories <strong>in</strong>side the small cycle spiraltoward the stable equilibrium, while trajectories outside the cycle spiral away andeventually lead to susta<strong>in</strong>ed spik<strong>in</strong>g activity. When a neuronal model is far from thesubcritical Andronov-Hopf bifurcation, its phase portrait may look similar to the onecorrespond<strong>in</strong>g to the supercritical Andronov-Hopf bifurcation. The narrow shadedband <strong>in</strong> the figure is not a threshold manifold but a fuzzy threshold set called “quasithreshold”by FitzHugh (1955). Many resonators, <strong>in</strong>clud<strong>in</strong>g the Hodgk<strong>in</strong>-Huxley modelhave quasi-thresholds. The width of the quasi-threshold <strong>in</strong> the Hodgk<strong>in</strong>-Huxley modelis so narrow, that it may be assumed to be just a curve for all practical reasons.Integrators <strong>in</strong>tegrate, resonators resonate. Now consider <strong>in</strong>puts consist<strong>in</strong>g of multiplepulses, e.g., a burst of spikes. Integrators prefer high-frequency <strong>in</strong>puts; the higherthe frequency, the sooner they fire. Indeed, the first spike of such an <strong>in</strong>put, marked“1” <strong>in</strong> the top-right phase portrait <strong>in</strong> Fig. 1.15, <strong>in</strong>creases the membrane potential andshifts the state to the right toward the threshold. S<strong>in</strong>ce the state of the system isstill <strong>in</strong> the white area, it slowly converges back to the stable equilibrium. To crossthe threshold manifold, the second pulse must arrive shortly after the first one. Thereaction of a resonator to a pair of pulses is quite different. The first pulse <strong>in</strong>itiates adamped subthreshold oscillation of the membrane potential, which looks like a spiral<strong>in</strong> the bottom-right phase portrait <strong>in</strong> Fig. 1.15. The effect of the second pulse dependson its tim<strong>in</strong>g. If it arrives after the trajectory makes half a rotation, marked as “2”<strong>in</strong> the figure, it cancels the effect of the first pulse. If it arrives after the trajectorymakes a full rotation, marked “3” <strong>in</strong> the figure, it adds to the first pulse and either<strong>in</strong>creases the amplitude of subthreshold oscillation or evokes a spike response. Thus,the response of the resonator neuron depends on the frequency content of the <strong>in</strong>put,as <strong>in</strong> Fig. 1.7.Integrators and resonators constitute two major modes of activity of neurons. Mostcortical pyramidal neurons, <strong>in</strong>clud<strong>in</strong>g the regular spik<strong>in</strong>g (RS), <strong>in</strong>tr<strong>in</strong>sically burst<strong>in</strong>g


Introduction 19(IB), and chatter<strong>in</strong>g (CH) types considered <strong>in</strong> Chap. 8, are <strong>in</strong>tegrators. So are thalamocorticalneurons <strong>in</strong> the relay mode of fir<strong>in</strong>g and neostriatal sp<strong>in</strong>y projection neurons.Most cortical <strong>in</strong>hibitory <strong>in</strong>terneurons, <strong>in</strong>clud<strong>in</strong>g the fast spik<strong>in</strong>g type, are resonators.So are bra<strong>in</strong>stem mesencephalic V neurons and stellate neurons of the entorh<strong>in</strong>al cortex.Some cortical pyramidal neurons and low-threshold spik<strong>in</strong>g (LTS) <strong>in</strong>terneurons can beat the border of transition between <strong>in</strong>tegrator and resonator modes. Such a transitioncorresponds to another bifurcation, which has co-dimension-2, and hence it is less likelyto be encountered experimentally. We consider this and other uncommon bifurcations<strong>in</strong> detail later. The phase portrait near the bifurcation is depicted <strong>in</strong> Fig. 1.16 and it isa good exercise for the reader to expla<strong>in</strong> why such a system has damped oscillations andpost-<strong>in</strong>hibitory responses yet a well-def<strong>in</strong>ed threshold, all-or-none spikes with possiblylong latencies.Of course, figures 1.15 and 1.16 cannot encompass all the richness of neuronalbehavior, otherwise this book would be only 19-pages long 1 . Many aspects of neuronaldynamics depend on other bifurcations, e.g., those correspond<strong>in</strong>g to appearance anddisappearance of spik<strong>in</strong>g limit cycles. These bifurcations describe the transitions fromspik<strong>in</strong>g to rest<strong>in</strong>g, and they are especially important when we consider burst<strong>in</strong>g activity.In addition, we need to take <strong>in</strong>to account the relative geometry of equilibria, limitcycles, and other relevant trajectories, and how they depend on the parameters of thesystem, such as maximal conductances, activation time constants, etc. We explore allthese issues systematically <strong>in</strong> subsequent chapters.In Chap. 2 we review some of the most fundamental concepts of neuron electrophysiology,culm<strong>in</strong>at<strong>in</strong>g with the Hodgk<strong>in</strong>-Huxley model. This chapter is aimed atmathematicians learn<strong>in</strong>g neuroscience. In Chapters 3 and 4 we use one- and twodimensionalneuronal models, respectively, to review some of the most fundamentalconcepts of dynamical systems, such as equilibria, limit cycles, stability, attractiondoma<strong>in</strong>, nullcl<strong>in</strong>es, phase portrait, bifurcation, etc. The material <strong>in</strong> these chapters,aimed at biologists learn<strong>in</strong>g the language of dynamical systems, is presented with theemphasis on geometrical rather than mathematical <strong>in</strong>tuition. In fact, the spirit of theentire book is to expla<strong>in</strong> concepts us<strong>in</strong>g pictures, not equations. Chap. 5 exploresphase portraits of various conductance-based models and the relations between ioniccurrents and dynamic behavior. In Chap. 6 we use the I Na,p +I K -model to systematically<strong>in</strong>troduce the geometric bifurcation theory. Chap. 7, probably the most importantchapter of the book, applies the theory to expla<strong>in</strong> many computational properties ofneurons. In fact, all the material <strong>in</strong> the previous chapters is given so that the readercan understand this chapter. In Chap. 8 we use a simple phenomenological model tosimulate many cortical and thalamic neurons. This chapter conta<strong>in</strong>s probably the mostcomprehensive up to date review of various fir<strong>in</strong>g patterns exhibited by mammalianneurons. In Chap. 9 we <strong>in</strong>troduce the electrophysiological and topological classificationof burst<strong>in</strong>g dynamics, as well as some useful methods to study the bursters. F<strong>in</strong>ally,the last and the most mathematically advanced chapter of the book, Chap. 10, deals1 This book is actually quite short; Most of the space is taken by figures, exercises, and solutions.


20 Introductionwith coupled neurons. There we show how the details of spike-generation mechanismof neurons affect their collective properties, such as synchronization.1.2.5 Build<strong>in</strong>g models (revisited)To have a good model of a neuron, it is not enough to put the right k<strong>in</strong>d of currents togetherand tune the parameters so that the model can fire spikes. It is not even enoughto reproduce the right <strong>in</strong>put resistance, rheobase, and fir<strong>in</strong>g frequencies. The modelhas to reproduce all the neuro-computational features of the neuron, start<strong>in</strong>g withthe co-existence of rest<strong>in</strong>g and spik<strong>in</strong>g states, spike latencies, subthreshold oscillations,rebound spikes, etc.A good way to start is to determ<strong>in</strong>e what k<strong>in</strong>d of bifurcations the neuron underconsideration undergoes and how the bifurcations depend on neuromodulators andpharmacological blockers. Instead of or <strong>in</strong> addition to measur<strong>in</strong>g neuronal responsesto get the k<strong>in</strong>etic parameters, we need to measure them to get the right bifurcationbehavior. Only <strong>in</strong> this case we can be sure that the behavior of the model is equivalentto that of the neuron, even if we omitted a current or guessed some of the parameters<strong>in</strong>correctly.Implementation of this research program is still a pipe dream. The people whounderstand the mathematical aspects of neuron dynamics, those who see beyond conductancesand currents, those people do not usually have the opportunity to do experiments.Conversely, those who study neurons <strong>in</strong> vitro or <strong>in</strong> vivo on a daily basis,those who see spik<strong>in</strong>g, burst<strong>in</strong>g, oscillations, those who can manipulate the experimentalsetup to test practically any aspect of neuronal activity, those people do notusually see the value of study<strong>in</strong>g phase portraits, bifurcations, and nonl<strong>in</strong>ear dynamics<strong>in</strong> general. One of the goals of this book is to change this state and br<strong>in</strong>g these twogroups of people closer together.


Introduction 21Review of Important Concepts• Neurons are dynamical systems.• Rest<strong>in</strong>g state of neurons corresponds to a stable equilibrium, tonicspik<strong>in</strong>g state corresponds to a limit cycle attractor.• Neurons are excitable because the equilibrium is near a bifurcation.• There are many ionic mechanisms of spike-generation, but only fourgeneric bifurcations of equilibria.• These bifurcations divide neurons <strong>in</strong>to four categories: <strong>in</strong>tegratorsor resonators, monostable or bistable.• Analyses of phase portraits at bifurcations expla<strong>in</strong> why some neuronshave well-def<strong>in</strong>ed thresholds, all-or-none spikes, post-<strong>in</strong>hibitoryspikes, frequency preference, hysteresis, etc., while others do not.• These features, and not ionic currents per se, determ<strong>in</strong>e the neuronalresponses, i.e., the k<strong>in</strong>d of computations neurons do.• A good neuronal model must reproduce not only electrophysiologybut also the bifurcation dynamics of neurons.Bibliographical NotesRichard FitzHugh at the National Institutes of Health (NIH) pioneered the phase planeanalysis of neuronal models with the view to understand their neuro-computationalproperties. He was the first to analyze the Hodgk<strong>in</strong>-Huxley model (FitzHugh 1955;years before they received the Nobel prize) and to prove that it has neither thresholdnor all-or-none spikes. FitzHugh (1961) <strong>in</strong>troduced a simplified model of excitabilityand showed that one can get the right k<strong>in</strong>d of neuronal dynamics <strong>in</strong> models lack<strong>in</strong>gconductances and currents. Nagumo et al. (1962) designed a correspond<strong>in</strong>g tunneldiode circuit, so the model is called the FitzHugh-Nagumo oscillator. Chapter 8 dealswith such simplified models.FitzHugh research program was further developed by John R<strong>in</strong>zel and G. BardErmentrout. In their 1989 sem<strong>in</strong>al paper, R<strong>in</strong>zel and Ermentrout revived Hodgk<strong>in</strong>’sclassification of excitability and po<strong>in</strong>ted out to the connection between the behavior ofneuronal models and the bifurcations they exhibit. (They also refer to the excitabilityas “type I” and “type II”). Unfortunately, many people treat the connection <strong>in</strong> asimplem<strong>in</strong>ded fashion and <strong>in</strong>correctly identify “type I = saddle-node, type II = Hopf”.


22 IntroductionFigure 1.17: John R<strong>in</strong>zel <strong>in</strong> 2004. Incidentally, his T-shirt shows the cover of the firstissue of Journal of Computational <strong>Neuroscience</strong>, <strong>in</strong> which the P<strong>in</strong>sky-R<strong>in</strong>zel (1994)model appeared.Figure 1.18: G. Bard Ermentrout (G. stands for George) with his parrot Junior <strong>in</strong>1983.


Introduction 23Figure 1.19: Richard FitzHugh <strong>in</strong> 1984.If only life was so simple!The geometrical analysis of neuronal models was further developed, among others,by Izhikevich (2000), who stressed the <strong>in</strong>tegrator and resonator modes of operationand made connections to other neuro-computational properties.The neuroscience and mathematics parts of this book are standard, though manyconnections are new. We po<strong>in</strong>t to the literature sources at the end of each chapter.Among many outstand<strong>in</strong>g books on computational neuroscience, we especially recommendSpikes, Decisions, and Actions by Wilson (1999), Biophysics of Computation byKoch (1999), Theoretical <strong>Neuroscience</strong> by Dayan and Abbott (2001), and Foundationsof Cellular Neurophysiology by Johnston and Wu (1995). The present monographcomplements these excellent books <strong>in</strong> the sense that it is more ambitious, focused,and thorough <strong>in</strong> deal<strong>in</strong>g with neurons as dynamical systems. Though the views <strong>in</strong>this monograph may be biased by the author’s philosophy and taste, the payoffs <strong>in</strong>understand<strong>in</strong>g neuronal dynamics are immense, provided that the reader has enoughpatience and perseverance to follow the author’s l<strong>in</strong>e of thought.NEURON simulation environment is described by H<strong>in</strong>es (1989) (http://www.neuron.yale.edu),GENESIS environment is described by Bower and Beeman (1995) (http://www.genesis-sim.org),XPP environment is described by Ermentrout (2002). The author of this book usesMATLAB (version 6.5 for W<strong>in</strong>dows), which has become a standard computational tool<strong>in</strong> science and eng<strong>in</strong>eer<strong>in</strong>g (MATLAB is registered trademark of The MathWorks, Inc.;see http://www.mathworks.com).


24 Introduction


Chapter 2Electrophysiology of NeuronsIn this chapter we rem<strong>in</strong>d the reader of some fundamental concepts of neuronal electrophysiology,which are necessary to understand the rest of the book. We start with ionsand currents, and move quickly toward the dynamics of the Hodgk<strong>in</strong>-Huxley model.If the reader is already familiar with the Hodgk<strong>in</strong>-Huxley formalism, this chapter canbe skipped. Our exposition is brief, and it cannot substitute a good <strong>in</strong>troductory neurosciencecourse or read<strong>in</strong>g of such excellent textbooks as Theoretical <strong>Neuroscience</strong> byDayan and Abbott (2001), Foundations of Cellular Neurophysiology by Johnston andWu (1995), Biophysics of Computation by Koch (1999) or Ion Channels of ExcitableMembranes by Hille (2001).2.1 IonsElectrical activity <strong>in</strong> neurons is susta<strong>in</strong>ed and propagated via ionic currents throughneuron membranes. Most of these transmembrane currents <strong>in</strong>volve four ionic species:sodium (Na + ), potassium (K + ), calcium (Ca 2+ ), and chloride (Cl − ). The first threehave a positive charge (cations) and the fourth has a negative charge (anion). Theconcentrations of these ions are different on the <strong>in</strong>side and outside of a cell, whichcreates electrochemical gradients — the major driv<strong>in</strong>g forces of neural activity. Theextracellular medium has high concentration of Na + and Cl − (salty like seawater) and arelatively high concentration of Ca 2+ . The <strong>in</strong>tracellular medium has high concentrationof K + and negatively charged molecules (denoted by A − ), as we illustrate <strong>in</strong> Fig. 2.1.The cell membrane has large prote<strong>in</strong> molecules form<strong>in</strong>g channels through whichions (but not A − ) can flow accord<strong>in</strong>g to their electrochemical gradients. The flow ofNa + and Ca 2+ ions is not significant, at least at rest, but the flow of K + and Cl − ionsis. This, however, does not elim<strong>in</strong>ate the concentration asymmetry due to the follow<strong>in</strong>gtwo reasons.• Passive redistribution. The impermeable anions A − attract more K + <strong>in</strong>to the cell(opposites attract) and repel more Cl − out of the cell, thereby creat<strong>in</strong>g concentrationgradients.• Active transport. Ions are pumped <strong>in</strong> and out of the cell via ionic pumps. Forexample, the Na + -K + pump depicted <strong>in</strong> Fig. 2.1 pumps out three Na + ions for25


26 Electrophysiology of NeuronsInsideNa + (5-15 mM)K + (140 mM)Cl-(4 mM)Ca 2+ (0.1 µ M)A-(147 mM)ActiveTransportNa +Na + Na +PumpK + K +K +A -OutsideK + A - Na + Cl -Na + (145 mM)K + (5 mM)Cl-(110 mM)Ca 2+ (2.5-5 mM)A-(25 mM)Cl -PassiveRedistributionEquilibrium PotentialsNa + 62 log 1455= 90 mV62 log 14515= 61 mVK + 62 log 5140= −90 mVCl − −62 log 1104= −89 mVCa 2+ 31 log 2.510 −4 = 136 mV31 log 510 −4 = 146 mVFigure 2.1: Ion concentrations and Nernst equilibrium potentials (2.1) <strong>in</strong> a typicalmammalian neuron (modified from Johnston and Wu 1995). A − are membraneimpermeantanions. Temperature T = 37 ◦ C (310 ◦ K).every two K + ions pumped <strong>in</strong>, thereby ma<strong>in</strong>ta<strong>in</strong><strong>in</strong>g concentration gradients.2.1.1 Nernst potentialThere are two forces that drive each ion species through the membrane channel: concentrationand electric potential gradients. First, the ions diffuse down the concentrationgradient. For example, the K + ions depicted <strong>in</strong> Fig. 2.2a diffuse out of the cell becauseK + concentration <strong>in</strong>side is higher than that outside. While exit<strong>in</strong>g the cell, K + ionscarry positive charge with them and leave a net negative charge <strong>in</strong>side the cell (consist<strong>in</strong>gmostly of impermeable anions A − ), thereby produc<strong>in</strong>g the outward current. Thepositive and negative charges accumulate on the opposite sides of the membrane surfacecreat<strong>in</strong>g an electric potential gradient across the membrane – transmembrane potentialor membrane voltage. This potential slows down the diffusion of K + , s<strong>in</strong>ce K + ions areattracted to the negatively charged <strong>in</strong>terior and repelled from the positively chargedexterior of the membrane, as we illustrate <strong>in</strong> Fig. 2.2b. At some po<strong>in</strong>t an equilibriumis achieved: The concentration gradient and the electric potential gradient exert equaland opposite forces that counterbalance each other, and the net cross-membrane currentis zero, as <strong>in</strong> Fig. 2.2c. The value of such an equilibrium potential depends on theionic species, and it is given by the Nernst equation (Hille 2001)E ion = RTzF ln [Ion] out[Ion] <strong>in</strong>, (2.1)


Electrophysiology of Neurons 27InsideOutsideInsideOutsideInsideOutsideK + Na + Cl -A - Na +Cl -DiffusionK +Cl - Na +K + A -A -K +K +Cl -A -Na + Na +A - Cl -K +Na +A -K + Na + Na +A -Cl -A -K +Cl -A - Na +A -K +Na +Cl - A - K + Cl -DiffusionDiffusionK + A -K + Cl - K + A - Na + Cl -K +A - K +A -K +Na +A - A -K +K +K +A - K +A - K +Cl -A - K +A - ClA - Electric-ElectricNaPotential+ K + A -KPotential+Na +A - Na +A - K +A -K +Cl -K +A - A - K+Na +Na +K +Cl -A - A -K +Na +A -Cl -Cl -K +Cl -a b cFigure 2.2: Diffusion of K + ions down the concentration gradient though the membrane(a) creates an electric potential force po<strong>in</strong>t<strong>in</strong>g <strong>in</strong> the opposite direction (b) until thediffusion and electrical forces counter each other (c). The result<strong>in</strong>g transmembranepotential (2.1) is referred to as the Nernst equilibrium potential for K + .where [Ion] <strong>in</strong> and [Ion] out are concentrations of the ions <strong>in</strong>side and outside the cell,respectively, R is the universal gas constant (8, 315 mJ/(K ◦·Mol)), T is temperature <strong>in</strong>degrees Kelv<strong>in</strong> (K ◦ = 273.16+C ◦ ), F is Faraday’s constant (96, 480 Coulombs/Mol),z is the valence of the ion (z = 1 for Na + and K + , z = −1 for Cl − , and z = 2 forCa 2+ ). Substitut<strong>in</strong>g the numbers, tak<strong>in</strong>g log 10 <strong>in</strong>stead of natural ln and us<strong>in</strong>g bodytemperature T = 310 ◦ K (37 ◦ C) results <strong>in</strong>E ion ≈ 62 log [Ion] out[Ion] <strong>in</strong>(mV)for monovalent (z = 1) ions. Nernst equilibrium potentials <strong>in</strong> a typical mammalianneuron are summarized <strong>in</strong> Fig. 2.1.2.1.2 Ionic currents and conductancesIn the rest of the book V denotes the membrane potential and E Na , E Ca , E K , and E Cldenote the Nernst equilibrium potentials. When the membrane potential equals theequilibrium potential, say E K , the net K + current, denoted as I K (µA/cm 2 ), is zero(this is the def<strong>in</strong>ition of the Nernst equilibrium potential for K + ). Otherwise, the netK + current is proportional to the difference of potentials, i.e.I K = g K (V − E K ) ,


28 Electrophysiology of NeuronsoutsideI NagNaI CagCaI KgKI ClgClCCVENaECaEKEClFigure 2.3: Equivalent circuit representationof a patch of cell membrane.<strong>in</strong>sidewhere the positive parameter g K (mS/cm 2 ) is the K + conductance and (V − E K ) is theK + driv<strong>in</strong>g force. The other major ionic currentsI Na = g Na (V − E Na ) , I Ca = g Ca (V − E Ca ) , I Cl = g Cl (V − E Cl )could also be expressed as products of non-l<strong>in</strong>ear conductances and correspond<strong>in</strong>gdriv<strong>in</strong>g forces. A better description of membrane currents, especially Ca 2+ current, isprovided by the Goldman-Hodgk<strong>in</strong>-Katz equation (Hille 2001), which we do not use <strong>in</strong>this book.When the conductance is constant, the current is said to be Ohmic. In general,ionic currents <strong>in</strong> neurons are not Ohmic, s<strong>in</strong>ce the conductances may depend on time,membrane potential, and pharmacological agents, e.g., neurotransmitters, neuromodulators,second-messengers, etc. It is the time-dependent variation <strong>in</strong> conductances thatallows a neuron to generate an action potential, or spike.2.1.3 Equivalent circuitIt is traditional to represent electrical properties of membranes <strong>in</strong> terms of equivalentcircuits similar to the one depicted <strong>in</strong> Fig. 2.3. Accord<strong>in</strong>g to Kirchhoff’s law, the totalcurrent, I, flow<strong>in</strong>g across a patch of a cell membrane is the sum of the membranecapacitive current C ˙V (the capacitance C ≈ 1.0 µF/cm 2 <strong>in</strong> the squid axon) and all theionic currentsI = C ˙V + I Na + I Ca + I K + I Cl ,where ˙V = dV/dt is the derivative of the voltage variable V with respect to time t.The derivative arises because it takes time to charge the membrane. This is the firstdynamic term <strong>in</strong> the book! We write this equation <strong>in</strong> the standard “dynamical system”formC ˙V = I − I Na − I Ca − I K − I Cl (2.2)orC ˙V = I − g Na (V − E Na ) − g Ca (V − E Ca ) − g K (V − E K ) − g Cl (V − E Cl ) . (2.3)


Electrophysiology of Neurons 29If there are no additional current sources or s<strong>in</strong>ks, such as synaptic current, axialcurrent, tangential current along the membrane surface, or current <strong>in</strong>jected via anelectrode, then I = 0. In this case, the membrane potential is typically bounded bythe equilibrium potentials <strong>in</strong> the follow<strong>in</strong>g order (see Fig. 2.4):E K < E Cl < V (at rest)< E Na < E Ca ,so that I Na , I Ca < 0 (<strong>in</strong>ward currents) and I K , I Cl > 0 (outward currents). From(2.2) it follows that <strong>in</strong>ward currents <strong>in</strong>crease the membrane potential, i.e., make itmore positive (depolarization), whereas outward currents decrease it, i.e., make it morenegative (hyperpolarization). Notice that I Cl is called an outward current even thoughthe flow of Cl − ions is <strong>in</strong>ward; the ions br<strong>in</strong>g negative charge <strong>in</strong>side the membrane,which is equivalent to positively charged ions leav<strong>in</strong>g the cell, as <strong>in</strong> I K .2.1.4 Rest<strong>in</strong>g potential and <strong>in</strong>put resistanceIf there were only K + channels, as <strong>in</strong> Fig. 2.2, the membrane potential would quicklyapproach the K + equilibrium potential, E K , which is around −90 mV. Indeed,C ˙V = −I K = −g K (V − E K )<strong>in</strong> this case. However, most membranes conta<strong>in</strong> a diversity of channels. For example,Na + channels would produce an <strong>in</strong>ward current and pull the membrane potential towardsthe Na + equilibrium potential, E Na , which could be as large as +90 mV. Thevalue of the membrane potential at which all <strong>in</strong>ward and outward currents balance eachother so that the net membrane current is zero corresponds to the rest<strong>in</strong>g membranepotential. It can be found from the equation (2.3, I = 0) by sett<strong>in</strong>g ˙V = 0. Theresult<strong>in</strong>g expressionV rest = g NaE Na + g Ca E Ca + g K E K + g Cl E Clg Na + g Ca + g K + g Cl(2.4)has a nice mechanistic <strong>in</strong>terpretation: V rest is the center of mass of the balance depicted<strong>in</strong> Fig. 2.4. Incidentally, the entire equation (2.3) can be written <strong>in</strong> the formwhereC ˙V = I − g <strong>in</strong>p (V − V rest ) , (2.5)g <strong>in</strong>p = g Na + g Ca + g K + g Clis the total membrane conductance, called <strong>in</strong>put conductance. The quantity R <strong>in</strong>p =1/g <strong>in</strong>p is the <strong>in</strong>put resistance of the membrane, and it measures the asymptotic sensitivityof the membrane potential to <strong>in</strong>jected or <strong>in</strong>tr<strong>in</strong>sic currents. Indeed, from (2.5) itfollows thatV → V rest + IR <strong>in</strong>p , (2.6)


30 Electrophysiology of NeuronsRestgKgClE K-100 E -50 mV0 mV 50 mV100 mVClNaEgNag CaE CaVrestAction PotentialgNagKE KgClg Ca-100E Cl-50 mV0 mV 50 mV100 mVNaEE CaVrestFigure 2.4: Mechanistic <strong>in</strong>terpretation of the rest<strong>in</strong>g membrane potential (2.4) as thecenter of mass. Na + conductance <strong>in</strong>creases dur<strong>in</strong>g the action potential.so greater values of R <strong>in</strong>p imply greater steady-state displacement of V due to the<strong>in</strong>jection of dc current I.A remarkable property of neuronal membranes is that ionic conductances, and hencethe <strong>in</strong>put resistance, are functions of V and time. We can use (2.6) to trace an actionpotential <strong>in</strong> a quasi-static fashion, i.e., assum<strong>in</strong>g that time is frozen. When a neuronis quiescent, Na + and Ca 2+ conductances are relatively small, V rest is near E K and E Clas <strong>in</strong> Fig. 2.4,top, and so is V . Dur<strong>in</strong>g the upstroke of an action potential, the Na +or Ca 2+ conductance becomes very large, V rest is near E Na , as <strong>in</strong> Fig. 2.4,bottom, andV <strong>in</strong>creases try<strong>in</strong>g to catch V rest . This event is however quite brief, due to the reasonsexpla<strong>in</strong>ed <strong>in</strong> subsequent sections.2.1.5 Voltage-clamp and I-V relationIn the next section we will study how the membrane potential affects ionic conductancesand currents, assum<strong>in</strong>g that the potential is fixed at certa<strong>in</strong> value V c controlled by anexperimenter. To ma<strong>in</strong>ta<strong>in</strong> the membrane potential constant (clamped), one <strong>in</strong>sertsa metallic conductor to short-circuit currents along the membrane (space-clamp), andthen <strong>in</strong>jects a current proportional to the difference V c − V (voltage-clamp), as <strong>in</strong>Fig. 2.5. From (2.2) and the clamp condition ˙V = 0 it follows that the <strong>in</strong>jected currentI equals the net current generated by the membrane conductances.In a typical voltage-clamp experiment the membrane potential is held at a certa<strong>in</strong>rest<strong>in</strong>g value V c and then reset to a new value V s , as <strong>in</strong> Fig. 2.6a. The <strong>in</strong>jected membrane


Electrophysiology of Neurons 31V cVI=(V c -V)/RII(V c )analysisaxon<strong>in</strong>jectFigure 2.5: Two-wire voltage-clamp experiment on the axon: The top wire is used tomonitor the membrane potential V . The bottom wire is used to <strong>in</strong>ject the current Iproportional to the difference V c − V to keep the membrane potential at V c .(a)presteppotentialstep potentials(b)V cV sV scurrent, I (pA)2000150010005000- 500<strong>in</strong>stantaneous I-Vsteady-state I-VI 0 (V c ,V s )0 1 2 3 4 5time, msI (V s )current, I (pA)<strong>in</strong>ward outward6004002000I (V s )-200I 0 (V c ,V s )-400-100 -50 0 50membrane potential, V (mV)Figure 2.6: Voltage-clamp experiment to measure <strong>in</strong>stantaneous and steady-state I-Vrelation. Shown are simulations of the I Na +I K -model (see Fig. 4.1b); the cont<strong>in</strong>uouscurves are theoretically found I-V relations.


32 Electrophysiology of NeuronsFigure 2.7: To tease out neuronal currents, biologists employ an arsenal of sophisticated“clamp” methods, such as current-, voltage-, conductance-, and dynamic-clamp.current needed to stabilize the potential at the new value is a function of time, the prestephold<strong>in</strong>g potential V c , and the step potential V s . First, the current jumps to a newvalue to accommodate the <strong>in</strong>stantaneous voltage change from V c to V s . From (2.5) wef<strong>in</strong>d that the amplitude of the jump is g <strong>in</strong>p (V s −V c ). Then, time- and voltage-dependentprocesses start to occur and the current decreases and then <strong>in</strong>creases. The value at thenegative peak, marked by circle “o” <strong>in</strong> Fig. 2.6, depends only on V c and V s and it iscalled the <strong>in</strong>stantaneous current-voltage (I-V) relation, or I 0 (V c , V s ). The asymptotic(t → ∞) value depends only on V s and it is called the steady-state current-voltage (I-V)relation, or I ∞ (V s ).Both relations, depicted <strong>in</strong> Fig. 2.6b, can be found experimentally (black dots) ortheoretically (curves). The <strong>in</strong>stantaneous I-V relation usually has a non-monotone N-shape reflect<strong>in</strong>g non-l<strong>in</strong>ear auto-catalytic (positive feedback) transmembrane processes,which are fast enough on the time scale of the action potential so that they could beassumed to have <strong>in</strong>stantaneous k<strong>in</strong>etics. The steady-state I-V relation measures theasymptotic values of all transmembrane processes, and it may be monotone, as <strong>in</strong>the figure, or not, depend<strong>in</strong>g on the properties of the membrane currents. Both I-Vrelations provide an <strong>in</strong>valuable quantitative <strong>in</strong>formation about the currents operat<strong>in</strong>gon fast and slow time scale, and both are useful <strong>in</strong> build<strong>in</strong>g mathematical models ofneurons. F<strong>in</strong>ally, when I ∞ (V ) = 0, the net membrane current is zero, and the potentialis at rest or equilibrium, which may still be unstable as we discuss <strong>in</strong> the next chapter.2.2 ConductancesIonic channels are large transmembrane prote<strong>in</strong>s hav<strong>in</strong>g aqueous pores through whichions can flow down their electrochemical gradients. The electrical conductance of <strong>in</strong>di-


Electrophysiology of Neurons 33ExtracellularNa +voltagesensorselectivityfilterchannelprote<strong>in</strong>membrane+ + + +++ ++ ++ +++ + ++ +activationgateIntracellular<strong>in</strong>activationgateClosed(not activated)Open(activated)Closed(<strong>in</strong>activated)Figure 2.8: Structure of voltage-gated ion channels. Voltage sensors open activationgate and allow selected ions to flow through the channel accord<strong>in</strong>g to their electrochemicalgradients. The <strong>in</strong>activation gate blocks the channel (modified from Armstrong andHille 1998).vidual channels may be controlled by gat<strong>in</strong>g particles (gates), which switch the channelsbetween open and closed states. The gates are sensitive to one or more of the follow<strong>in</strong>gfactors:• Membrane potential. Example: voltage-gated Na + or K + channels.• Intracellular agents (second-messengers). Example: Ca 2+ -gated K + channels.• Extracellular agents (neurotransmitters and neuromodulators). Example: AMPA,NMDA, or GABA receptors.Despite the stochastic nature of transitions between open and closed states <strong>in</strong> <strong>in</strong>dividualchannels, the net current generated by a large population or ensemble of identicalchannels can reasonably be described by the equationI = ḡ p (V − E) (2.7)where p is the average proportion of channels <strong>in</strong> the open state, ḡ is the maximalconductance of the population, and E is the reverse potential of the current, i.e., thepotential at which the current reverses its direction. If the channels are selective for as<strong>in</strong>gle ionic species, then the reverse potential E equals the Nernst equilibrium potential(2.1) for that ionic species, see Ex. 2.2.2.1 Voltage-gated channelsWhen the gat<strong>in</strong>g particles are sensitive to the membrane potential, the channels aresaid to be voltage-gated. The gates are divided <strong>in</strong>to two types: Those that activate or


34 Electrophysiology of Neuronsm (V)1.00.80.60.40.20.0-80 -40 0 40 80V (mV)τ(V) (ms)1.61.41.21.00.80.60.40.2-120 -80 -40 0 40 80V (mV)Figure 2.9: The activation function m ∞ (V ) and the time constant τ(V ) of the fasttransient K + current <strong>in</strong> layer 5 neocortical pyramidal neurons (modified from Korngreenand Sakmann 2000).open the channels, and those that <strong>in</strong>activate or close them; see Fig. 2.8. Accord<strong>in</strong>gto the tradition <strong>in</strong>itiated <strong>in</strong> the middle of 20th century by Hodgk<strong>in</strong> and Huxley, theprobability of an activation gate to be <strong>in</strong> the open state is denoted by the variable m(sometimes the variable n is used for K + and Cl − channels). The probability of an<strong>in</strong>activation gate to be <strong>in</strong> the open state is denoted by the variable h. The proportionof open channels <strong>in</strong> a large population isp = m a h b , (2.8)where a is the number of activation gates and b is the number of <strong>in</strong>activation gates perchannel. The channels can be partially (0 < m < 1) or completely activated (m = 1);not activated or deactivated (m = 0); <strong>in</strong>activated (h = 0); released from <strong>in</strong>activationor de<strong>in</strong>activated (h = 1). Some channels do not have <strong>in</strong>activation gates (b = 0), hencep = m a . Such channels do not <strong>in</strong>activate, and they result <strong>in</strong> persistent currents. Incontrast, channels that do <strong>in</strong>activate result <strong>in</strong> transient currents.Below we describe voltage- and time-dependent k<strong>in</strong>etics of gates. This descriptionis often referred to as the Hodgk<strong>in</strong>-Huxley gate model of membrane channels.2.2.2 Activation of persistent currentsThe dynamics of the activation variable m is described by the first-order differentialequationṁ = (m ∞ (V ) − m)/τ(V ) (2.9)where the voltage-sensitive steady-state activation function m ∞ (V ) and the time constantτ(V ) can be measured experimentally: They have sigmoid and unimodal shapes,respectively, as <strong>in</strong> Fig. 2.9; see also Fig. 2.20. The steady-state activation functionm ∞ (V ) gives the asymptotic value of m when the potential is fixed (voltage-clamp).Smaller values of τ(V ) result <strong>in</strong> faster dynamics of m.


Electrophysiology of Neurons 35-30 mV(a)-60 mVpresteppotential V 0-100 mVstep potentialsV scurrent, I (pA)0-20-40-60I s(c)I(V)-80(b)I(t)-5 pAV s-100-80 -60 -40 -20 0membrane voltage, V (mV)10.5(e)activationactivationI s-85 pAm sgat<strong>in</strong>g varitable, m10.80.60.40.2m s(d)m (V)m(t)00 2 4 6 8 10time (ms)V s0-80 -60 -40 -20 0membrane voltage, V (mV)Figure 2.10: An experiment to determ<strong>in</strong>e m ∞ (V ). Shown are simulations of the persistentNa + current <strong>in</strong> Purk<strong>in</strong>je cells (see Sect. 2.3.5).


36 Electrophysiology of Neurons10.8h sm (V)0.60.40.2V sh (V)0-80 -60 -40 -20 0membrane voltage, V (mV)Figure 2.11: Steady-state activation functionm ∞ (V ) from Fig. 2.10, <strong>in</strong>activation functionh ∞ (V ), and values h s from Fig. 2.12. Theiroverlap (shaded region) produces a noticeablepersistent “w<strong>in</strong>dow” current.In Fig. 2.10 we depict a typical experiment to determ<strong>in</strong>e m ∞ (V ) of a persistentcurrent, i.e., a current hav<strong>in</strong>g no <strong>in</strong>activation variable. Initially we hold the membranepotential at a hyperpolarized value V 0 so that all activation gates are closed and I ≈ 0.Then we step-<strong>in</strong>crease V to a greater value V s (s = 1, . . . , 7; see Fig. 2.10a) and holdit there until the current is essentially equal to its asymptotic value, which we denotehere as I s (s stands for “step”; see Fig. 2.10b). Repeat<strong>in</strong>g the experiment for variousstepp<strong>in</strong>g potentials V s , one can easily determ<strong>in</strong>e the correspond<strong>in</strong>g I s , and hence theentire steady-state I-V relation, which we depict <strong>in</strong> Fig. 2.10c. Accord<strong>in</strong>g to (2.7),I(V ) = ḡm ∞ (V )(V − E), and the steady-state activation curve m ∞ (V ) depicted <strong>in</strong>Fig. 2.10d is just I(V ) divided by the driv<strong>in</strong>g force (V − E) and normalized so thatmax m ∞ (V ) = 1. To determ<strong>in</strong>e the time constant τ(V ), one needs to analyze theconvergence rates. In Ex. 6 we describe an efficient method to determ<strong>in</strong>e m ∞ (V ) andτ(V ).2.2.3 Inactivation of transient currentsThe dynamics of the <strong>in</strong>activation variable h can also be described by the first-orderdifferential equationḣ = (h ∞ (V ) − h)/τ(V ) . (2.10)where h ∞ (V ) is the voltage-sensitive steady-state <strong>in</strong>activation function depicted <strong>in</strong>Fig. 2.11. In Fig. 2.12 we present a typical voltage-clamp experiment to determ<strong>in</strong>eh ∞ (V ) <strong>in</strong> the presence of activation m ∞ (V ). It relies on the observation that <strong>in</strong>activationk<strong>in</strong>etics is usually slower than activation k<strong>in</strong>etics. First, we hold the membranepotential at a certa<strong>in</strong> pre-step potential V s for a sufficiently long time so that the activationand <strong>in</strong>activation variables are essentially equal to their steady-state values m ∞ (V s )and h ∞ (V s ), respectively, which have yet to be determ<strong>in</strong>ed. Then we step-<strong>in</strong>crease Vto a sufficiently high value V 0 chosen so that m ∞ (V 0 ) ≈ 1. If activation is much fasterthan <strong>in</strong>activation, m approaches 1 after a first few milliseconds while h cont<strong>in</strong>ues tobe near its asymptotic value h s = h ∞ (V s ), which can be found from the peak valueof the current I s ≈ ḡ · 1 · h s (V s − E). Repeat<strong>in</strong>g this experiment for various pre-steppotentials, one can determ<strong>in</strong>e the steady-state <strong>in</strong>activation curve h ∞ (V ) <strong>in</strong> Fig. 2.11.


Electrophysiology of Neurons 37pre-steppotentials-20 mVV s-80 mVstep potentialV 0I(t)peakI s<strong>in</strong>activation10.5010.50h s<strong>in</strong>stantaneousactivation0 10 20 30 40 50 60 70time (ms)m(t)h(t)Figure 2.12: Dynamics of the current(I), activation (m) and <strong>in</strong>activation(h) variables <strong>in</strong> the voltage-clampexperiment aimed at measur<strong>in</strong>g h ∞ (V )<strong>in</strong> Fig. 2.11.In Ex. 6 we describe a better method to determ<strong>in</strong>e h ∞ (V ) that does not rely on thedifference between the activation and <strong>in</strong>activation time scales.The voltage-sensitive steady-state activation and <strong>in</strong>activation functions overlap <strong>in</strong>a shaded w<strong>in</strong>dow depicted <strong>in</strong> Fig. 2.11. Depend<strong>in</strong>g on the size of the shaded area <strong>in</strong>the figure, the overlap may result <strong>in</strong> a noticeable “w<strong>in</strong>dow” current.2.2.4 Hyperpolarization-activated channelsMany neurons <strong>in</strong> various parts of the bra<strong>in</strong> have channels that are opened by hyperpolarization.These channels produce currents that are turned on by hyperpolarizationand turned off by depolarization. Biologists refer to such currents as be<strong>in</strong>g“exceptional” or “weird”, and denoted them as I Q (“queer”), I f (“funny”), I h(“hyperpolarization-activated”), or I Kir (K + <strong>in</strong>ward rectifier). We will consider the lasttwo currents <strong>in</strong> detail <strong>in</strong> the next chapter. Most neuroscience textbooks classify thesecurrents <strong>in</strong> a special category — hyperpolarization-activated currents. However, fromthe theoretical po<strong>in</strong>t of view it is <strong>in</strong>convenient to create special categories. In this bookwe treat these currents as “normal” transient currents with the understand<strong>in</strong>g that theyare always activated (either a = 0 or variable m = 1 <strong>in</strong> (2.8)), but can be <strong>in</strong>activatedby depolarization (variable h → 0) or de<strong>in</strong>activated by hyperpolarization (variableh → 1). Moreover, there is biophysical evidence suggest<strong>in</strong>g that clos<strong>in</strong>g/open<strong>in</strong>g ofI Kir is <strong>in</strong>deed related to the <strong>in</strong>activation/de<strong>in</strong>activation process (Lopat<strong>in</strong> et al. 1994).


38 Electrophysiology of Neurons2.3 Hodgk<strong>in</strong>-Huxley ModelIn Sect. 2.1 we have studied how the membrane potential depends on the membranecurrents assum<strong>in</strong>g that ionic conductances are fixed. In Sect. 2.2 we have used theHodgk<strong>in</strong>-Huxley gate model to study how the conductances and currents depend onthe membrane potential assum<strong>in</strong>g that the potential is clamped at different values.In this section we put it all together and study how the potential ↔ current nonl<strong>in</strong>ear<strong>in</strong>teractions lead to many <strong>in</strong>terest<strong>in</strong>g phenomena such as generation of actionpotentials.2.3.1 Hodgk<strong>in</strong>-Huxley equationsOne of the most important models <strong>in</strong> computational neuroscience is the Hodgk<strong>in</strong>-Huxley model of the squid giant axon. Us<strong>in</strong>g pioneer<strong>in</strong>g experimental techniques of thattime, Hodgk<strong>in</strong> and Huxley (1952) determ<strong>in</strong>ed that the squid axon curries three majorcurrents: voltage-gated persistent K + current with four activation gates (result<strong>in</strong>g <strong>in</strong>the term n 4 <strong>in</strong> the equation below, where n is the activation variable for K + ), voltagegatedtransient Na + current with three activation gates and one <strong>in</strong>activation gate (termm 3 h below), and Ohmic leak current, I L , which is carried mostly by Cl − ions. Thecomplete set of space-clamped Hodgk<strong>in</strong>-Huxley equations isC ˙V = I −I K{ }} {ḡ K n 4 (V − E K ) −ṅ = α n (V )(1 − n) − β n (V )nṁ = α m (V )(1 − m) − β m (V )mḣ = α h (V )(1 − h) − β h (V )h ,I Na{ }} {ḡ Na m 3 h(V − E Na ) −I L{ }} {g L (V − E L )where10 − Vα n (V ) = 0.01exp( 10−V ) − 1 ,10( ) −Vβ n (V ) = 0.125 exp ,8025 − Vα m (V ) = 0.1exp( 25−V− 1 ,10( ) −Vβ m (V ) = 4 exp ,18( ) −Vα h (V ) = 0.07 exp ,201β h (V ) =exp( 30−V ) + 1 .10


Electrophysiology of Neurons 391h (V)8τ (V) hn (V)m (V)τ (V) n0-40 0 100V (mV)τm(V)0-40 0 100V (mV)Figure 2.13: Steady-state (<strong>in</strong>)activation functions (left) and voltage-dependent timeconstants (right) <strong>in</strong> the Hodgk<strong>in</strong>-Huxley model.These parameters, provided <strong>in</strong> the orig<strong>in</strong>al Hodgk<strong>in</strong> and Huxley paper, correspond tothe membrane potential shifted by approximately 65 mV so that the rest<strong>in</strong>g potentialis at V ≈ 0. Hodgk<strong>in</strong> and Huxley did that for the sake of convenience, but the shifthas led to a lot of confusion over the years. The shifted Nernst equilibrium potentialsareE K = −12 mV , E Na = 120 mV , E L = 10.6 mV;see also Ex. 1. Typical values of maximal conductances areḡ K = 36 mS/cm 2 , ḡ Na = 120 mS/cm 2 , g L = 0.3 mS/cm 2 .C = 1 µF/cm 2 is the membrane capacitance and I = 0 µA/cm 2 is the applied current.The functions α(V ) and β(V ) describe the transition rates between open and closedstates of the channels. We present this notation only for historical reasons. In the restof the book, we use the standard formwhereṅ = (n ∞ (V ) − n)/τ n (V ) ,ṁ = (m ∞ (V ) − m)/τ m (V ) ,ḣ = (h ∞ (V ) − h)/τ h (V ) ,n ∞ = α n /(α n + β n ) , τ n = 1/(α n + β n ) ,m ∞ = α m /(α m + β m ) , τ m = 1/(α m + β m ) ,h ∞ = α h /(α h + β h ) , τ h = 1/(α h + β h )are depicted <strong>in</strong> Fig. 2.13. These functions can be approximated by the Boltzmann andGaussian functions; see Ex. 4. We also shift the membrane potential back to its truevalue, so that the rest<strong>in</strong>g state is near -65 mV.The membrane of the squid giant axon carries only two major currents: transientNa + and persistent K + . Most neurons <strong>in</strong> the central nervous system have additionalcurrents with diverse activation and <strong>in</strong>activation dynamics, which we summarize <strong>in</strong>Sect. 2.3.5. The Hodgk<strong>in</strong>-Huxley formalism is the most accepted model to describetheir k<strong>in</strong>etics.


40 Electrophysiology of NeuronsFigure 2.14: Studies of spike-generation mechanism <strong>in</strong> “giant squid axons” won AlanHodgk<strong>in</strong> and Andrew Huxley the 1963 Nobel prize for physiology or medic<strong>in</strong>e (sharedwith Sir John Eccles); see also Fig. 4.1 <strong>in</strong> Keener and Sneyd (1998).S<strong>in</strong>ce we are <strong>in</strong>terested <strong>in</strong> geometrical and qualitative methods of analysis of neuronalmodels, we assume that all variables and parameters have appropriate scalesand dimensions, but we do not explicitly state them. An exception is the membranepotential V , whose mV scale is stated <strong>in</strong> every figure.2.3.2 Action PotentialRecall that when V = V rest , which is 0 mV <strong>in</strong> the Hodgk<strong>in</strong>-Huxley model, all <strong>in</strong>wardand outward currents balance each other so the net current is zero, as <strong>in</strong> Fig. 2.15.The rest state is stable: A small pulse of current applied via I(t) produces a smallpositive perturbation of the membrane potential (depolarization), which results <strong>in</strong> asmall net current that drives V back to rest (repolarization). However, an <strong>in</strong>termediatesize pulse of current produces a perturbation that is amplified significantly becausemembrane conductances depend on V . Such a non-l<strong>in</strong>ear amplification causes V todeviate considerably from V rest — a phenomenon referred to as an action potential orspike.In Fig. 2.15 we show a typical time course of an action potential <strong>in</strong> the Hodgk<strong>in</strong>-Huxley system. Strong depolarization <strong>in</strong>creases activation variables m and n and decreases<strong>in</strong>activation variable h. S<strong>in</strong>ce τ m (V ) is relatively small, variable m is relativelyfast. Fast activation of Na + conductance drives V toward E Na , result<strong>in</strong>g <strong>in</strong> furtherdepolarization and further activation of g Na . This positive feedback loop, depicted <strong>in</strong>Fig. 2.16 results <strong>in</strong> the upstroke of V . While V moves toward E Na , the slower gat<strong>in</strong>gvariables catch up. Variable h → 0, caus<strong>in</strong>g <strong>in</strong>activation of Na + current, and variablen → 1, caus<strong>in</strong>g slow activation of outward K + current. The latter and the leak current


Electrophysiology of Neurons 41E Na10050E L0E KMembrane voltage (mV)smalldepolarizationRestRepolarizationExcited(regenerative)largedepolarizationUpstrokeActionpotential(spike)RepolarizationAbsoluterefractorydepolarizationresthyperpolarizationAfter-hyperpolarizationRelativerefractoryV(t)10.50Gat<strong>in</strong>g variablesactivationh(t)n(t)m(t)<strong>in</strong>activationdeactivationde<strong>in</strong>activation20Conductances (mS/cm2)g (t)Nag (t)K05000-500Currents( µ A/cm 2)I Na+I K+ILI (t)KI (t)Na200Applied current( µ A/cm 2)0 2 4 6 8 10 12 14 16 18 20time (ms)Figure 2.15: Action potential <strong>in</strong> the Hodgk<strong>in</strong>-Huxley model.I(t)


42 Electrophysiology of NeuronsDepolarizationIncrease <strong>in</strong>gNaIncrease<strong>in</strong> gNaIncrease<strong>in</strong> gKNa +InflowDepolarizationRepolarizationHyperpolarizationFigure 2.16: Positive and negative feedback loops result<strong>in</strong>g <strong>in</strong> excited (regenerative)behavior <strong>in</strong> neurons.repolarize the membrane potential toward V rest .When V is near V rest , the voltage-sensitive time constants τ n (V ) and τ h (V ) arerelatively large, as one can see <strong>in</strong> Fig. 2.13. Therefore, recovery of variables n and his slow. In particular, outward K + current cont<strong>in</strong>ues to be activated (n is large) evenafter the action potential downstroke, thereby caus<strong>in</strong>g V to go below V rest toward E K— a phenomenon known as afterhyperpolarization.In addition, Na + current cont<strong>in</strong>ues to be <strong>in</strong>activated (h is small) and not availablefor any regenerative function. The Hodgk<strong>in</strong>-Huxley system cannot generate anotheraction potential dur<strong>in</strong>g this absolute refractory period. While the current de<strong>in</strong>activates,the system becomes able to generate an action potential provided that the stimulus isrelatively strong (relative refractory period).To study the relationship between these refractory periods, we stimulate the Hodgk<strong>in</strong>-Huxley model with 1-ms pulses of current hav<strong>in</strong>g various amplitudes and latencies. Them<strong>in</strong>imal amplitude of the stimulation needed to evoke a second spike <strong>in</strong> the model isdepicted <strong>in</strong> Fig. 2.17, bottom. Notice that around 14 ms after the first spike, themodel is hyper-excitable, that is, the stimulation amplitude is less than the basel<strong>in</strong>eamplitude A p ≈ 6 needed to evoke a spike from the rest<strong>in</strong>g state. This occurs becausethe Hodgk<strong>in</strong>-Huxley model exhibits damped oscillations of membrane potential, whichwe discuss <strong>in</strong> Chap. 7.2.3.3 Propagation of the action potentialsThe space-clamped Hodgk<strong>in</strong>-Huxley model of squid giant axon describes non-propagat<strong>in</strong>gaction potentials s<strong>in</strong>ce V (t) does not depend on the location, x, along the axon. To describepropagation of action potentials (pulses) along the axon hav<strong>in</strong>g potential V (x, t),radius a (cm) and <strong>in</strong>tracellular resistivity R (Ω·cm), the partial derivative V xx is addedto the voltage equation to account for axial currents along the membrane. The result<strong>in</strong>gnon-l<strong>in</strong>ear parabolic partial differential equationC V t = a2R V xx + I − I K − I Na − I Lis often referred to as the Hodgk<strong>in</strong>-Huxley cable or propagat<strong>in</strong>g equation. Its importanttype of solution, a travel<strong>in</strong>g pulse, is depicted <strong>in</strong> Fig. 2.18. Study<strong>in</strong>g this equation goes


Electrophysiology of Neurons 43pulse size, Ap, needed toproduce 2nd spike (µA/cm2) memrane potential (mV)10050020100absoluterefractoryrelativerefractoryhyperexcitability0 5 10 15 20 25time, tp, after 1st spike (ms)tpApFigure 2.17: Refractory periods <strong>in</strong> the Hodgk<strong>in</strong>-Huxley model with I = 3.beyond the scope of this book, and the reader can consult Keener and Sneyd (1998)and references there<strong>in</strong>.2.3.4 Dendritic compartmentsModifications of the Hodgk<strong>in</strong>-Huxley model, often called Hodgk<strong>in</strong>-Huxley-type models,or conductance-based models, can describe the dynamics of spike-generation of manyif not all neurons recorded <strong>in</strong> nature. However, there is more to the computationalproperty of neurons than just the spike-generation mechanism. Many neurons havean extensive dendritic tree that can sample the synaptic <strong>in</strong>put arriv<strong>in</strong>g at differentlocations and <strong>in</strong>tegrate it over space and time.Many dendrites have voltage-gated currents, so the synaptic <strong>in</strong>tegration is nonl<strong>in</strong>ear,sometimes result<strong>in</strong>g <strong>in</strong> dendritic spikes that can propagate forward to the somaof the neuron or backwards to distant dendritic locations. Dendritic spikes are prom<strong>in</strong>ent<strong>in</strong> <strong>in</strong>tr<strong>in</strong>sically burst<strong>in</strong>g (IB) and chatter<strong>in</strong>g (CH) neocortical neurons considered<strong>in</strong> Chap. 8. In that chapter we also model regular spik<strong>in</strong>g (RS) pyramidal neurons,the most numerous class of neurons <strong>in</strong> mammalian neocortex, and show that theirspike-generation mechanism is one of the simplest. The computation complexity of RSneurons must be hidden then <strong>in</strong> the arbors of their dendritic trees.It is not feasible at present to study analytically or geometrically the dynamics ofmembrane potential <strong>in</strong> dendritic trees, unless dendrites are assumed to be passive (l<strong>in</strong>-


44 Electrophysiology of NeuronsV(x,t )1V(x,t )2V(x,t )3V(x,t )4100Travel<strong>in</strong>g pulseMembrane voltage (mV)x0AxonFigure 2.18: Travel<strong>in</strong>g pulse solution of the Hodgk<strong>in</strong>-Huxley cable equation at foursuccessive moments.(a) (b) (c) (d)neuron 1neuron 2dendriteVdsomaVsFigure 2.19: Dendritic tree of a neuron (a) is replaced by a network of compartments(b), each modeled by a Hodgk<strong>in</strong>-Huxley-type model. Two-compartment neuronalmodel (c) may be equivalent to two neurons coupled via gap junctions (electricalsynapse).


Electrophysiology of Neurons 45ear), semi-<strong>in</strong>f<strong>in</strong>ite, and satisfy Rall’s branch<strong>in</strong>g law (Rall 1959). Much of the <strong>in</strong>sightcan be obta<strong>in</strong>ed via simulations, which typically substitute the cont<strong>in</strong>uous dendriticstructure <strong>in</strong> Fig. 2.19a by a network of discrete compartments <strong>in</strong> Fig. 2.19b. Dynamicsof each compartment is simulated by a Hodgk<strong>in</strong>-Huxley-type model, and the compartmentsare coupled via conductances. For example, if V s and V d denote the membranepotential at the soma and <strong>in</strong> the dendritic tree, as <strong>in</strong> Fig. 2.19c, thenC s ˙Vs = −I s (V s , t) + g s (V d − V s ) , and C d ˙Vd = −I d (V d , t) + g d (V s − V d ) ,where each I(V, t) represents the sum of all voltage-, Ca 2+ -, and time-dependent currents<strong>in</strong> the compartment, and g s and g d are the coupl<strong>in</strong>g conductances that dependon the relative sizes of dendritic and somatic compartments. One can obta<strong>in</strong> manyspik<strong>in</strong>g and burst<strong>in</strong>g patterns by chang<strong>in</strong>g the conductances and keep<strong>in</strong>g all the otherparameters fixed (P<strong>in</strong>sky and R<strong>in</strong>zel 1994, Ma<strong>in</strong>en and Sejnowski 1996).Once we understand how to couple two compartments, we can do it for hundreds orthousands of compartments. GENESIS and NEURON simulation environments couldbe useful here, especially s<strong>in</strong>ce they conta<strong>in</strong> databases of dendritic trees reconstructedfrom real neurons.Interest<strong>in</strong>gly, the somatic-dendritic pair <strong>in</strong> Fig. 2.19c is equivalent to a pair ofneurons <strong>in</strong> Fig. 2.19d coupled via gap-junctions. These are electrical contacts thatallow ions and small molecules to pass freely between the cells. Gap junctions areoften called electrical synapses, because they allow potentials to be conducted directlyfrom one neuron to another.Computational study of multi-compartment dendritic process<strong>in</strong>g is outside of thescope of this book. We consider multi-compartment models of cortical pyramidal neurons<strong>in</strong> Chap. 8 and gap-junction coupled neurons <strong>in</strong> Chap. 10.2.3.5 Summary of voltage-gated currentsThroughout this book we model k<strong>in</strong>etics of various voltage-sensitive currents us<strong>in</strong>g theHodgk<strong>in</strong>-Huxley gate modelI = ḡ m a h b (V − E)whereI (µA/cm 2 ) currentV (mV) membrane voltageE (mV) reverse potentialḡ (mS/cm 2 ) maximal conductancemprobability of activation gate to be openhprobability of <strong>in</strong>activation gate to be openathe number of activation gates per channelbthe number of <strong>in</strong>activation gates per channelThe gat<strong>in</strong>g variables m and n satisfy l<strong>in</strong>ear first order differential equations (2.9) and(2.10), respectively. We approximate the steady-state activation curve m ∞ (V ) by the


46 Electrophysiology of Neurons1m (V)slopeC base+Campτ(V)0.502kV 1/2VC ampCσ C amp/ebaseV maxVFigure 2.20: Boltzmann (2.11) and Gaussian (2.12) functions and geometrical <strong>in</strong>terpretationsof their parameters.Boltzmann function depicted <strong>in</strong> Fig. 2.20,m ∞ (V ) =11 + exp {(V 1/2 − V )/k}(2.11)The parameter V 1/2 satisfies m ∞ (V 1/2 ) = 0.5, and k is the slope factor (negative for<strong>in</strong>activation curve h ∞ (V )). Smaller values of |k| result <strong>in</strong> steeper m ∞ (V ).The voltage-sensitive time constant τ(V ) can be approximated by the Gaussianfunctionτ(V ) = C base + C amp exp −(V max − V ) 2σ 2 , (2.12)see Fig. 2.20. The graph of the function is above C base with amplitude C amp . Themaximal value is achieved at V max . The parameter σ measures the characteristic widthof the graph, i.e., τ(V max ± σ) = C base + C amp /e. The Gaussian description is often notadequate, so we substitute it by other functions whenever appropriate.Below is the summary of voltage-gated currents whose k<strong>in</strong>etics were measured experimentally.The division <strong>in</strong>to persistent and transient is somewhat artificial s<strong>in</strong>cemost “persistent” currents can still <strong>in</strong>activate after seconds of prolonged depolarization.Hyperpolarization-activated currents, such as the h-current or K + <strong>in</strong>wardly rectify<strong>in</strong>gcurrent, are mathematically equivalent to currents that are always activated, but canbe <strong>in</strong>activated by depolarization. To avoid possible confusion we mark these currentsas “opened by hyperpolarization”.


Electrophysiology of Neurons 47Parameters (Fig. 2.20)Na + currents Eq. 2.11 Eq. 2.12V 1/2 k V max σ C amp C baseFast transient 1 I Na,t = ḡ m 3 h(V − E Na )activation −40 15 −38 30 0.46 0.04<strong>in</strong>activation −62 −7 −67 20 7.4 1.2Fast transient 2 I Na,t = ḡ m ∞ (V )h(V − E Na )activation −30 5.5 − − − −<strong>in</strong>activation −70 −5.8 τ h (V ) = 3 exp((−40 − V )/33)Fast transient 3 I Na,t = ḡ m ∞ (V )h(V − E Na )activation −28 6.7 − − − −<strong>in</strong>activation −66 −6 τ h (V ) = 4 exp((−30 − V )/29)Fast persistent 4,a I Na,p = ḡ m ∞ (V )h(V − E Na )activation −50 4 − − − −<strong>in</strong>activation −49 −10 −66 35 4.5 sec 2 secFast persistent 5,a I Na,p = ḡ m ∞ (V )(0.14 + 0.86h)(V − E Na )activation −50 6 − − − −<strong>in</strong>activation −56 −7 τ h (V ) = 63.2 + 25 exp(−V/25.5)Fast persistent 2 I Na,p = ḡ m(V − E Na )activation −54 9 − − − 0.8Fast persistent 6 I Na,p = ḡ m(V − E Na )activation −42 4 − − − 0.81. Squid giant axon (Hodgk<strong>in</strong> and Huxley 1954); see Ex. 4.2. Thalamocortical neurons <strong>in</strong> rats (Parri and Crunelli 1999).3. Thalamocortical neurons <strong>in</strong> cats (Parri and Crunelli 1999).4. Layer-II pr<strong>in</strong>cipal neurons <strong>in</strong> entorh<strong>in</strong>al cortex (Magistretti and Alonso 1999).5. Large dorsal root ganglion neurons <strong>in</strong> rats (Baker and Bostock 1997, 1998).6. Purk<strong>in</strong>je cells (Kay et al. 1998).a Very slow <strong>in</strong>activation.


48 Electrophysiology of NeuronsParameters (Fig. 2.20)K + currents Eq. 2.11 Eq. 2.12V 1/2 k V max σ C amp C baseDelayed rectifier 1 I K = ḡ n 4 (V − E K )activation −53 15 −79 50 4.7 1.1Delayed rectifier 2,4 I K = ḡ mh(V − E K )activation −3 10 −50 30 47 5<strong>in</strong>activation −51 −12 −50 50 1000 360M current 3 I K(M) = ḡ m(V − E K )activation −44 8 −50 25 320 20Transient 4 I A = ḡ mh(V − E K )activation −3 20 −71 60 0.92 0.34<strong>in</strong>activation −66 −10 −73 23 50 8Transient 5 I A = ḡ mh(V − E K )activation −26 20 − − − −<strong>in</strong>activation −72 −9.6 − − − 15.5Transient 6 I A = ḡ m 4 h (V − E K )Fast component (60% of total conductance)activation −60 8.5 −58 25 2 0.37<strong>in</strong>activation −78 −6 −78 25 45 19Slow component (40% of total conductance)activation −36 20 −58 25 2 0.37<strong>in</strong>activation −78 −6 −78 25 45 19τ h (V ) = 60 when V > −73Inward rectifier 7 I Kir = ḡ h ∞ (V )(V − E K )(opened by hyperpolarization )<strong>in</strong>activation −80 −12 − − − < 11. Squid giant axon (Hodgk<strong>in</strong> and Huxley 1954); see Ex. 4.2. Neocortical pyramidal neurons (Bekkers 2000).3. Rodent neuroblastoma-glioma hybrid cells (Robb<strong>in</strong>s et al. 1992).4. Neocortical pyramidal neurons (Korngreen and Sakmann 2000).5. Hippocampal mossy fiber boutons (Geiger and Jonas 2000).6. Thalamic relay neurons (Huguenard and McCormick 1992).7. Horizontal cells <strong>in</strong> catfish ret<strong>in</strong>a (Dong and Werbl<strong>in</strong> 1995); AP cell of leech (Wessel et al.1999); rat locus coeruleus neurons (Williams et al. 1988, V 1/2 = E K ).


Electrophysiology of Neurons 49Parameters (Fig. 2.20)Cation currents Eq. 2.11 Eq. 2.12V 1/2 k V max σ C amp C baseI h current 1 I h = ḡ h (V − E h ), E h = −43 mV(opened by hyperpolarization )<strong>in</strong>activation −75 −5.5 −75 15 1000 100I h current 2 I h = ḡ h (V − E h ), E h = −1 mV<strong>in</strong>act. (soma) −82 −9 −75 20 50 10<strong>in</strong>act. (dendrite) −90 −8.5 −75 20 40 10I h current 3 I h = ḡ h (V − E h ), E h = −21 mVfast <strong>in</strong>act. (65%) −67 −12 −75 30 50 20slow <strong>in</strong>act. (35%) −58 −9 −65 30 300 1001. Thalamic relay neurons (McCormick and Pape 1990; Huguenard and McCormick 1992).2. Hippocampal pyramidal neurons <strong>in</strong> CA1 (Magee 1998).3. Entorh<strong>in</strong>al cortex Layer II neurons (Dickson et al. 2000).1000slowI hvoltage-gated currentsactivation<strong>in</strong>activationI hI K(M)100I K(M)time constant, (ms)10IAI AI K(M)I h IhI AI KI AI NatI NatI NatI KI NatI A1fastI KirI NapI Naplow-threshold-100 -50high-threshold0half-voltage, V 1/2 (mV)Figure 2.21: Summary of current k<strong>in</strong>etics. Each oval (rectangle) denotes the voltageand temporal scale of activation (<strong>in</strong>activation) of a current. Transient currents arerepresented by arrows connect<strong>in</strong>g ovals and rectangles.


50 Electrophysiology of NeuronsFigure 2.22: Alan Hodgk<strong>in</strong> (right) and Andrew Huxley (left) <strong>in</strong> their Plymouth Mar<strong>in</strong>eLab <strong>in</strong> 1949 (photo was k<strong>in</strong>dly provided by National Mar<strong>in</strong>e Biological Library,Plymouth, UK).Review of Important Concepts• Electrical signals <strong>in</strong> neurons are carried by Na + , Ca 2+ , K + , and Cl −ions, which move through membrane channels accord<strong>in</strong>g to theirelectrochemical gradients.• The membrane potential V is determ<strong>in</strong>ed by the membrane conductancesg i and correspond<strong>in</strong>g reversal potentials E i :C ˙V = I − ∑ ig i · (V − E i ) .• Neurons are excitable because the conductances depend on the membranepotential and time.• The most accepted description of k<strong>in</strong>etics of voltage-sensitive conductancesis the Hodgk<strong>in</strong>-Huxley gate model.• Voltage-gated activation of <strong>in</strong>ward Na + or Ca 2+ current depolarizes(<strong>in</strong>creases) the membrane potential.• Voltage-gated activation of outward K + or Cl − current hyperpolarizes(decreases) the membrane potential.• An action potential or spike is a brief regenerative depolarization ofthe membrane potential followed by its repolarization and possiblyhyperpolarization, as <strong>in</strong> Fig. 2.16.


Electrophysiology of Neurons 51Bibliographical NotesOur summary of the membrane electrophysiology is limited: We present only those conceptsthat are necessary to understand the Hodgk<strong>in</strong>-Huxley description of generation ofaction potentials. We have omitted such important topics as Goldman-Hodgk<strong>in</strong>-Katzequation, cable theory, dendritic and synaptic function, etc., although some of thosewill be <strong>in</strong>troduced later <strong>in</strong> the book.The standard textbook on membrane electrophysiology is the second edition ofIon Channels of Excitable Membranes by B. Hille (2001). An excellent <strong>in</strong>troductorytextbook with an emphasis on the quantitative approach is Foundations of CellularNeurophysiology by D. Johnston and S. Wu (1995). A detailed <strong>in</strong>troduction <strong>in</strong>to mathematicalaspects of cellular biophysics can be found <strong>in</strong> Mathematical Physiology byJ. Keener and J. Sneyd (1998). The latter two books complement rather than repeateach other. Biophysics of Computation by Koch (1999) and Chapters 5 and 6 of Theoretical<strong>Neuroscience</strong> by Dayan and Abbott (2001) provide a good <strong>in</strong>troduction <strong>in</strong>tobiophysics of excitable membranes.The first book devoted exclusively to dendrites is Dendrites by Stuart et al. (1999).It emphasizes the active nature of dendritic dynamics. Arshavsky et al. (1971; Russianlanguage edition - 1969) make the first and probably still the best theoretical attempt tounderstand the neuro-computational properties of branch<strong>in</strong>g dendritic trees endowedwith voltage-gated channels and capable of generat<strong>in</strong>g action potentials. Had theypublished their results <strong>in</strong> the 90s, they would have been considered classics <strong>in</strong> the field;Unfortunately, the computational neuroscience community of the 70s was not ready toaccept the “heretic” idea that dendrites can fire spikes, that spikes can propagate backand forward along the dendritic tree, that EPSPs can be scaled-up with distance, that<strong>in</strong>dividual dendritic branches can perform co<strong>in</strong>cidence detection and branch<strong>in</strong>g po<strong>in</strong>tscan perform non-l<strong>in</strong>ear summation, and that different and <strong>in</strong>dependent computationscan be carried out at different parts of the neuronal dendritic tree. We touch some ofthese issues <strong>in</strong> Chap. 8.Exercises1. Determ<strong>in</strong>e Nernst equilibrium potentials for the membrane of the squid giantaxon us<strong>in</strong>g the follow<strong>in</strong>g dataand T = 20 ◦ C.Inside (mM) Outside (mM)K + 430 20Na + 50 440Cl − 65 5602. Show that a non-selective cation currentI = ḡ Na p (V − E Na ) + ḡ K p (V − E K )


52 Electrophysiology of Neuronsvoltage steps-100 mV-60 mV0 mV-10 mVcurrentstime (ms)0 1 2 3 4 5Figure 2.23: Current tracescorrespond<strong>in</strong>g to voltage stepsof various amplitudes; see Ex. 6.can be written <strong>in</strong> the form (2.7) withḡ = ḡ Na + ḡ K and E = ḡNaE Na + ḡ K E Kḡ Na + ḡ K.3. Show that apply<strong>in</strong>g a dc-current I <strong>in</strong> the neuronal modelC ˙V = I − g L (V − E L ) − I other (V )is equivalent to chang<strong>in</strong>g the leak reverse potential E L .4. Steady-state (<strong>in</strong>)activation curves and voltage-sensitive time constants can beapproximated by the Boltzmann (2.11) and Gaussian (2.12) functions, respectively,depicted <strong>in</strong> Fig. 2.20. Expla<strong>in</strong> the mean<strong>in</strong>g of the parameters V 1/2 , k,C base , C amp , V max and σ and f<strong>in</strong>d their values that provide satisfactory fit near therest state V = 0 to the Hodgk<strong>in</strong>-Huxley functions depicted <strong>in</strong> Fig. 2.13.5. (Willms et al. 1999) Consider the curve m p ∞(V ), where m ∞ (V ) is the Boltzmannfunction with parameters V 1/2 and k, and p > 1. This curve can be approximatedby another Boltzmann function with some parameters Ṽ1/2 and ˜k (and p = 1).F<strong>in</strong>d the formulae that relate Ṽ1/2 and ˜k to V 1/2 , k, and p.6. (Willms et al. 1999) Write a MATLAB program that determ<strong>in</strong>es activationand <strong>in</strong>activation parameters via a simultaneous fitt<strong>in</strong>g of current traces from avoltage-clamp experiment similar to the one <strong>in</strong> Fig. 2.23. Assume that the valuesof the voltage pairs, e.g., −60, −10; −100, 0 (mV); are <strong>in</strong> the file v.dat. Thevalues of the current (circles <strong>in</strong> Fig. 2.23) are <strong>in</strong> the file current.dat, and thesampl<strong>in</strong>g times, e.g., 0, 0.25, 0.5, 1, 1.5, 2, 3, 5 (ms), are <strong>in</strong> the file times.dat.7. Modify the MATLAB program from the previous exercise to handle multi-step(Fig. 2.24) and ramp protocols.8. [M.S.] F<strong>in</strong>d the best sequence of step potentials that can determ<strong>in</strong>e activationand <strong>in</strong>activation parameters (a) <strong>in</strong> the shortest time, (b) with the highestprecision.


Electrophysiology of Neurons 53+50 mV-10 mV 10 ms-40 mV-60 mV-80 mV-100 mV-20 mV Figure 2.24: Multiple voltage steps are oftenneeded to determ<strong>in</strong>e time constants of<strong>in</strong>activation; see Ex. 7.9. [M.S.] Modify the MATLAB program from Ex. 6 to handle multiple currents.10. [M.S.] Add a PDE solver to the MATLAB program from exercise 6 to simulatepoor space and voltage clamp conditions.11. [Ph.D.] Introduce numerical optimization <strong>in</strong>to the dynamic clamp protocol toanalyze experimentally <strong>in</strong> real time the (<strong>in</strong>)activation parameters of membranecurrents.12. [Ph.D.] Use new classification of families of channels (Kv3,1, Na v 1.2, etc. seeHille 2001), determ<strong>in</strong>e the k<strong>in</strong>etics of each subgroup, and provide a completetable similar to those <strong>in</strong> Sect. 2.3.5.


54 Electrophysiology of Neurons


Chapter 3One-Dimensional <strong>Systems</strong>In this chapter we describe geometrical methods of analysis of one-dimensional dynamicalsystems, i.e., systems hav<strong>in</strong>g only one variable. An example of such a system isthe space-clamped membrane hav<strong>in</strong>g Ohmic leak current I LC ˙V = −g L (V − E L ) . (3.1)Here the membrane voltage V is a time-dependent variable, and the capacitance C, leakconductance g L and leak reverse potential E L are constant parameters described <strong>in</strong> theprevious chapter. We use this and other one-dimensional neural models to <strong>in</strong>troduceand illustrate the most important concepts of dynamical system theory: equilibrium,stability, attractor, phase portrait, and bifurcation.3.1 Electrophysiological ExamplesThe Hodgk<strong>in</strong>-Huxley description of dynamics of membrane potential and voltage-gatedconductances can be reduced to a one-dimensional system when all transmembraneconductances have fast k<strong>in</strong>etics. For the sake of illustration, let us consider a spaceclampedmembrane hav<strong>in</strong>g leak current and a fast voltage-gated current I fast hav<strong>in</strong>gonly one gat<strong>in</strong>g variable p,Leak I L IC ˙V{ }} { { }} fast{= − g L (V − E L ) − g p (V − E) (3.2)ṗ = (p ∞ (V ) − p)/τ(V ) (3.3)with dimensionless parameters C = 1, g L = 1, and g = 1. Suppose that the gat<strong>in</strong>gk<strong>in</strong>etics (3.3) is much faster than the voltage k<strong>in</strong>etics (3.2), which means that thevoltage-sensitive time constant τ(V ) is very small, i.e. τ(V ) ≪ 1, <strong>in</strong> the entire biophysicalvoltage range. Then, the gat<strong>in</strong>g process may be treated as be<strong>in</strong>g <strong>in</strong>stantaneous,and the asymptotic value p = p ∞ (V ) may be used <strong>in</strong> the voltage equation (3.2) to55


56 One-Dimensional <strong>Systems</strong>Activation Variable10.90.80.70.60.50.40.30.2m(t)Membrane Voltage (mV)20100-10-20-30-40-50-60V(t)τ(V) < 0.5τ(V) < 0.1τ(V) < 0.01Instantaneous0.1-7000 1 2 3 4 5 6 7 8 9 10Time (ms)-800 1 2 3 4 5 6 7 8 9 10Time (ms)Figure 3.1: Solution of the full system (3.2, 3.3) converges to that of the reducedone-dimensional system (3.4) as τ(V ) → 0reduce the two-dimensional system (3.2, 3.3) to a one-dimensional equationC ˙V = −g L (V − E L ) −<strong>in</strong>stantaneous I fast{ }} {g p ∞ (V ) (V − E) . (3.4)This reduction <strong>in</strong>troduces a small error of the order τ(V ) ≪ 1, as one can see <strong>in</strong>Fig. 3.1.S<strong>in</strong>ce the hypothetical current I fast can be either <strong>in</strong>ward (E > E L ) or outward(E < E L ), and the gat<strong>in</strong>g process can be either activation (p is m, as <strong>in</strong> the Hodgk<strong>in</strong>-Huxley model) or <strong>in</strong>activation (p is h), there are four fundamentally different choicesfor I fast (V ), which we summarize <strong>in</strong> Fig. 3.2 and elaborate on below.Gat<strong>in</strong>gactivation<strong>in</strong>activationCurrent<strong>in</strong>ward outwardI Na,pI hI KI KirFigure 3.2: Four fundamental examples of voltagegatedcurrents with one gat<strong>in</strong>g variable. In thisbook we treat “hyperpolarization-activated” currentsI h and I Kir as be<strong>in</strong>g <strong>in</strong>activat<strong>in</strong>g currents,which are turned off (<strong>in</strong>activated via h) by depolarizationand turned on (de<strong>in</strong>activated) by hyperpolarization;see discussion <strong>in</strong> Sect. 2.2.4.3.1.1 I-V relations and dynamicsThe four choices <strong>in</strong> Fig. 3.2 result <strong>in</strong> four simple one-dimensional models of the form(3.4):I Na,p -model , I K -model , I h -model , and I Kir -model .These models might seem too simple to biologists, who can easily understand theirbehavior just by look<strong>in</strong>g at the I-V relations of the currents depicted <strong>in</strong> Fig. 3.3 without


One-Dimensional <strong>Systems</strong> 57currents<strong>in</strong>wardoutwardgat<strong>in</strong>gactivation, m<strong>in</strong>activation, hI Na,pnegativeconductanceI hI(V)I(V)VVI KI KirI(V)I(V)VnegativeconductanceVFigure 3.3: Typical currentvoltage(I-V) relations of the fourcurrents considered <strong>in</strong> this chapter.Shaded boxes correspond tonon-monotonic I-V relations hav<strong>in</strong>ga region of negative conductance(I ′ (V ) < 0) <strong>in</strong> the biophysicallyrelevant voltage range.us<strong>in</strong>g any dynamical systems theory. The models might also appear too simple tomathematicians, who can easily understand their dynamics just by look<strong>in</strong>g at thegraphs of the right-hand side of (3.4) without us<strong>in</strong>g any electrophysiological <strong>in</strong>tuition.In fact, the models provide an <strong>in</strong>valuable learn<strong>in</strong>g tool, s<strong>in</strong>ce they establish a bridgebetween electrophysiology and dynamical systems.In Fig. 3.3 we plot typical steady-state current-voltage (I-V) relations of the fourcurrents considered above. Notice that the I-V curve is non-monotonic for I Na,p andI Kir but monotonic for I K and I h , at least <strong>in</strong> the biophysically relevant voltage range.This subtle difference is an <strong>in</strong>dication of the fundamentally different roles these currentsplay <strong>in</strong> neuron dynamics: The I-V relation <strong>in</strong> the first group has a region of “negativeconductance”, i.e., I ′ (V ) < 0, which creates positive feedback between the voltage andthe gat<strong>in</strong>g variable (Fig. 3.4), and plays an amplify<strong>in</strong>g role <strong>in</strong> neuron dynamics. Werefer to such currents as amplify<strong>in</strong>g currents. In contrast, the currents <strong>in</strong> the secondgroup have negative feedback between voltage and gat<strong>in</strong>g variable, and they often result<strong>in</strong> damped oscillation of the membrane potential, as we show <strong>in</strong> the next chapter. Werefer to such currents as resonant currents. Most neural models <strong>in</strong>volve a comb<strong>in</strong>ationof at least one amplify<strong>in</strong>g and one resonant current, as we discuss <strong>in</strong> Chap. 5. Theway these currents are comb<strong>in</strong>ed determ<strong>in</strong>es whether the neuron is an <strong>in</strong>tegrator or aresonator.3.1.2 Leak + <strong>in</strong>stantaneous I Na,pTo ease our <strong>in</strong>troduction <strong>in</strong>to dynamical systems, we will use the I Na,p -modelC ˙V = I − g L (V − E L ) −<strong>in</strong>stantaneous I Na,p{ }} {g Na m ∞ (V ) (V − E Na ) (3.5)withm ∞ (V ) = 1/(1 + exp {(V 1/2 − V )/k})


58 One-Dimensional <strong>Systems</strong><strong>in</strong>wardcurrentsoutwardgat<strong>in</strong>gactivation, mdepolarization+ +<strong>in</strong>crease+ <strong>in</strong> m<strong>in</strong>wardcurrent+depolarization- +<strong>in</strong>crease- <strong>in</strong> moutwardcurrent+<strong>in</strong>activation, h-hyperpolarization- +<strong>in</strong>crease<strong>in</strong> h<strong>in</strong>wardcurrent+-negative feedback,resonant current +hyperpolarization+ +<strong>in</strong>crease+ <strong>in</strong> houtwardcurrent+positive feedback,amplify<strong>in</strong>g currentFigure 3.4: Feedback loopsbetween voltage and gat<strong>in</strong>gvariables <strong>in</strong> the four modelspresented above; see also Fig. 5.2.throughout the rest of this chapter. (Some biologists refer to transient Na + currentswith very slow <strong>in</strong>activation as be<strong>in</strong>g persistent, s<strong>in</strong>ce the current does not change muchon the time scale of 1 sec.) We obta<strong>in</strong> the experimental parameter valuesC = 10 µF , I = 0 pA , g L = 19 mS , E L = −67 mV ,g Na = 74 mS , V 1/2 = 1.5 mV , k = 16 mV , E Na = 60 mVus<strong>in</strong>g whole-cell patch clamp record<strong>in</strong>gs of a layer 5 pyramidal neuron <strong>in</strong> the visualcortex of a rat at room temperature. We prove <strong>in</strong> Ex. 3.3.8 and illustrate <strong>in</strong> Fig. 3.15that the model approximates the action potential upstroke dynamics of this neuron.The model’s I-V relation, I(V ), is depicted <strong>in</strong> Fig. 3.5a. Due to the negativeconductance region <strong>in</strong> the I-V curve, this one-dimensional model can exhibit a numberof <strong>in</strong>terest<strong>in</strong>g non-l<strong>in</strong>ear phenomena, such as bistability, i.e. co-existence of rest<strong>in</strong>gand excited states. From a mathematical po<strong>in</strong>t of view, bistability occurs becausethe right-hand side function <strong>in</strong> the differential equation (3.5), depicted <strong>in</strong> Fig. 3.5b,is not monotonic. In Fig. 3.6 we depict typical voltage time courses of the model(3.5) with two values of <strong>in</strong>jected dc-current I and 16 different <strong>in</strong>itial conditions. Thequalitative behavior <strong>in</strong> Fig. 3.6a is clearly bistable: depend<strong>in</strong>g on the <strong>in</strong>itial condition,the trajectory of the membrane potential goes either up to the excited state or downto the rest<strong>in</strong>g state. In contrast, the behavior <strong>in</strong> Fig. 3.6b is monostable, s<strong>in</strong>ce therest<strong>in</strong>g state does not exist. The goal of the dynamical system theory reviewed <strong>in</strong> thischapter is to understand why and how the behavior depends on the <strong>in</strong>itial conditionsand the parameters of the system.3.2 <strong>Dynamical</strong> <strong>Systems</strong>In general, dynamical systems can be cont<strong>in</strong>uous or discrete, depend<strong>in</strong>g on whetherthey are described by differential or difference equations. Cont<strong>in</strong>uous one-dimensional


One-Dimensional <strong>Systems</strong> 59current (nA)I1L (V)I(V)0-1I Na,p (V)-2-60 -40 -20 0 20 40 60membrane potential (mV)aderivative of membrane potential (mV/ms)100V=F(V)50F(V)=-I(V)/C0-50-60 -40 -20 0 20 40 60membrane potential, V (mV)bFigure 3.5: a. I-V relations of the leak current, I L , fast Na + current, I Na , and comb<strong>in</strong>edcurrent I(V ) = I L (V ) + I Na (V ) <strong>in</strong> the I Na,p -model (3.5). Dots denote I 0 (V ) data fromlayer 5 pyramidal cell <strong>in</strong> rat visual cortex. b. The right-hand side of the I Na,p -model(3.5).dynamical systems are usually written <strong>in</strong> the form˙V = F (V ) , V (0) = V 0 ∈ R , (3.6)for example,˙V = −80 − V , V (0) = −20 ,where V is a scalar time-dependent variable denot<strong>in</strong>g the current state of the system,˙V = V t = dV/dt is its derivative with respect to time t, F is a scalar function (itsoutput is one-dimensional) that determ<strong>in</strong>es the evolution of the system, e.g., the righthandside of (3.5) divided by C; see Fig. 3.5b. V 0 ∈ R is an <strong>in</strong>itial condition, and R isthe real l<strong>in</strong>e, i.e., a l<strong>in</strong>e of real numbers (R n would be the n-dimensional real space).In the context of dynamical systems, the real l<strong>in</strong>e R is called phase l<strong>in</strong>e or state l<strong>in</strong>e(phase space or state space for R n ) to stress the fact that each po<strong>in</strong>t <strong>in</strong> R correspondsto a certa<strong>in</strong>, possibly <strong>in</strong>admissible state of the system, and each state of the systemcorresponds to a certa<strong>in</strong> po<strong>in</strong>t <strong>in</strong> R. For example, the state of the Ohmic membrane(3.1) is just its membrane potential V ∈ R. The state of the Hodgk<strong>in</strong>-Huxley model(see Sect. 2.3) is the four-dimensional vector (V, m, n, h) ∈ R 4 . The state of the I Na,p -model (3.5) is aga<strong>in</strong> its membrane potential V ∈ R, because the value m = m ∞ (V ) isunequivocally def<strong>in</strong>ed by V .When all parameters are constant, the dynamical system is called autonomous.When at least one of the parameters is time-dependent, the system is non-autonomous,denoted as ˙V = F (V, t).To solve (3.6) means to f<strong>in</strong>d a function V (t) whose <strong>in</strong>itial value is V (0) = V 0and whose derivative is F (V (t)) at each moment t ≥ 0. For example, the functionV (t) = V 0 + at is an explicit analytical solution to the dynamical system ˙V = a. Theexponentially decay<strong>in</strong>g function V (t) = E L + (V 0 − E L )e −g Lt/C depicted <strong>in</strong> Fig. 3.7,


60 One-Dimensional <strong>Systems</strong>40bistability (I=0)excited40monostability (I=60)excitedmembrane potential, V (mV)200-20-40V(t)rest<strong>in</strong>gmembrane potential, V (mV)200-20-40V(t)-600 1 2 3 4 5time (ms)a-600 1 2 3 4 5time (ms)bFigure 3.6: Typical voltage trajectories of the I Na,p -model (3.5) hav<strong>in</strong>g different valuesof I.membrane potential, VV 0V(0)V(h)V(2h)V(3h)V(t)=E L+(V 0-E L)e -g L t/CE Ltime, tFigure 3.7: Explicit analytical solution (V (t) = E L + (V 0 − E L )e −g Lt/C ) of the l<strong>in</strong>earequation (3.1) and correspond<strong>in</strong>g numerical approximation (dots) us<strong>in</strong>g Euler’s method(3.7).solid curve, is an explicit analytical solution to the l<strong>in</strong>ear equation (3.1) (check bydifferentiat<strong>in</strong>g).F<strong>in</strong>d<strong>in</strong>g explicit solutions is often impossible even for such simple systems as (3.5),so quantitative analysis is carried out mostly via numerical simulations. The simplestprocedure to solve (3.6) numerically, known as first-order Euler method, replaces (3.6)by the discretized system[V (t + h) − V (t)]/h = F (V (t))where t = 0, h, 2h, 3h, . . . , is the discrete time and h is a small time step. Know<strong>in</strong>g thecurrent state V (t), we can f<strong>in</strong>d the next state po<strong>in</strong>t viaV (t + h) = V (t) + hF (V (t)) . (3.7)Iterat<strong>in</strong>g this difference equation start<strong>in</strong>g with V (0) = V 0 , we can approximate the


One-Dimensional <strong>Systems</strong> 61membrane potential, V (mV) graph of F(V)=V100500-50I LV>0 V=0 V


62 One-Dimensional <strong>Systems</strong>at every po<strong>in</strong>t V where F (V ) is negative, the derivative ˙V is negative, and hence thestate variable V decreases. In contrast, at every po<strong>in</strong>t where F (V ) is positive, ˙V ispositive, and the state variable V <strong>in</strong>creases; the greater the value of F (V ), the fasterV <strong>in</strong>creases. Thus, the direction of movement of the state variable V , and hence theevolution of the dynamical system, is determ<strong>in</strong>ed by the sign of the function F (V ).The right-hand side of the I leak -model (3.1) or the I Na,p -model (3.5) <strong>in</strong> Fig. 3.8 is thesteady-state current-voltage (I-V) relation, I L (V ) or I L (V )+I Na,p (V ) respectively, takenwith the m<strong>in</strong>us sign, see Fig. 3.5. Positive values of the right-hand side F (V ) meannegative I-V, correspond<strong>in</strong>g to a net <strong>in</strong>ward current that depolarizes the membrane.Conversely, negative values mean positive I-V, correspond<strong>in</strong>g to a net outward currentthat hyperpolarizes the membrane.3.2.2 EquilibriaThe next step <strong>in</strong> the qualitative analysis of any dynamical system is to f<strong>in</strong>d its equilibriaor rest po<strong>in</strong>ts, i.e., the values of the state variable whereF (V ) = 0(V is an equilibrium).At each such po<strong>in</strong>t ˙V = 0, the state variable V does not change. In the context ofmembrane potential dynamics, equilibria correspond to the po<strong>in</strong>ts where the steadystateI-V curve passes zero. At each such po<strong>in</strong>t there is a balance of the <strong>in</strong>ward andoutward currents so that the net transmembrane current is zero, and the membranevoltage does not change. (Incidentally, the part l ībra <strong>in</strong> the Lat<strong>in</strong> word aequil ībriummeans balance).The I K - and I h -models mentioned <strong>in</strong> Sect. 3.1 can have only one equilibrium becausetheir I-V relations I(V ) are monotonic <strong>in</strong>creas<strong>in</strong>g functions. The correspond<strong>in</strong>gfunctions F (V ) are monotonic decreas<strong>in</strong>g and can have only one zero.In contrast, the I Na,p - and I Kir -models can have many equilibria because their I-Vcurves are not monotonic, and hence there is a possibility for multiple <strong>in</strong>tersectionswith the V -axis. For example, there are three equilibria <strong>in</strong> Fig. 3.8b correspond<strong>in</strong>g tothe rest state (around −53 mV), threshold state (around −40 mV) and the excitedstate (around 30 mV). Each equilibrium corresponds to the balance of the outwardleak current and partially (rest), moderately (threshold) or fully (excited) activatedpersistent Na + <strong>in</strong>ward current. Throughout this book we denote equilibria as smallopen or filled circles depend<strong>in</strong>g on their stability, as <strong>in</strong> Fig. 3.8.3.2.3 StabilityIf the <strong>in</strong>itial value of the state variable is exactly at equilibrium, then ˙V = 0 and thevariable will stay there forever. If the <strong>in</strong>itial value is near the equilibrium, the statevariable may approach the equilibrium or diverge from it. Both cases are depicted <strong>in</strong>Fig. 3.8. We say that an equilibrium is asymptotically stable if all solutions start<strong>in</strong>gsufficiently near the equilibrium will approach it as t → ∞.


One-Dimensional <strong>Systems</strong> 63F(V)stableequilibriumunstableequilibriumF(V)VVnegative slopeF (V)0Figure 3.9: The sign of the slope,λ = F ′ (V ), determ<strong>in</strong>es the stabilityof the equilibrium.Stability of an equilibrium is determ<strong>in</strong>ed by the signs of the function F around it.The equilibrium is stable when F (V ) changes the sign from “+” to “−” as V <strong>in</strong>creases,as <strong>in</strong> Fig. 3.8a. Obviously, all solutions start<strong>in</strong>g near such an equilibrium converge toit. Such an equilibrium “attracts” all nearby solutions, and it is called an attractor. Astable equilibrium po<strong>in</strong>t is the only type of attractor that can exist <strong>in</strong> one-dimensionalcont<strong>in</strong>uous dynamical systems def<strong>in</strong>ed on a state l<strong>in</strong>e R. Multidimensional systems canhave other attractors, e.g., limit cycles.The differences between stable, asymptotically stable, and exponentially stableequilibria are discussed <strong>in</strong> Ex. 18 <strong>in</strong> the end of the chapter. The reader is also encouragedto solve Ex. 4 (piece-wise cont<strong>in</strong>uous F (V )).3.2.4 EigenvaluesA sufficient condition for an equilibrium to be stable is that the derivative of thefunction F with respect to V at the equilibrium is negative, provided that the functionis differentiable. We denote this derivative here by the Greek letterλ = F ′ (V ) , (V is an equilibrium; that is, F (V ) = 0)and note that it is just the slope of graph of F at the po<strong>in</strong>t V ; see Fig. 3.9. Obviously,when the slope, λ, is negative, the function changes the sign from “+” to “−”, andthe equilibrium is stable. Positive slope λ implies <strong>in</strong>stability. The parameter λ def<strong>in</strong>edabove is the simplest example of an eigenvalue of an equilibrium. We <strong>in</strong>troduce eigenvaluesformally <strong>in</strong> the next chapter and show that eigenvalues play an important role<strong>in</strong> def<strong>in</strong><strong>in</strong>g the types of equilibria of multi-dimensional systems.3.2.5 Unstable equilibriaIf a one-dimensional system has two stable equilibrium po<strong>in</strong>ts, then they must beseparated by at least one unstable equilibrium po<strong>in</strong>t, as we illustrate <strong>in</strong> Fig. 3.10.(This may not be true <strong>in</strong> multidimensional systems.) Indeed, a cont<strong>in</strong>uous functionF has to change the sign from “−” to “+” somewhere <strong>in</strong> between those equilibria;that is, it has to cross the V axis <strong>in</strong> some po<strong>in</strong>t, as <strong>in</strong> Fig. 3.8b. This po<strong>in</strong>t would be


64 One-Dimensional <strong>Systems</strong>?F(V)>0+ - - ? + + -??F(V)


¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡One-Dimensional <strong>Systems</strong> 65VF(V)¡ ¡ ¡ ¡- F(v)dv¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡Figure 3.11: Mechanistic <strong>in</strong>terpretation of stable and unstable equilibria. A massless(<strong>in</strong>ertia free) ball moves toward energy m<strong>in</strong>ima with the speed proportional to the slope.A one-dimensional system ˙V = F (V ) has the energy landscape E(V ) = − ∫ VF (v) dv;−∞see Ex. 17. Zeros of F (V ) with negative (positive) slope correspond to m<strong>in</strong>ima (maxima)of E(V ).3.2.7 Threshold and action potentialUnstable equilibria play the role of thresholds <strong>in</strong> one-dimensional bistable systems,i.e., <strong>in</strong> systems hav<strong>in</strong>g two attractors. We illustrate this <strong>in</strong> Fig. 3.13, which is believedto describe the essence of the mechanism of bistability <strong>in</strong> many neurons. Supposethe state variable is <strong>in</strong>itially at the stable equilibrium po<strong>in</strong>t marked as “state A” <strong>in</strong>the figure, and suppose that perturbations can kick it around the equilibrium. Smallperturbations may not kick it over the unstable equilibrium so that the state variablecont<strong>in</strong>ues to be <strong>in</strong> the attraction doma<strong>in</strong> of “state A”. We refer to such perturbationsas be<strong>in</strong>g subthreshold. In contrast, we refer to perturbations as be<strong>in</strong>g superthreshold(also known as suprathreshold) if they are large enough to push the state variable overthe unstable equilibrium so that it becomes attracted to the “state B”. We see thatthe unstable equilibrium acts as a threshold that separates two states.The transition between two stable states separated by a threshold is relevant to themechanism of excitability and generation of action potentials <strong>in</strong> many neurons, whichwe illustrate <strong>in</strong> Fig. 3.14. In the I Na,p -model (3.5) with the I-V relation <strong>in</strong> Fig. 3.5the existence of the rest state is largely due to the leak current I L , while the existenceof the excited state is largely due to the persistent <strong>in</strong>ward Na + current I Na,p . Small


66 One-Dimensional <strong>Systems</strong>UnstableEquilibriumVF(V)Attraction Doma<strong>in</strong>Attraction Doma<strong>in</strong>Figure 3.12: Two attraction doma<strong>in</strong>s <strong>in</strong> a one-dimensional system are separated by theunstable equilibrium.(subthreshold) perturbations leave the state variable <strong>in</strong> the attraction doma<strong>in</strong> of therest state, while large (superthreshold) perturbations <strong>in</strong>itiate the regenerative process— the upstroke of an action potential, and the voltage variable becomes attracted to theexcited state. Generation of the action potential must be completed via repolarization,which moves V back to the rest state. Typically, repolarization occurs because of arelatively slow <strong>in</strong>activation of Na + current and/or slow activation of an outward K +current, which are not taken <strong>in</strong>to account <strong>in</strong> the one-dimensional system (3.5). Toaccount for such processes, we consider two-dimensional systems <strong>in</strong> the next chapter.Recall that the parameters of the I Na,p -model (3.5) were obta<strong>in</strong>ed from a corticalpyramidal neuron. In Fig. 3.15, left, we stimulate (<strong>in</strong> vitro) the cortical neuron by short(0.1 ms) strong pulses of current to reset its membrane potential to various <strong>in</strong>itial valuesand <strong>in</strong>terpret the results us<strong>in</strong>g the I Na,p -model. S<strong>in</strong>ce activation of Na + current is not<strong>in</strong>stantaneous <strong>in</strong> real neurons, we allow variable m to converge to m ∞ (V ), and ignorethe 0.3-ms transient activity that follows each pulse. We also ignore the <strong>in</strong>itial segmentof the downstroke of the action potential, and plot the magnification of the voltagetraces <strong>in</strong> Fig. 3.15, right. Compar<strong>in</strong>g this figure with Fig. 3.8b, we see that the I Na,p -model is a reasonable one-dimensional approximation of the action potential upstrokedynamics; It predicts the value of the rest<strong>in</strong>g (−53 mV), <strong>in</strong>stantaneous threshold (−40mV), and the excited (+30 mV) states of the cortical neuron.3.2.8 Bistability and hysteresis<strong>Systems</strong> hav<strong>in</strong>g two (many) co-exist<strong>in</strong>g attractors are called bistable (multi-stable).Many neurons and neuronal models, such as the Hodgk<strong>in</strong>-Huxley model, exhibit bistabilitybetween rest<strong>in</strong>g (equilibrium) and spik<strong>in</strong>g (limit cycle) attractors. Some neuronscan exhibit bistability of two stable rest<strong>in</strong>g states <strong>in</strong> the subthreshold voltage range,e.g., −59 mV and −75 mV <strong>in</strong> the thalamocortical neurons (Hughes et al. 1999) depicted


¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡One-Dimensional <strong>Systems</strong> 67Vstate Athresholdstate BAttraction doma<strong>in</strong>of state AAttraction doma<strong>in</strong>of state BF(V)subthresholdperturbation¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡state Athresholdstate B¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡superthresholdperturbation¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡state Athresholdstate B¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡Figure 3.13: Unstable equilibrium plays the role of a threshold that separates twoattraction doma<strong>in</strong>s.


¡ ¡ ¡¡ ¡ ¡¡ ¡ ¡¡ ¡ ¡¡ ¡ ¡¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡68 One-Dimensional <strong>Systems</strong>repolarization(another variable)superthresholdperturbation¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡I Khyperpolarization¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡regenerativeupstrokedepolarization¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡rest statethresholdI NaexcitedFigure 3.14: Mechanistic illustration of the mechanism of generation of an action potential.strongdepolarizationmembrane potential (mV)-40 mV20 mV1 ms+30 mVpulses of currentactionpotentialsmembrane potential (mV)4030stable equilibrium20100-10-20-30unstable equilibrium-40-50stable equilibrium-600 0.2 0.4 0.6 0.8 1 1.2time (ms)Figure 3.15: Upstroke dynamics of layer 5 pyramidal neuron <strong>in</strong> vitro (compare withthe I Na,p -model (3.5) <strong>in</strong> Fig. 3.8b).


One-Dimensional <strong>Systems</strong> 6915 mV-59 mV-75 mVmembranepotential100 pA2 sFigure 3.16: Membrane potential bistability <strong>in</strong> a cat TC neuron <strong>in</strong> the presence ofZD7288 (pharmacological blocker of I h ; modified from Fig. 6B of Hughes et al. 1999).(a)(d)(a)(b)(b)(b)(c)I+F(V)VIV(c)Figure 3.17: Bistability and hysteresis loop as I changes.<strong>in</strong> Fig. 3.16, or −50 mV and −60 mV <strong>in</strong> mitral cells of the olfactory bulb (Heyward etal. 2001), or −45 mV and −60 mV <strong>in</strong> Purk<strong>in</strong>je neurons. Brief <strong>in</strong>puts can switch suchneurons from one state to the other, as <strong>in</strong> Fig. 3.16. Though the ionic mechanisms ofbistability are different <strong>in</strong> the three neurons, the mathematical mechanism is the same.Consider a one-dimensional system ˙V = I+F (V ) with function F (V ) hav<strong>in</strong>g a cubicN-shape. Injection of a dc-current I shifts the function I + F (V ) up or down. When Iis negative, the system has only one equilibrium, depicted <strong>in</strong> Fig. 3.17a. As we removethe <strong>in</strong>jected current I, the system becomes bistable, as <strong>in</strong> Fig. 3.17b, but its state isstill at the left equilibrium. As we <strong>in</strong>ject positive current, the left stable equilibriumdisappears via another saddle-node bifurcation, and the state of the system jumps tothe right equilibrium, as <strong>in</strong> Fig. 3.17c. But as we slowly remove the <strong>in</strong>jected currentthat caused the jump and go back to Fig. 3.17b, the jump to the left equilibrium doesnot occur until a much lower value correspond<strong>in</strong>g to Fig. 3.17a is reached. The failureof the system to return to the orig<strong>in</strong>al value when the <strong>in</strong>jected current is removed iscalled hysteresis. If I were a slow V -depended variable, then the system could exhibitrelaxation oscillations depicted <strong>in</strong> Fig. 3.17d and described <strong>in</strong> the next chapter.


70 One-Dimensional <strong>Systems</strong>+ - - + + - - + + - - + + -VFunction F(V)Attraction doma<strong>in</strong>sVPhase PortraitFigure 3.18: Phase portrait of a one-dimensional system ˙V = F (V ).F (V)1VF (V)2VFigure 3.19: Two “seem<strong>in</strong>gly different”dynamical systems ˙V = F1 (V ) and˙V = F 2 (V ) are topologically equivalent,hence they have qualitativelysimilar dynamics.3.3 Phase PortraitsAn important component <strong>in</strong> the qualitative analysis of any dynamical system is reconstructionof its phase portrait. For this one depicts all stable and unstable equilibria(as black and white circles respectively), representative trajectories, and correspond<strong>in</strong>gattraction doma<strong>in</strong>s <strong>in</strong> the system’s state/phase space, as we illustrate <strong>in</strong> Fig. 3.18. Thephase portrait is a geometrical representation of system dynamics. It depicts all possibleevolutions of the state variable and how they depend on the <strong>in</strong>itial state. Look<strong>in</strong>g atthe phase portrait, one immediately gets all important <strong>in</strong>formation about the system’squalitative behavior without even know<strong>in</strong>g the equation for F .


One-Dimensional <strong>Systems</strong> 71F (V)1? F (V)2VVFigure 3.20: Two “seem<strong>in</strong>gly alike”dynamical systems ˙V = F1 (V ) and˙V = F 2 (V ) are not topologicallyequivalent, hence they do not havequalitatively similar dynamics. (Thefirst system has three equilibria, whilethe second system has only one.)3.3.1 Topological equivalencePhase portraits can be used to determ<strong>in</strong>e qualitative similarity of dynamical systems.In particular, two one-dimensional systems are said to be topologically equivalent whenthe phase portrait of one of them treated as a piece of rubber can be stretched orshrunk to fit the other one, as <strong>in</strong> Fig. 3.19. Topological equivalence is a mathematicalconcept that clarifies the imprecise notion of “qualitative similarity”, and its rigorousdef<strong>in</strong>ition is provided, e.g., by Guckenheimer and Holmes (1983).The stretch<strong>in</strong>g and shr<strong>in</strong>k<strong>in</strong>g of the “rubber” phase space are topological transformationsthat do not change the number of equilibria or their stability. Thus, twosystems hav<strong>in</strong>g different number of equilibria cannot be topologically equivalent, hencethey have qualitatively different dynamics, as we illustrate <strong>in</strong> Fig. 3.20. Indeed, the topsystem is bistable because it has two stable equilibria separated by an unstable one.The evolution of the state variable depends on which attraction doma<strong>in</strong> the <strong>in</strong>itial conditionis <strong>in</strong> <strong>in</strong>itially. Such a system has “memory” of the <strong>in</strong>itial condition. Moreover,sufficiently strong perturbations can switch it from one equilibrium state to another.In contrast, the bottom system <strong>in</strong> Fig. 3.20 has only one equilibrium, which is a globalattractor, and the state variable converges to it regardless of the <strong>in</strong>itial condition. Sucha system has quite primitive dynamics, and it is topologically equivalent to the l<strong>in</strong>earsystem (3.1).3.3.2 Local equivalence and the Hartman-Grobman theoremIn computational neuroscience, we usually face quite complicated systems describ<strong>in</strong>gneuronal dynamics. A useful strategy is to substitute such systems by simpler oneshav<strong>in</strong>g topologically equivalent phase portraits. For example, both systems <strong>in</strong> Fig. 3.19are topologically equivalent to ˙V = V − V 3 (please, check this), which is easier to dealwith analytically.Quite often we cannot f<strong>in</strong>d a simpler system that is topologically equivalent to ourneuronal model on the entire state l<strong>in</strong>e R. In this case, we make a sacrifice: we restrict


72 One-Dimensional <strong>Systems</strong>λ(V-Veq)VeqF(V)VFigure 3.21: Hartman-Grobman theorem:The non-l<strong>in</strong>ear system ˙V = F (V )is topologically equivalent to the l<strong>in</strong>earone ˙V = λ(V −V eq ) <strong>in</strong> the local (shaded)neighborhood of the hyperbolic equilibriumV eq .our analysis to a small neighborhood of the l<strong>in</strong>e R, e.g., a neighborhood of the rest<strong>in</strong>gstate or of the threshold, and study behavior locally <strong>in</strong> this neighborhood.An important tool <strong>in</strong> the local analysis of dynamical systems is the Hartman-Grobman theorem, which says that a non-l<strong>in</strong>ear one-dimensional system˙V = F (V )sufficiently near an equilibrium V = V eq is locally topologically equivalent to the l<strong>in</strong>earone˙V = λ(V − V eq ) (3.8)provided that the eigenvalueλ = F ′ (V eq )at the equilibrium is non-zero, i.e., the slope of F (V ) is non-zero. Such an equilibriumis called hyperbolic. Thus, nonl<strong>in</strong>ear systems near hyperbolic equilibria behave as ifthere were l<strong>in</strong>ear, as <strong>in</strong> Fig. 3.21.It is easy to f<strong>in</strong>d the exact solution of the l<strong>in</strong>earized system (3.8) with an <strong>in</strong>itialcondition V (0) = V 0 . It is V (t) = V eq + e λt (V 0 − V eq ) (check by differentiat<strong>in</strong>g).If the eigenvalue λ < 0, then e λt → 0 and V (t) → V eq as t → ∞, so that theequilibrium is stable. Conversely, if λ > 0, then e λt → ∞ mean<strong>in</strong>g that the <strong>in</strong>itialdisplacement, V 0 − V eq , grows with the time, and the equilibrium is unstable. Thus,the l<strong>in</strong>earization predicts qualitative dynamics at the equilibrium and quantitative rateof convergence/divergence to/from the equilibrium.If the eigenvalue λ = 0, then the equilibrium is non-hyperbolic, and analysis ofthe l<strong>in</strong>earized system ˙V = 0 cannot describe the behavior of the nonl<strong>in</strong>ear system.Typically, non-hyperbolic equilibria arise when the system undergoes a bifurcation,i.e., a qualitative change of behavior, which we consider next. To study stability, weneed to consider higher-order terms of the Taylor series of F (V ) at V eq .3.3.3 BifurcationsThe f<strong>in</strong>al and the most advanced step <strong>in</strong> the qualitative analysis of any dynamicalsystem is the bifurcation analysis. In general, a system is said to undergo a bifurcationwhen its phase portrait changes qualitatively. For example, the energy landscape <strong>in</strong>Fig. 3.22 changes so that the system is no longer bistable. The precise mathematicaldef<strong>in</strong>ition of a bifurcation will be given later.


¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡One-Dimensional <strong>Systems</strong> 73Bistability¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡ ¡ ¡ ¡Bifurcation¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡¡ ¡ ¡ ¡ ¡ ¡Monostability¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡ ¡Figure 3.22: Mechanistic illustration of a bifurcation as a change of the landscape.Qualitative change of the phase portrait may or may not necessarily reveal itself<strong>in</strong> a qualitative change of behavior, depend<strong>in</strong>g on the <strong>in</strong>itial conditions. For example,there is a bifurcation <strong>in</strong> Fig. 3.23, left, but no change of behavior because the ballrema<strong>in</strong>s <strong>in</strong> the attraction doma<strong>in</strong> of the right equilibrium. To see the change, we needto drop the ball at different <strong>in</strong>itial conditions and observe the disappearance of the leftequilibrium. In the same va<strong>in</strong>, there is no bifurcation Fig. 3.23, middle and right, (thephase portraits <strong>in</strong> each column are topologically equivalent) but the apparent change ofbehavior is caused by the expansion of the attraction doma<strong>in</strong> of the left equilibrium orby the external <strong>in</strong>put. Dropp<strong>in</strong>g the ball at different locations would result <strong>in</strong> the samequalitative picture — two stable equilibria whose attraction doma<strong>in</strong>s are separated bythe unstable equilibrium. When mathematicians talk about bifurcations, they assumethat all <strong>in</strong>itial conditions could be sampled, <strong>in</strong> which case bifurcations do result <strong>in</strong> aqualitative change of behavior of the system as a whole.To illustrate the importance of sampl<strong>in</strong>g all <strong>in</strong>itial conditions, let us consider the <strong>in</strong>vitro record<strong>in</strong>gs of a pyramidal neuron <strong>in</strong> Fig. 3.24. We <strong>in</strong>ject 0.1-ms strong pulses ofcurrent of various amplitude to set the membrane potential to different <strong>in</strong>itial values.Right after each pulse, we <strong>in</strong>ject a 4 ms step of dc-current of amplitude I = 0, I = 16or I = 60 pA. The case I = 0 pA is the same as <strong>in</strong> Fig. 3.15, so some <strong>in</strong>itial conditions


74 One-Dimensional <strong>Systems</strong>bifurcationbut no change of behaviorchange of behavior but no bifurcationpulsed<strong>in</strong>putFigure 3.23: Bifurcations are not equivalent to qualitative change of behavior if thesystem is started with the same <strong>in</strong>itial condition or subject to external <strong>in</strong>put.bistablebistablemonostablemembrane potential, mV10 mV1 ms-40 mVdc-current I=0 dc-current I=16dc-current I=60pulses pulses pulsesFigure 3.24: Qualitative change of the up-stroke dynamics of a layer 5 pyramidal neuronfrom rat visual cortex (the same neuron as <strong>in</strong> Fig. 3.15).


One-Dimensional <strong>Systems</strong> 75result <strong>in</strong> upstroke of the action potential, while others do not. When I = 60 pA, all<strong>in</strong>itial conditions result <strong>in</strong> the generation of an action potential. Clearly, a change ofqualitative behavior occurs for some I between 0 and 60.To understand the qualitative dynamics <strong>in</strong> Fig. 3.24, we consider the one-dimensionalI Na,p -model (3.5) hav<strong>in</strong>g different values of the parameter I and depict its trajectories<strong>in</strong> Fig. 3.25. One can clearly see that the qualitative behavior of the model dependson whether I is greater or less than 16. When I = 0 (top of Fig. 3.25), the system isbistable. The rest and the excited states coexist. When I is large (bottom of Fig. 3.25)the rest state no longer exists because the leak outward current cannot balance thelarge <strong>in</strong>jected dc-current I and the <strong>in</strong>ward Na + current.What happens when we change I past 16? The answer lies <strong>in</strong> the details of thegeometry of the right-hand side function F (V ) of (3.5) and how it depends on theparameter I. Increas<strong>in</strong>g I elevates the graph of F (V ). The higher the graph of F (V )is, the closer its <strong>in</strong>tersections with the V -axis are, as we illustrate <strong>in</strong> Fig. 3.26 depict<strong>in</strong>gonly the low-voltage range of the system. When I approaches 16, the distance betweenthe stable and unstable equilibria vanishes; the equilibria coalesce and annihilate eachother. The value I = 16 at which the equilibria coalesce is called the bifurcation value.This value separates two qualitatively different regimes: When I is near but less than16, the system has three equilibria and bistable dynamics. The quantitative features,such as the exact locations of the equilibria, depend on the particular values of I, butthe qualitative behavior rema<strong>in</strong>s unchanged no matter how close I to the bifurcationvalue is. In contrast, when I is near but greater than 16 the system has only oneequilibrium and monostable dynamics.In general, a dynamical system may depend on a vector of parameters, say p. Apo<strong>in</strong>t <strong>in</strong> the parameter space, say p = a, is said to be a regular or non-bifurcation po<strong>in</strong>t,if the system’s phase portrait at p = a is topologically equivalent to the phase portraitat p = c for any c sufficiently close to a. For example, the value I = 13 <strong>in</strong> Fig. 3.26 isregular, s<strong>in</strong>ce the system has topologically equivalent phase portraits for all I near 13.Similarly, the value I = 18 is also regular. Any po<strong>in</strong>t <strong>in</strong> the parameter space that isnot regular is called a bifurcation po<strong>in</strong>t. Namely, a po<strong>in</strong>t p = b is a bifurcation po<strong>in</strong>t,if the system’s phase portrait at p = b is not topologically equivalent to the phaseportrait at a po<strong>in</strong>t p = c no matter how close c to b is. The value I = 16 <strong>in</strong> Fig. 3.26 isa bifurcation po<strong>in</strong>t. It corresponds to the saddle-node (also known as fold or tangent)bifurcation for reasons described later. It is one of the simplest bifurcations considered<strong>in</strong> this book.3.3.4 Saddle-node (fold) bifurcationIn general, a one-dimensional system˙V = F (V, I)hav<strong>in</strong>g an equilibrium po<strong>in</strong>t V = V sn for some value of the parameter I = I sn (i.e.,F (V sn , I sn ) = 0) is said to be at a saddle-node bifurcation (sometimes called a fold


76 One-Dimensional <strong>Systems</strong>100806040200100806040200100806040200bistabilityF(V)rest thresholdbifurcationF(V)tangent po<strong>in</strong>tmonostabilityF(V)-50 0 50membrane potential, V (mV)I=0I=16I=60membrane potential, V (mV) membrane potential, V (mV) membrane potential, V (mV)40200-20-40-6040200-20-40-6040200-20-40V(t)excited statethresholdrestV(t)excited stateV(t)excited state-600 1 2 3 4 5time (ms)Figure 3.25: Bifurcation <strong>in</strong> the I Na,p -model (3.5): The rest state and the thresholdstate coalesce and disappear when the parameter I <strong>in</strong>creases.


One-Dimensional <strong>Systems</strong> 77F(V)I=18I=17I=16I=15I=14I=13I=12I=11no equilibriabifurcationtwo equilibriatangentpo<strong>in</strong>tVstable equilibriaunstable equilibriaFigure 3.26: Saddle-node bifurcation: As the graph of the function F (V ) is lifted up,the stable and unstable equilibria approach each other, coalesce at the tangent po<strong>in</strong>t,and then disappear.saddle-nodenot saddle-nodeFVnon-hyperbolichyperbolichyperbolicnon-degenerate degenerate degeneratetransversalnot transversalnot transversalFigure 3.27: Geometrical illustration of the three conditions def<strong>in</strong><strong>in</strong>g saddle-node bifurcations.Arrows denote the direction of displacement of the function F (V, I) as thebifurcation parameter I changes.


78 One-Dimensional <strong>Systems</strong>bifurcation) if the follow<strong>in</strong>g mathematical conditions, illustrated <strong>in</strong> Fig. 3.27, are satisfied:• (Non-hyperbolicity) The eigenvalue λ at V sn is zero; that is,λ = F V (V, I sn ) = 0 (at V = V sn ),where F V denotes the derivative of F with respect to V , that is, F V = ∂F/∂V .Equilibria with zero or pure imag<strong>in</strong>ary eigenvalues are called non-hyperbolic.Geometrically, this condition implies that the graph of F has horizontal slope atthe equilibrium.• (Non-degeneracy) The second order derivative with respect to V at V sn is nonzero;that is,F V V (V, I sn ) ≠ 0 (at V = V sn ).Geometrically, this means that the graph of F looks like the square parabola V 2<strong>in</strong> Fig. 3.27.• (Transversality) The function F (V, I) is non-degenerate with respect to the bifurcationparameter I; that is,F I (V sn , I) ≠ 0 (at I = I sn ),where F I denotes the derivative of F with respect to I. Geometrically, thismeans that as I changes past I sn , the graph of F approaches, touches, and then<strong>in</strong>tersects the V axis.Saddle-node bifurcation results <strong>in</strong> appearance or disappearance of a pair of equilibria,as <strong>in</strong> Fig. 3.26. None of the six examples on the right-hand side of Fig. 3.27 can undergoa saddle-node bifurcation because at least one of the conditions above is violated.The number of conditions <strong>in</strong>volv<strong>in</strong>g strict equality (“=”) is called the co-dimensionof a bifurcation. The saddle-node bifurcation has co-dimension-1 because there is onlyone condition <strong>in</strong>volv<strong>in</strong>g “=”, and the other two conditions <strong>in</strong>volve <strong>in</strong>equalities (“≠”).Co-dimension-1 bifurcations can be reliably observed <strong>in</strong> systems with one parameter.It is an easy exercise to check that the one-dimensional system˙V = I + V 2 (3.9)is at saddle-node bifurcation when V = 0 and I = 0 (please, check all three conditions).This system is called the topological normal form for saddle-node bifurcation. Thephase portraits of this system are topologically equivalent to those depicted <strong>in</strong> Fig. 3.26,except that the bifurcation occurs at I = 0, and not at I = 16.


One-Dimensional <strong>Systems</strong> 79F(V)10 mV0.5 msExcited0Attractorru<strong>in</strong>smembrane potential (mV)ExcitedV-40 mVslow transitionAttractorru<strong>in</strong>sFigure 3.28: Slow transition through the ghost of the rest<strong>in</strong>g state attractor <strong>in</strong> a corticalpyramidal neuron with I = 30 pA (the same neuron as <strong>in</strong> Fig. 3.15). Even thoughthe rest<strong>in</strong>g state has already disappeared, the function F (V ), and hence the rate ofchange, ˙V , is still small when V ≈ −46 mV.3.3.5 Slow transitionAll physical, chemical, and biological systems near saddle-node bifurcations possesscerta<strong>in</strong> universal features that do not depend on particulars of the systems. Consequently,all neural systems near such a bifurcation share common neuro-computationalproperties, which we will discuss <strong>in</strong> detail <strong>in</strong> Chapter 7. Here we take a look at onesuch property — slow transition through the ru<strong>in</strong>s (or ghost) of the rest state attractor,which is relevant to the dynamics of many neocortical neurons.In Fig. 3.28 we show the function F (V ) of the system (3.5) with I = 30 pA,which is greater than the bifurcation value 16 pA, and the correspond<strong>in</strong>g behaviorof a cortical neuron; compare with Fig. 3.15. The system has only one attractor— the excited state, and any solution start<strong>in</strong>g from an arbitrary <strong>in</strong>itial conditionshould quickly approach this attractor. However, the solutions start<strong>in</strong>g from the <strong>in</strong>itialconditions around −50 mV do not seem to hurry. Instead, they slow down near −46mV and spend a considerable amount of time <strong>in</strong> the voltage range correspond<strong>in</strong>g to therest<strong>in</strong>g state, as if the state were still present. The closer is I to the bifurcation value,the more time the membrane potential spends <strong>in</strong> the neighborhood of the rest<strong>in</strong>g state.Obviously, such a slow transition cannot be expla<strong>in</strong>ed by a slow activation of the <strong>in</strong>wardNa + current, s<strong>in</strong>ce Na + activation <strong>in</strong> a cortical neuron is practically <strong>in</strong>stantaneous.The slow transition occurs because the neuron or the system (3.5) <strong>in</strong> Fig. 3.28 isnear a saddle-node bifurcation. Even though I is greater than the bifurcation value,and the rest state attractor is already annihilated, the function F (V ) is barely abovethe V -axis at the “annihilation site”. In other words, the rest state attractor hasalready been ru<strong>in</strong>ed, but its “ru<strong>in</strong>s” (or its “ghost”) can still be felt because˙V = F (V ) ≈ 0(at attractor ru<strong>in</strong>s, V ≈ −46 mV),as one can see <strong>in</strong> Fig. 3.28. In Chapter 7 we will show how this property expla<strong>in</strong>s the


80 One-Dimensional <strong>Systems</strong>20 mV100 ms-60 mVslow transition0 pA43.1 pAFigure 3.29: A 400-ms latency <strong>in</strong> a layer 5 pyramidal neuron of rat visual cortex.-40unstable equilibriamembrane potential, V (mV)-45-50stable equilibriasaddle-node(fold) bifurcation-550 5 10 15 20 25<strong>in</strong>jected dc-current I, (pA)Figure 3.30: Bifurcation diagram of the system <strong>in</strong> Fig. 3.26.ability of many neocortical neurons, such as the one <strong>in</strong> Fig. 3.29, to generate repetitiveaction potentials with small frequency, and how it predicts that all such neurons,considered as dynamical systems, reside near saddle-node bifurcations.3.3.6 Bifurcation diagramThe f<strong>in</strong>al step <strong>in</strong> the geometrical bifurcation analysis of one-dimensional systems isthe analysis of bifurcation diagrams, which we do <strong>in</strong> Fig. 3.30 for the saddle-nodebifurcation shown <strong>in</strong> Fig. 3.26. To draw the bifurcation diagram, we determ<strong>in</strong>e thelocations of the stable and unstable equilibria for each value of the parameter I andplot them as white or black circles <strong>in</strong> the (I, V ) plane <strong>in</strong> Fig. 3.30. The equilibria formtwo branches that jo<strong>in</strong> at the fold po<strong>in</strong>t correspond<strong>in</strong>g to the saddle-node bifurcation(hence the alternative name of fold bifurcation). The branch correspond<strong>in</strong>g to theunstable equilibria is dashed to stress its <strong>in</strong>stability. As the bifurcation parameter Ivaries from left to right pass<strong>in</strong>g through the bifurcation po<strong>in</strong>t, the stable and unstable


One-Dimensional <strong>Systems</strong> 81200I-V relationsteady-state current (pA)0-200-400I=16I=-100I (V)-600-800-100 -80 -60 -40 -20 0 20 40membrane potential, V (mV)Figure 3.31: Equilibria are <strong>in</strong>tersectionsof the steady-state I-V curveI ∞ (V ) and a horizontal l<strong>in</strong>e I = const.equilibria coalesce and annihilate each other. As the parameter varies from right toleft, two equilibria — one stable and one unstable — appear from a s<strong>in</strong>gle po<strong>in</strong>t. Thus,depend<strong>in</strong>g on the direction of movement of the bifurcation parameter, the saddle-nodebifurcation expla<strong>in</strong>s disappearance or appearance of a new stable state. In any case,the qualitative behavior of the systems changes exactly at the bifurcation po<strong>in</strong>t.3.3.7 Bifurcations and I-V record<strong>in</strong>gsIn general, determ<strong>in</strong><strong>in</strong>g saddle-node bifurcation diagrams of neurons may be a daunt<strong>in</strong>gmathematical task. However, it is a trivial exercise when the bifurcation parameteris the <strong>in</strong>jected dc-current I. In this case, the bifurcation diagram, such as the one<strong>in</strong> Fig. 3.30, is just the steady-state I-V relation I ∞ (V ) plotted on the (I, V )-plane.Indeed, the equationC ˙V = I − I ∞ (V ) = 0states that V is an equilibrium if and only if the net membrane current, I − I ∞ (V ), iszero. For example, equilibria of the I Na,p -model are solutions of the equation0 = I −I ∞ (V ){ }} {(g L (V − E L ) + g Na m ∞ (V )(V − E Na )) ,which follows directly from (3.5). In Fig. 3.31 we illustrate how to f<strong>in</strong>d the equilibriageometrically: We plot the steady-state I-V curve I ∞ (V ) and draw a horizontal l<strong>in</strong>ewith altitude I. Any <strong>in</strong>tersection satisfies the equation I = I ∞ (V ), and hence is anequilibrium (stable or unstable). Obviously, when I <strong>in</strong>creases past 16, the saddle-nodebifurcation occurs.Notice that the equilibria are po<strong>in</strong>ts on the curve I ∞ (V ), so flipp<strong>in</strong>g and rotat<strong>in</strong>g thecurve by 90 ◦ , as we do <strong>in</strong> Fig. 3.32, left, results <strong>in</strong> a complete saddle-node bifurcationdiagram. The diagram conveys <strong>in</strong> a very condensed manner all important <strong>in</strong>formationabout the qualitative behavior of the I Na,p -model. The three branches of the S-shapedcurve, which is the 90 ◦ -rotated and flipped copy of the N-shaped I-V curve, correspondto the rest, threshold, and excited states of the model. Each slice I = const represents


82 One-Dimensional <strong>Systems</strong>membrane potential, V (mV)40200-20-40-60-80-100-120excited statesthreshold statessaddle-node(fold) bifurcationrest statessaddle-node(fold) bifurcationI (V)16-890membrane potential, V (mV)40200-20-40-60-80-100-120-1000 -500 0<strong>in</strong>jected dc-current, I (pA)-1000 -500 0<strong>in</strong>jected dc-current, I (pA)Figure 3.32: Bifurcation diagram of the I Na,p -model (3.5).the phase portrait of the system, as we illustrate <strong>in</strong> Fig. 3.32, right. Each po<strong>in</strong>t wherethe branches fold (max or m<strong>in</strong> of I ∞ (V )) corresponds to a saddle-node bifurcation.S<strong>in</strong>ce there are two such folds, at I = 16 pA and at I = −890 pA, there are two saddlenodebifurcations <strong>in</strong> the system. The first one, studied <strong>in</strong> Fig. 3.25, corresponds to thedisappearance of the rest state. The other one, illustrated <strong>in</strong> Fig. 3.33, correspondsto the disappearance of the excited state. It occurs because I becomes so negativethat the Na + <strong>in</strong>ward current is no longer strong enough to balance the leak outwardcurrent and the negative <strong>in</strong>jected dc-current to keep the membrane <strong>in</strong> the depolarized(excited) state.Below the reader can f<strong>in</strong>d more examples of bifurcation analysis of the I Na,p - andI Kir -models, which have non-monotonic I-V relations and can exhibit multi-stabilityof states. The I K - and I h -models have monotonic I-V relations and hence only oneequilibrium state. These models cannot have saddle-node bifurcations, as the readeris asked to prove <strong>in</strong> Ex. 13 and 14.3.3.8 Quadratic <strong>in</strong>tegrate-and-fire neuronLet us consider the topological normal form for the saddle-node bifurcation (3.9). From0 = I + V 2 we f<strong>in</strong>d that there are two equilibria, V rest = − √ |I| and V thresh = + √ |I|when I < 0. The equilibria approach and annihilate each other via saddle-node bifurcationwhen I = 0, so there are no equilibria when I > 0. In this case, ˙V ≥ I and V (t)<strong>in</strong>creases to <strong>in</strong>f<strong>in</strong>ity. Because of the quadratic term, the rate of <strong>in</strong>crease also <strong>in</strong>creases,result<strong>in</strong>g <strong>in</strong> a positive feedback loop correspond<strong>in</strong>g to the regenerative activation ofNa + current. In Ex. 15 we show that V (t) escapes to <strong>in</strong>f<strong>in</strong>ity <strong>in</strong> a f<strong>in</strong>ite time, whichcorresponds to the up-stroke of the action potential. The same up-stroke is generatedwhen I < 0, if the voltage variable is pushed beyond the threshold value V thresh .


One-Dimensional <strong>Systems</strong> 8300BistabilityRestExcitedThresholdF(V)Rest BifurcationTangent Po<strong>in</strong>tF(V)I=-400I=-890membrane potential, V membrane potential, VV(t)V(t)ExcitedThresholdRestRest0RestF(V)MonostabilityI=-1000membrane potential, VV(t)Restmembrane potential, VtimeFigure 3.33: Bifurcation <strong>in</strong> the I Na,p -model (3.5): The excited state and the thresholdstate coalesce and disappear when the parameter I is sufficiently small.


84 One-Dimensional <strong>Systems</strong>steady-state current (pA)200-20-40-60I (V)(V sn , I sn )I sn -k(V-V sn )2I-V relation-60 -40 -20membrane potential, V (mV)Figure 3.34: Magnification of the I-V curve <strong>in</strong> Fig. 3.31 at the left knee shows that itcan be approximated by a square parabola.Consider<strong>in</strong>g <strong>in</strong>f<strong>in</strong>ite values of the membrane potential may be convenient from apurely mathematical po<strong>in</strong>t of view, but this has no physical mean<strong>in</strong>g and no way tosimulate it on a digital computer. Instead, we fix a sufficiently large constant V peakand say that (3.9) generated a spike when V (t) reached V peak . After the peak of thespike is reached, we reset V (t) to a new value V reset . The topological normal form forthe saddle-node bifurcation with the after-spike resett<strong>in</strong>g˙V = I + V 2 , if V ≥ V peak , then V ← V reset (3.10)is called the quadratic <strong>in</strong>tegrate-and-fire neuron. It is the simplest model of a spik<strong>in</strong>gneuron. The name stems from its resemblance to the leaky <strong>in</strong>tegrate-and-fire neuron˙V = I −V considered <strong>in</strong> Chap. 8. In contrast to the common folklore, the leaky neuronis not a spik<strong>in</strong>g model because it does not have a spike-generation mechanism, i.e., aregenerative up-stroke of the membrane potential, whereas the quadratic neuron does.We discuss this and other issues <strong>in</strong> detail <strong>in</strong> Chap. 8.In general, quadratic <strong>in</strong>tegrate-and-fire model could be derived directly from theequation C ˙V = I − I ∞ (V ) by approximat<strong>in</strong>g the steady-state I-V curve near therest<strong>in</strong>g state by the square parabola I ∞ (V ) ≈ I sn − k(V − V sn ) 2 , where k > 0 andthe peak of the curve (V sn , I sn ) could be easily found experimentally; see Fig. 3.34.Approximat<strong>in</strong>g the I-V curve by other functions, for example I ∞ (V ) = g leak (V −V rest )−ke pV , results <strong>in</strong> other forms of the model, e.g., the exponential <strong>in</strong>tegrate-and-fire model(Fourcaud-Trocme et al. 2003), which has certa<strong>in</strong> advantages over the quadratic form.Unfortunately, the model is not solvable analytically, and it is expensive to simulate.The form I ∞ (V ) = g leak (V −V leak )−k(V −V th ) 2 +, where x + = x when x > 0 and x + = 0otherwise, comb<strong>in</strong>es the advantages of both models. The parameters V peak and V resetare derived from the shape of the spike. Normalization of variables and parametersresults <strong>in</strong> the form (3.10) with V peak = 1.In Fig. 3.35 we simulated the quadratic <strong>in</strong>tegrate-and-fire neuron to illustrate a


One-Dimensional <strong>Systems</strong> 85000I 0, not shown <strong>in</strong> the figure, the membrane potential is reset to the right ofthe ghost, no slow transition is <strong>in</strong>volved, and the tonic spik<strong>in</strong>g starts with a non-zerofrequency. As an exercise, expla<strong>in</strong> why there is a noticeable latency (delay) to the first


86 One-Dimensional <strong>Systems</strong>spike right after the bifurcation. This type of behavior is typical <strong>in</strong> sp<strong>in</strong>y projectionneurons of neostriatum and basal ganglia, as we show <strong>in</strong> Chap. 8.Review of Important Concepts• The one-dimensional dynamical system ˙V = F (V ) describes howthe rate of change of V depends on V . Positive F (V ) means V<strong>in</strong>creases, negative F (V ) means V decreases.• In the context of neuronal dynamics, V is often the membrane potential,and F (V ) is the steady-state I-V curve taken with the m<strong>in</strong>ussign.• A zero of F (V ) corresponds to an equilibrium of the system. (Indeed,if F (V ) = 0, then the state of the system, V , neither <strong>in</strong>creases nordecreases.)• An equilibrium is stable when F (V ) changes the sign from “+” to“−”. A sufficient condition for stability is that the eigenvalue λ =F ′ (V ) at the equilibrium be negative.• A phase portrait is a geometrical representation of the system’s dynamics.It depicts all equilibria, their stability, representative trajectories,and attraction doma<strong>in</strong>s.• A bifurcation is a qualitative change of the system’s phase portrait.• The saddle-node (fold) is a typical bifurcation <strong>in</strong> one-dimensionalsystems: As a parameter changes, a stable and an unstable equilibriumapproach, coalesce, and then annihilate each other.Bibliographical NotesThere is no standard textbook on dynamical systems theory. The classical book Nonl<strong>in</strong>earOscillations, <strong>Dynamical</strong> <strong>Systems</strong>, and Bifurcations of Vector Fields by Guckenheimerand Holmes (1983) plays the same role <strong>in</strong> the dynamical systems communityas the book Ion Channels of Excitable Membranes by Hille (2001) <strong>in</strong> the neurosciencecommunity. A common feature of these books is that they are not suitable as a firstread<strong>in</strong>g on the subject.Most textbooks on differential equations, such as Differential Equations and <strong>Dynamical</strong><strong>Systems</strong> by Perko (1996), develop the theory start<strong>in</strong>g with a comprehensiveanalysis of l<strong>in</strong>ear systems, then apply<strong>in</strong>g it to local analysis of non-l<strong>in</strong>ear systems, andthen discuss<strong>in</strong>g global behavior. By the time the reader gets to bifurcations, he hasto go through a lot of daunt<strong>in</strong>g math, which is fun only for mathematicians. Here we


One-Dimensional <strong>Systems</strong> 87follow approach similar to that <strong>in</strong> Nonl<strong>in</strong>ear Dynamics and Chaos by Strogatz (1994):Instead of go<strong>in</strong>g from l<strong>in</strong>ear to non-l<strong>in</strong>ear systems, we go from one-dimensional nonl<strong>in</strong>earsystems (this chapter) to two-dimensional non-l<strong>in</strong>ear systems (next chapter).Rather than burden<strong>in</strong>g the theory with a lot of mathematics, we use the geometricalapproach to stimulate the reader’s <strong>in</strong>tuition. (There is plenty of fun math <strong>in</strong> exercisesand <strong>in</strong> the later chapters.)Exercises1. Consider a neuron hav<strong>in</strong>g Na + current with fast activation k<strong>in</strong>etics. Assume that<strong>in</strong>activation of this current, as well as (<strong>in</strong>)activations of the other currents <strong>in</strong> theneuron are much slower. Prove that the <strong>in</strong>itial segment of action potential upstrokeof this neuron can be approximated by the I Na,p -model (3.5). Use Fig. 3.15to discuss the applicability of this approximation.2. Draw phase portraits of the systems <strong>in</strong> Fig. 3.36. Clearly mark all equilibria,their stability, attraction doma<strong>in</strong>s, and direction of trajectories. Determ<strong>in</strong>e thesigns of eigenvalues at each equilibrium.F(V) F(V) F(V)V V Va b cFigure 3.36: Draw phase portrait of the system ˙V = F (V ) with shown F (V ).3. Draw phase portraits of the follow<strong>in</strong>g systems(a) ẋ = −1 + x 2 ,(b) ẋ = x − x 3 .Determ<strong>in</strong>e the eigenvalues at each equilibrium.4. Determ<strong>in</strong>e stability of the equilibrium x = 0 and draw phase portraits of thefollow<strong>in</strong>g piece-wise cont<strong>in</strong>uous systems{ 2x, if x < 0(a) ẋ =x, if x ≥ 0⎧⎨ −1, if x < 0(b) ẋ = 0, if x = 0⎩1, if x > 0{ −2/x, if x ≠ 0(c) ẋ =0, if x = 0


88 One-Dimensional <strong>Systems</strong>VF (V)1VVF (V) 2F (V)1VF (V) 2VF (V) 2F (V)1VabcF (V)1VF (V)1VF (V)1VF (V) 2VF (V) 2VF (V) 2VdefFigure 3.37: Which of the pairs correspond to topologically equivalent dynamical systems?(All <strong>in</strong>tersections with the V axis are marked as dots.)5. Draw phase portraits of the systems <strong>in</strong> Fig. 3.37. Which of the pairs <strong>in</strong> the figurecorrespond to topologically equivalent dynamical systems?6. (Saddle-node bifurcation) Draw the bifurcation diagram and representative phaseportraits of the system ẋ = a + x 2 , where a is a bifurcation parameter. F<strong>in</strong>d theeigenvalues at each equilibrium.7. (Saddle-node bifurcation) Use def<strong>in</strong>ition <strong>in</strong> Sect. 3.3.4 to f<strong>in</strong>d saddle-node bifurcationpo<strong>in</strong>ts <strong>in</strong> the follow<strong>in</strong>g systems:(a) ẋ = a + 2x + x 2 ,(b) ẋ = a + x + x 2 ,(c) ẋ = a − x + x 2 ,(d) ẋ = a − x + x 3 (H<strong>in</strong>t: verify the non-hyperbolicity condition first),(e) ẋ = 1 + ax + x 2 ,(f) ẋ = 1 + 2x + ax 2 ,where a is the bifurcation parameter.8. (Pitchfork bifurcation) Draw the bifurcation diagram and representative phaseportraits of the system ẋ = bx − x 3 , where b is a bifurcation parameter. F<strong>in</strong>d theeigenvalues at each equilibrium.


One-Dimensional <strong>Systems</strong> 8910.80.60.40.2h (V)0-140 -70 0Membrane Voltage (mV)Current200-20-40-60I (V)LI (V)Kir-80E KE L-100-140 -70 0Membrane Voltage (mV)210-1-2F(V)-100 -50 0Membrane Voltage (mV)Figure 3.38: The I Kir -model hav<strong>in</strong>g <strong>in</strong>jected current (I), leak current (I L ), and <strong>in</strong>stantaneousK + <strong>in</strong>ward rectifier current (I Kir ) and described by (3.11). Inactivation curveh ∞ (V ) is modified from Wessel et. al (1999). Parameters: C = 1, I = 6, g L = 0.2,E L = −50, g Kir = 2, E K = −80, V 1/2 = −76, k = −12 (see Fig. 2.20).10.80.60.40.2m (V)0-100 -50 0 50 100Membrane Voltage (mV)Current100500-50I (V)LE L-100E Na-150-100 -50 0 50 100Membrane Voltage (mV)INa,p(V)500F(V)-50-100 -50 0 50 100Membrane Voltage (mV)Figure 3.39: The I Na,p -model with leak current (I L ) and persistent Na + current (I Na,p ),described by (3.5) with the right-hand side function F (V ). Parameters: C = 1, I = 0,g L = 1, E L = −80, g Na = 2.25, E Na = 60, V 1/2 = −20, k = 15 (see Fig. 2.20).9. Draw the bifurcation diagram of the I Kir -modelC ˙V = I − g L (V − E L ) −<strong>in</strong>stantaneous I Kir{ }} {g Kir h ∞ (V )(V − E K ) , (3.11)us<strong>in</strong>g parameters from Fig. 3.38 and treat<strong>in</strong>g I as a bifurcation parameter.10. Derive an explicit formula that relates the position of the equilibrium <strong>in</strong> theHodgk<strong>in</strong>-Huxley model to the magnitude of the <strong>in</strong>jected dc-current I. Are thereany saddle-node bifurcations?11. Draw the bifurcation diagram of the I Na,p -model (3.5) us<strong>in</strong>g parameters fromFig. 3.39 and treat<strong>in</strong>g(a) g L as a bifurcation parameter,(b) E L as a bifurcation parameter.


90 One-Dimensional <strong>Systems</strong>10.80.60.40.2m (V)0-120 -100 -80 -60 -40 -20 0Membrane Voltage (mV)Current100500E KE LI (V)LI (V)K-50-120 -100 -80 -60 -40 -20 0Membrane Voltage (mV)500F(V)-50-120 -100 -80 -60 -40 -20 0Membrane Voltage (mV)Figure 3.40: The I K -model with leak current (I L ) and persistent K + current (I K ),described by (3.12). Parameters: C = 1, g L = 1, E L = −80, g K = 1, E K = −90,V 1/2 = −53, k = 15 (see Fig. 2.20).10.80.60.40.2h (V)0-120 -100 -80 -60 -40 -20 0Membrane Voltage (mV)Current500-50I (V)LI (V)hE LE h-100-120 -100 -80 -60 -40 -20 0Membrane Voltage (mV)100500-50F(V)-100-120 -100 -80 -60 -40 -20 0Membrane Voltage (mV)Figure 3.41: The I h -model with leak current (I L ) and “hyperpolarization-activated”<strong>in</strong>ward current I h , described by (3.13). Parameters: C = 1, g L = 1, E L = −80, g h = 1,E h = −43, V 1/2 = −75, k = −5.5 (Huguenard and McCormick 1992).12. Draw the bifurcation diagram of the I Kir -model (3.11) us<strong>in</strong>g parameters fromFig. 3.38 and treat<strong>in</strong>g(a) g L as a bifurcation parameter,(b) g Kir as a bifurcation parameter.13. Show that the I K -model <strong>in</strong> Fig. 3.40C ˙V = −g L (V − E L ) −<strong>in</strong>stantaneous I K{ }} {g K m 4 ∞(V )(V − E K ) . (3.12)cannot exhibit saddle-node bifurcation for V > E K . (H<strong>in</strong>t: show that F ′ (V ) ≠ 0for all V > E K .)14. Show that the I h -model <strong>in</strong> Fig. 3.41C ˙V = −g L (V − E L ) −cannot exhibit saddle-node bifurcation for any V < E h .<strong>in</strong>stantaneous I h{ }} {g h h ∞ (V )(V − E h ) (3.13)


One-Dimensional <strong>Systems</strong> 9115. Prove that the upstroke of the spike <strong>in</strong> the quadratic <strong>in</strong>tegrate-and-fire neuron(3.9) has the asymptote 1/(c − t) for some c > 0.16. (Cusp bifurcation) Draw the bifurcation diagram and representative phase portraitsof the system ẋ = a + bx − x 3 , where a and b are bifurcation parameters.Plot the bifurcation diagram <strong>in</strong> the (a, b, x)-space and on the (a, b)-plane.17. (Gradient systems) An n-dimensional dynamical system ẋ = f(x), with x =(x 1 , . . . , x n ) ∈ R n is said to be gradient when there is a potential (energy) functionE(x) such thatẋ = − grad E(x) ,wheregrad E(x) = (E x1 , . . . , E xn )is the gradient of E(x). Show that all one-dimensional systems are gradient (H<strong>in</strong>t:see Fig. 3.11). F<strong>in</strong>d potential (energy) functions for the follow<strong>in</strong>g one-dimensionalsystemsa. ˙V = 0 , b. ˙V = 1 , c. ˙V = −V ,d. ˙V = −1 + V 2 , e. ˙V = V − V 3 , f. ˙V = − s<strong>in</strong> V .18. Consider a dynamical system ẋ = f(x) , x(0) = x 0 .(a) (Stability) An equilibrium y is stable if any solution x(t) with x 0 sufficientlyclose to y rema<strong>in</strong>s near y for all time. That is, for all ε > 0 there existsδ > 0 such that if |x 0 − y| < δ then |x(t) − y| < ε for all t ≥ 0.(b) (Asymptotic stability) A stable equilibrium y is asymptotically stable if allsolutions start<strong>in</strong>g sufficiently close to y approach it as t → ∞. That is, ifδ > 0 from the def<strong>in</strong>ition above can be chosen so that lim t→∞ x(t) = y.(c) (Exponential stability) A stable equilibrium y is said to be exponentiallystable when there is a constant a > 0 such that |x(t) − y| < exp(−at) for allx 0 near y and all t ≥ 0.Prove that (c) implies (b), and (b) implies (a). Show that (a) does not imply(b) and (b) does not imply (c); That is, present a system hav<strong>in</strong>g stable but notasymptotically stable equilibrium, and a system hav<strong>in</strong>g asymptotically but notexponentially stable equilibrium.19. (I NMDA -model) Show that voltage-depended activation of NMDA synaptic receptors<strong>in</strong> a passive dendritic tree with a constant concentration of glutamate ismathematically equivalent to the I Na,p -model.


92 One-Dimensional <strong>Systems</strong>


Chapter 4Two-Dimensional <strong>Systems</strong>In this chapter we <strong>in</strong>troduce methods of phase plane analysis of two-dimensional systems.Most concepts will be illustrated us<strong>in</strong>g the I Na,p +I K -model <strong>in</strong> Fig. 4.1:leak I L<strong>in</strong>stantaneous I Na,p IC ˙V{ }} { { }} { { }} K{= I − g L (V −E L ) − g Na m ∞ (V ) (V −E Na ) − g K n (V −E K ) , (4.1)ṅ = (n ∞ (V ) − n)/τ(V ) , (4.2)hav<strong>in</strong>g leak current I L , persistent Na + current I Na,p with <strong>in</strong>stantaneous activation k<strong>in</strong>eticand a relatively slower persistent K + current I K with either high (Fig. 4.1a) orlow (Fig. 4.1b) threshold (the two choices result <strong>in</strong> fundamentally different dynamics).The state of the I Na,p +I K -model is a two-dimensional vector (V, n) ∈ R 2 on the phaseplane R 2 . New types of equilibria, orbits, and bifurcations can exist on the phase planethat cannot exist on the phase l<strong>in</strong>e R. Many <strong>in</strong>terest<strong>in</strong>g features of s<strong>in</strong>gle neuron dynamicscan be illustrated or expla<strong>in</strong>ed us<strong>in</strong>g two-dimensional systems. Even neuronalburst<strong>in</strong>g, which occurs <strong>in</strong> multi-dimensional systems, can be understood via bifurcationanalysis of two-dimensional systems.This model is equivalent <strong>in</strong> many respects to the well-known and widely usedI Ca +I K -model proposed by Morris and Lecar (1981) to describe voltage oscillations<strong>in</strong> the barnacle giant muscle fiber.4.1 Planar Vector FieldsTwo-dimensional dynamical systems, also called planar systems, are often written <strong>in</strong>the formẋ = f(x, y) ,ẏ = g(x, y) ,where the functions f and g describe the evolution of the two-dimensional state variable(x(t), y(t)). For any po<strong>in</strong>t (x 0 , y 0 ) on the phase plane the vector (f(x 0 , y 0 ), g(x 0 , y 0 ))93


94 Two-Dimensional <strong>Systems</strong>I Na,pneuronI KI0high-thresholdI KI(V)=I L +I Na,p +I KILa bcurrent (pA)100-1000low-thresholdI(V)=I L +I Na,p +I KI K-100 E -50 0 50KE Namembrane potential, V (mV)-100E -50 0 50KE Namembrane potential, V (mV)Figure 4.1: The I Na,p +I K -model (4.1, 4.2). Parameters <strong>in</strong> (a): C = 1, I = 0, E L = −80mV, g L = 8, g Na = 20, g K = 10, m ∞ (V ) has V 1/2 = −20 and k = 15, n ∞ (V ) hasV 1/2 = −25 and k = 5, and τ(V ) = 1, E Na = 60 mV and E K = −90 mV. Parameters<strong>in</strong> (b) as <strong>in</strong> (a) except E L = −78 mV and n ∞ (V ) has V 1/2 = −45; see Sect. 2.3.5.Figure 4.2: Harold Lecar (back), RichardFitzHugh (front), and Cathy Morris atNIH Biophysics Lab, summer of 1983.


Two-Dimensional <strong>Systems</strong> 9510886644y20y20-2-2-4-4-6-6-8-8-1010-10 -8 -6 -4 -2 0 2 4 6 8 10xa-10-10 -8 -6 -4 -2 0 2 4 6 810xb886644y20-2-4-6-8-10-10 -8 -6 -4 -2 0 2 4 6 8 10xcy20-2-4-6-8-10-10 -8 -6 -4 -2 0 2 4 6 8 10xdFigure 4.3: Examples of vector fields.<strong>in</strong>dicates the direction of change of the state variable. For example, negative f(x 0 , y 0 )and positive g(x 0 , y 0 ) imply that x(t) decreases and y(t) <strong>in</strong>creases at this particularpo<strong>in</strong>t. S<strong>in</strong>ce each po<strong>in</strong>t on the phase plane (x, y) has its own vector (f, g), the systemabove is said to def<strong>in</strong>e a vector field on the plane, also known as direction field or velocityfield, see Fig. 4.3. Thus, the vector field def<strong>in</strong>es the direction of motion; depend<strong>in</strong>g onwhere you are, it tells you where you are go<strong>in</strong>g.Let us consider a few examples. The two-dimensional systemẋ = 1 ,ẏ = 0def<strong>in</strong>es a constant horizontal vector field <strong>in</strong> Fig. 4.3a s<strong>in</strong>ce each po<strong>in</strong>t has a horizontalvector (1, 0) attached to it. (Of course, we depict only a small sample of vectors.)Similarly, the systemẋ = 0 ,ẏ = 1


96 Two-Dimensional <strong>Systems</strong>def<strong>in</strong>es a constant vertical vector field depicted <strong>in</strong> Fig. 4.3b. The systemẋ = −x ,ẏ = −ydef<strong>in</strong>es a vector field that po<strong>in</strong>ts to the orig<strong>in</strong> (0, 0), as <strong>in</strong> Fig. 4.3c, and the systemẋ = −y , (4.3)ẏ = −x (4.4)def<strong>in</strong>es a saddle vector field, as <strong>in</strong> Fig. 4.3d. Vector fields provide geometrical <strong>in</strong>formationabout the jo<strong>in</strong>t evolution of state variables. For example, the vector field <strong>in</strong>Fig. 4.3d is directed rightward <strong>in</strong> the lower half-plane and leftward <strong>in</strong> the upper halfplane.Therefore the variable x(t) <strong>in</strong>creases when y < 0 and decreases otherwise, whichobviously follows from the equation (4.3). Quite often however geometrical analysis ofvector fields can provide <strong>in</strong>formation about the behavior of the system that may notbe obvious from the form of the functions f and g.4.1.1 Nullcl<strong>in</strong>esThe vector field <strong>in</strong> Fig. 4.3d is directed rightward (x <strong>in</strong>creases) or leftward (x decreases)<strong>in</strong> different regions of the phase plane. The set of po<strong>in</strong>ts where the vectorfield changes its horizontal direction is called x-nullcl<strong>in</strong>e, and it is def<strong>in</strong>ed by the equationf(x, y) = 0. Indeed, at any such po<strong>in</strong>t x neither <strong>in</strong>creases nor decreases becauseẋ = 0. The x-nullcl<strong>in</strong>e partitions the phase plane <strong>in</strong>to two regions where x moves <strong>in</strong>opposite directions. Similarly, the y-nullcl<strong>in</strong>e is def<strong>in</strong>ed by the equation g(x, y) = 0,and it denotes the set of po<strong>in</strong>ts where the vector field changes its vertical direction.This nullcl<strong>in</strong>e partitions the phase plane <strong>in</strong>to two regions where y either <strong>in</strong>creases ordecreases. The x- and y-nullcl<strong>in</strong>es partition the phase plane <strong>in</strong>to 4 different regions:(a) x and y <strong>in</strong>crease, (b) x decreases, y <strong>in</strong>creases, (c) x and y decrease, and (d) x<strong>in</strong>creases, y decreases, as we illustrate <strong>in</strong> Fig. 4.4.Each po<strong>in</strong>t of <strong>in</strong>tersection of the nullcl<strong>in</strong>es is an equilibrium po<strong>in</strong>t, s<strong>in</strong>ce f(x, y) =g(x, y) = 0 and hence ẋ = ẏ = 0. Conversely, every equilibrium of a two-dimensionalsystem is the po<strong>in</strong>t of <strong>in</strong>tersection of its nullcl<strong>in</strong>es. Because nullcl<strong>in</strong>es are so important,we consider two examples <strong>in</strong> detail below (the reader is urged to solve Ex. 1 at the endof this chapter).Let us determ<strong>in</strong>e nullcl<strong>in</strong>es of the system (4.3, 4.4) with the vector field shown <strong>in</strong>Fig. 4.3d. From (4.3) it follows that x-nullcl<strong>in</strong>e is the horizontal l<strong>in</strong>e y = 0, and from(4.4) it follows that y-nullcl<strong>in</strong>e is the vertical l<strong>in</strong>e x = 0. These nullcl<strong>in</strong>es (dashed l<strong>in</strong>es<strong>in</strong> Fig. 4.3d) partition the phase plane <strong>in</strong>to 4 quadrants, <strong>in</strong> each of which the vectorfield has a different direction. The <strong>in</strong>tersection of the nullcl<strong>in</strong>es is the equilibrium(0, 0). Later <strong>in</strong> this chapter we will study how to determ<strong>in</strong>e stability of equilibria <strong>in</strong>two-dimensional systems, though <strong>in</strong> this particular case one can easily guess that theequilibrium is not stable.


Two-Dimensional <strong>Systems</strong> 970.7K + activation variable, n0.60.50.40.30.20.1(d)relativerefractory(c)absoluterefractoryn-nullcl<strong>in</strong>e(b)spikedownstroke(a)spikeupstroke(regenerative)peakof spikeV-nullcl<strong>in</strong>e0rest<strong>in</strong>g-80 -70 -60 -50 -40 -30 -20 -10 0 10 20membrane voltage, V (mV)Figure 4.4: Nullcl<strong>in</strong>es of the I Na,p +I K -model (4.1, 4.2) with low-threshold K + current<strong>in</strong> Fig. 4.1b. (The vector field is slightly distorted for the sake of clarity of illustration).As another example, let us determ<strong>in</strong>e the nullcl<strong>in</strong>es of the I Na,p +I K -model (4.1,4.2). The V -nullcl<strong>in</strong>e is given by the equationwhich has the solutionI − g L (V −E L ) − g Na m ∞ (V ) (V −E Na ) − g K n (V −E K ) = 0 ,n = I − g L(V −E L ) − g Na m ∞ (V ) (V −E Na )g K (V − E K )(V -nullcl<strong>in</strong>e)depicted <strong>in</strong> Fig. 4.4. It typically has the form of a cubic parabola. The equationdef<strong>in</strong>es the n-nullcl<strong>in</strong>en = n ∞ (V )n ∞ (V ) − n = 0(n-nullcl<strong>in</strong>e),which co<strong>in</strong>cides with the K + steady-state activation function n ∞ (V ), though only an<strong>in</strong>itial segment of this curve fits <strong>in</strong> Fig. 4.4. It is easy to see how the V - and n-nullcl<strong>in</strong>es partition the phase plane <strong>in</strong>to four regions, <strong>in</strong> each of which the vector fieldhas a different direction:(a) Both V and n <strong>in</strong>crease: Both Na + and K + currents activate and lead to theupstroke of the action potential.(b) V decreases but n still <strong>in</strong>creases: Na + current deactivates but the slower K +current still activates and leads to the downstroke of the action potential.


98 Two-Dimensional <strong>Systems</strong>(f(x(t),y(t)),g(x(t),y(t)))y(x(t),y(t))(f(x 0 ,y 0 ),g(x 0 ,y 0 ))(x 0 ,y 0 )xFigure 4.5: Solutions are trajectories tangent to the vector field.(c) Both V and n decrease: Both Na + and K + currents deactivate while V is small,lead<strong>in</strong>g to a refractory period.(d) V <strong>in</strong>creases but n still decreases: Partial activation of Na + current comb<strong>in</strong>edwith further deactivation of the residual K + current lead to a relative refractoryperiod, then to an excitable period, and possibly to another action potential.The <strong>in</strong>tersection of the V - and n-nullcl<strong>in</strong>es <strong>in</strong> Fig. 4.4 is an equilibrium correspond<strong>in</strong>gto the rest state. The number and location of equilibria might be difficult to <strong>in</strong>fervia analysis of equations (4.1, 4.2), but it is a trivial geometrical exercise once the nullcl<strong>in</strong>esare determ<strong>in</strong>ed. Because nullcl<strong>in</strong>es are so useful and important <strong>in</strong> geometricalanalysis of dynamical systems, few scientists bother to plot vector fields. Follow<strong>in</strong>gthis tradition, we will not show vector fields <strong>in</strong> the rest of the book (except for thischapter). Instead, we plot nullcl<strong>in</strong>es and representative trajectories, which we discussnext.4.1.2 TrajectoriesA vector-function (x(t), y(t)) is a solution of the two-dimensional systemẋ = f(x, y) ,ẏ = g(x, y) ,start<strong>in</strong>g with an <strong>in</strong>itial condition (x(0), y(0)) = (x 0 , y 0 ) when dx(t)/dt = f(x(t), y(t))and dy(t)/dt = g(x(t), y(t)) at each t ≥ 0. This requirement has a simple geometrical<strong>in</strong>terpretation: A solution is a curve (x(t), y(t)) on the phase plane R 2 which is tangentto the vector field, as we illustrate <strong>in</strong> Fig 4.5. Such a curve is often called a trajectoryor an orbit.One can th<strong>in</strong>k of the vector field as a stationary flow of a fluid. Then a solution isjust a trajectory of a small particle dropped at a certa<strong>in</strong> (<strong>in</strong>itial) po<strong>in</strong>t and carried bythe flow. To study the flow, it is useful to drop a few particles and see where they are


Two-Dimensional <strong>Systems</strong> 99108642y0-2-4-6-8-10-10 -8 -6 -4 -2 0 2 4 6 8 10xFigure 4.6: Representative trajectories of the two-dimensional system (4.3,4.4).go<strong>in</strong>g. Thus, to understand the geometry of a vector field, it is always useful to plot afew representative trajectories start<strong>in</strong>g from various <strong>in</strong>itial po<strong>in</strong>ts, as we do <strong>in</strong> Fig. 4.6.Due to the uniqueness of the solutions, the trajectories cannot cross, so they partitionthe phase space <strong>in</strong>to various regions. This is an important step toward determ<strong>in</strong><strong>in</strong>gthe phase portrait of a two-dimensional system.Let us return to the I Na,p +I K -model (4.1, 4.2) with low-threshold K + current andexpla<strong>in</strong> two odd phenomena discussed <strong>in</strong> the first chapter: Failure to generate all-ornoneaction potentials (Fig. 1.5b) and <strong>in</strong>ability to have a fixed value of the thresholdvoltage. Brief and strong current pulses <strong>in</strong> Fig. 4.7 reset the value of the voltage variableV but do not change the value of the K + activation variable n. Thus, each voltagetrace after the pulse corresponds to a trajectory start<strong>in</strong>g with different values of V 0 butthe same value n 0 . We see that each trajectory makes a counter-clockwise excursionand returns to the rest state. However, the size of the excursion depends on the <strong>in</strong>itialvalue of the voltage variable and can be small (subthreshold response), <strong>in</strong>termediate, orlarge (action potential). This phenomenon was considered theoretically by FitzHugh <strong>in</strong>early sixties (see bibliography) and demonstrated experimentally by Cole et al. (1970)us<strong>in</strong>g squid giant axon at higher than normal temperatures.In Fig. 4.8 we apply a long pre-pulse current of various amplitudes to reset theK + activation variable n to various values, and then a brief strong pulse to reset Vto exactly −48 mV. Each voltage trace after the pulse corresponds to a trajectorystart<strong>in</strong>g with the same V 0 = −48 mV, but different values of n 0 . We see that sometrajectories return immediately to the rest state while others do so after generat<strong>in</strong>g anaction potential. Therefore, V = −48 mV is a subthreshold value when n 0 is large,and a superthreshold one otherwise.


100 Two-Dimensional <strong>Systems</strong>0.70action potential0.6K + activation variable, n0.50.40.30.20.1subthresholdn-nullcl<strong>in</strong>eV-nullcl<strong>in</strong>emembrane voltage, V (mV)-20-40-60<strong>in</strong>termediateamplitudeaction potentials0n 0-80 -70 -60 -50 -40 -30 -20 -10 0 10 200 2 4 6membrane voltage, V (mV)-80time (ms)Figure 4.7: Failure to generate all-or-none action potentials <strong>in</strong> the I Na,p +I K -model (4.1,4.2)K + activation variable, n0.70.60.50.40.30.20.10n-nullcl<strong>in</strong>esubthresholdaction potentialV-nullcl<strong>in</strong>e-80 -70 -60 -50 -40 -30 -20 -10 0 10 20membrane voltage, V (mV)0I=-10I=+10 pre-pulse-10I=0-20-30-40-50-60I=0I=-10membrane voltage, V (mV)I=+101 msFigure 4.8: Failure to have a fixed value of threshold voltage <strong>in</strong> the I Na,p +I K -model(4.1, 4.2).


Two-Dimensional <strong>Systems</strong> 101stableunstable4.1.3 Limit cyclesFigure 4.9: Limit cycles (periodic orbits).A trajectory that forms a closed loop is called a periodic trajectory or a periodic orbit(the latter is usually reserved for mapp<strong>in</strong>gs, which we do not consider here). Sometimesperiodic trajectories are isolated, as <strong>in</strong> Fig. 4.9, sometimes they are part of a cont<strong>in</strong>uum,as <strong>in</strong> Fig. 4.13, left. An isolated periodic trajectory is called a limit cycle. The existenceof limit cycles is a major feature of two-dimensional systems that cannot exist <strong>in</strong> R 1 .If the <strong>in</strong>itial po<strong>in</strong>t is on a limit cycle, then the solution (x(t), y(t)) stays on the cycleforever, and the system exhibits periodic behavior; i.e.,x(t) = x(t + T ) and y(t) = y(t + T ) (for all t)for some T > 0. The m<strong>in</strong>imal T for which this equality holds is called the periodof the limit cycle. A limit cycle is said to be asymptotically stable if any trajectorywith the <strong>in</strong>itial po<strong>in</strong>t sufficiently near the cycle approaches the cycle as t → ∞. Suchasymptotically stable limit cycles are often called limit cycle attractors, s<strong>in</strong>ce they“attract” all nearby trajectories. The stable limit cycle <strong>in</strong> Fig. 4.9 is an attractor.The limit cycle <strong>in</strong> Fig. 4.10 is also an attractor; It corresponds to the periodic (tonic)spik<strong>in</strong>g of the I Na,p +I K -model (4.1, 4.2). The unstable limit cycle <strong>in</strong> Fig. 4.9 is oftencalled a repeller, s<strong>in</strong>ce it repels all nearby trajectories. Notice that there is always atleast one equilibrium <strong>in</strong>side any limit cycle on a plane.In Fig. 4.11 we depict limit cycles of three types of neurons recorded <strong>in</strong> vitro.S<strong>in</strong>ce we do not know the state of the <strong>in</strong>ternal variables, such as the magnitude of theactivation and <strong>in</strong>activation of Na + and K + currents, we plot the cycles on the (V, V ′ )-plane, where V ′ is the time derivative of V . The cycles look jerky because of the poordata sampl<strong>in</strong>g rate dur<strong>in</strong>g each spike.4.1.4 Relaxation OscillatorsMany models <strong>in</strong> science and eng<strong>in</strong>eer<strong>in</strong>g can be reduced to two-dimensional fast/slowsystems of the formẋ = f(x, y) (fast variable)


102 Two-Dimensional <strong>Systems</strong>0.70.6K + activation variable, n0.50.40.3n-nullcl<strong>in</strong>eV-nullcl<strong>in</strong>e0.20.1membrane voltage, V (mV)00-20-40-60-80 -70 -60 -50 -40 -30 -20 -10 0 10 20membrane voltage, V (mV)-800 5 10 15 20 25 30time (ms)Figure 4.10: Stable limit cycle <strong>in</strong> the I Na,p +I K -model (4.1, 4.2) with low-threshold K +current and I = 40.


Two-Dimensional <strong>Systems</strong> 103derivative, V' (mV/ms)membranepotential (mV)2001000-100-80 -60 -40 -20 0 20 40membrane potential, V (mV)40200-20-40cortical pyramidal neuron cortical <strong>in</strong>terneuron bra<strong>in</strong>stem neuron-600 50 100-80 -60 -40 -20 0 20 40membrane potential, V (mV)-80 -60 -40 -20 0 20 40membrane potential, V (mV)0 20 40 60 0 10 20time (ms)Figure 4.11: Limit cycles correspond<strong>in</strong>g to tonic spik<strong>in</strong>g of three types of neuronsrecorded <strong>in</strong> vitro.ẏ = µg(x, y) (slow variable)where small parameter µ describes the ratio of time scales of variables x and y. Typically,fast variable x has a cubic-like nullcl<strong>in</strong>e that <strong>in</strong>tersects the y-nullcl<strong>in</strong>e somewhere<strong>in</strong> the middle branch, as <strong>in</strong> Fig. 4.12a, result<strong>in</strong>g <strong>in</strong> relaxation oscillations. The periodictrajectory of the system slides down along the left (stable) branch of the cubic nullcl<strong>in</strong>euntil it reaches the left knee A. At this moment, it quickly jumps to the po<strong>in</strong>t B andthen slowly slides up along the right (also stable) branch of the cubic nullcl<strong>in</strong>e. Uponreach<strong>in</strong>g the right knee C, the system jumps to the left branch and starts to slide downaga<strong>in</strong>, thereby complet<strong>in</strong>g one oscillation. Relaxation oscillations are easy to graspconceptually, but some of their features are quite difficult to study mathematically.We consider relaxation oscillations <strong>in</strong> detail <strong>in</strong> Sect. 6.3.4.Notice that the jumps <strong>in</strong> Fig. 4.12a are nearly horizontal — a dist<strong>in</strong>ctive signature ofrelaxation oscillations that is due to the disparately different time scales <strong>in</strong> the system.Although many neuronal models have fast and slow time scales and could be reducedto the fast/slow form above, they do not exhibit relaxation oscillations because theparameter µ is not small enough. Anybody who records from neurons would probablynotice the weird square shape of “spikes” <strong>in</strong> Fig. 4.12b, someth<strong>in</strong>g that most biologicalneurons do not exhibit. Nevertheless, relaxation oscillations <strong>in</strong> fast/slow systems areimportant when we consider neuronal burst<strong>in</strong>g <strong>in</strong> Chap. 9, though the fast variable xis two-dimensional there.


104 Two-Dimensional <strong>Systems</strong>y1.510.50Dy=x-x3/3x=0C(a)x(t)2.521.510.50BCBC(b)-0.51-1.5AB-2 -1 0 1 2x-0.5-1AA-1.5-2DD-2.50 100 200 300 400time, tFigure 4.12: Relaxation oscillations <strong>in</strong> the van der Pol model ẋ = x − x 3 /3 − y, ẏ = µxwith µ = 0.01.4.2 EquilibriaAn important step <strong>in</strong> the analysis of any dynamical system is to f<strong>in</strong>d its equilibria, i.e.,po<strong>in</strong>ts wheref(x, y) = 0 ,g(x, y) = 0(po<strong>in</strong>t (x, y) is an equilibrium).As we mentioned before, equilibria are <strong>in</strong>tersections of nullcl<strong>in</strong>es. If the <strong>in</strong>itial po<strong>in</strong>t(x 0 , y 0 ) is an equilibrium, then ẋ = 0 and ẏ = 0, and the trajectory stays at equilibrium;that is, x(t) = x 0 and y(t) = y 0 for all t ≥ 0. If the <strong>in</strong>itial po<strong>in</strong>t is near the equilibrium,then the trajectory may converge to or diverge from the equilibrium depend<strong>in</strong>g on itsstability.From the electrophysiological po<strong>in</strong>t of view, any equilibrium of a neuronal modelis the zero cross<strong>in</strong>g of its steady-state I-V relation I ∞ (V ). For example, the I Na,p +I K -model (4.1, 4.2) with high-threshold K + current has an I-V curve with three zeroes(Fig. 4.1a), hence it has three equilibria: around −66 mV, −56 mV, and −28 mV. Incontrast, the same model with low-threshold K + current has a monotonic I-V curvewith only one zero (Fig. 4.1b), hence it has a unique equilibrium, which is around −61mV.4.2.1 StabilityIn Chap. 3, Exercise 18 we provide rigorous def<strong>in</strong>itions of stability of equilibria <strong>in</strong>one-dimensional systems. The same def<strong>in</strong>itions apply to higher-dimensional systems.


Two-Dimensional <strong>Systems</strong> 105Figure 4.13: Neutrally stable equilibria.diverge from the equilibria.Some trajectories neither converge to norabFigure 4.14: Unstable equilibria.Briefly, an equilibrium is stable if any trajectory start<strong>in</strong>g sufficiently close to the equilibriumrema<strong>in</strong>s near it for all t ≥ 0. If, <strong>in</strong> addition, all such trajectories convergeto the equilibrium as t → ∞, the equilibrium is asymptotically stable, as <strong>in</strong> Fig. 4.3c.When the convergence rate is exponential or faster, then the equilibrium is said tobe exponentially stable. Notice that stability does not imply asymptotic stability. Forexample, all equilibria <strong>in</strong> Fig. 4.13 are stable but not asymptotically stable. They areoften referred to as be<strong>in</strong>g neutrally stable.An equilibrium is called unstable, if it is not stable. Obviously, if all nearby trajectoriesdiverge from the equilibrium, as <strong>in</strong> Fig. 4.14a, then it is unstable. This, however,is an exceptional case. For <strong>in</strong>stability it suffices to have at least one trajectory thatdiverges from the equilibrium no matter how close the <strong>in</strong>itial condition to the equilibriumis, as <strong>in</strong> Fig. 4.14b. Indeed, any trajectory start<strong>in</strong>g <strong>in</strong> the shaded area (attractiondoma<strong>in</strong>) converges to the equilibrium, but any trajectory start<strong>in</strong>g <strong>in</strong> the white areadiverges from it regardless of how close the <strong>in</strong>itial po<strong>in</strong>t to the equilibrium is.In contrast to the one-dimensional case, the stability of a two-dimensional equilibriumcannot be <strong>in</strong>ferred from the slope of the steady-state I-V curve. For example, theequilibrium around −28 mV <strong>in</strong> Fig. 4.1a is unstable even though the I-V curve haspositive slope.To determ<strong>in</strong>e the stability of an equilibrium, we need to look at the behavior of thetwo-dimensional vector field <strong>in</strong> a small neighborhood of the equilibrium. Quite oftenvisual <strong>in</strong>spection of the vector field does not give conclusive <strong>in</strong>formation about stability.


106 Two-Dimensional <strong>Systems</strong>For example, is the equilibrium <strong>in</strong> Fig. 4.4 stable? What about the equilibrium <strong>in</strong>Fig. 4.10? The vector fields <strong>in</strong> the neighborhoods of the two equilibria exhibit subtledifferences that are difficult to spot without the help of analytical tools, which wediscuss next.4.2.2 Local l<strong>in</strong>ear analysisBelow we rem<strong>in</strong>d the reader some basic concepts of l<strong>in</strong>ear algebra, assum<strong>in</strong>g that thereader has some familiarity with matrices, eigenvectors and eigenvalues. Consider atwo-dimensional dynamical systemẋ = f(x, y) (4.5)ẏ = g(x, y) (4.6)hav<strong>in</strong>g an equilibrium po<strong>in</strong>t (x 0 , y 0 ). The nonl<strong>in</strong>ear functions f and g can be l<strong>in</strong>earizednear the equilibrium; i.e., written <strong>in</strong> the formf(x, y) = a(x − x 0 ) + b(y − y 0 ) + higher-order terms,g(x, y) = c(x − x 0 ) + d(y − y 0 ) + higher-order terms,where higher-order terms <strong>in</strong>clude (x − x 0 ) 2 , (x − x 0 )(y − y 0 ), (x − x 0 ) 3 , etc., anda = ∂f∂x (x 0, y 0 ),c = ∂g∂x (x 0, y 0 ),b = ∂f∂y (x 0, y 0 ),d = ∂g∂y (x 0, y 0 ),are the partial derivatives of f and g with respect of the state variables x and yevaluated at the equilibrium (x 0 , y 0 ) (first, evaluate the derivatives, then substitutex = x 0 and y = y 0 ). Many questions regard<strong>in</strong>g the stability of the equilibrium can beanswered by consider<strong>in</strong>g the correspond<strong>in</strong>g l<strong>in</strong>ear system˙u = au + bw , (4.7)ẇ = cu + dw , (4.8)where u = x − x 0 and w = y − y 0 are the deviations from the equilibrium, and thehigher-order terms, u 2 , uw, w 3 , etc., are neglected. We can write this system <strong>in</strong> thevector form( ) ( ) ( )˙u a b u=.ẇ c d wThe l<strong>in</strong>earization matrix( ) a bL =c d


Two-Dimensional <strong>Systems</strong> 107is called the Jacobian matrix of the system (4.5, 4.6) at the equilibrium (x 0 , y 0 ). Forexample, the Jacobian matrix of the system (4.3, 4.4) at the orig<strong>in</strong> is( ) 0 −1. (4.9)−1 0It is important to remember that Jacobian matrices are def<strong>in</strong>ed for equilibria, andthat a nonl<strong>in</strong>ear system can have many equilibria and hence many different Jacobianmatrices.4.2.3 Eigenvalues and eigenvectorsA non-zero vector v ∈ R 2 is said to be an eigenvector of the matrix L correspond<strong>in</strong>gto the eigenvalue λ ifLv = λv (matrix notation) .For example, the matrix (4.9) has two eigenvectors( ) 1v 1 =and v12 =( 1−1correspond<strong>in</strong>g to the eigenvalues λ 1 = −1 and λ 2 = 1, respectively. Any textbook onl<strong>in</strong>ear algebra expla<strong>in</strong>s how to f<strong>in</strong>d eigenvectors and eigenvalues of an arbitrary matrix.It is important for the reader to get comfortable with these notions, s<strong>in</strong>ce they are usedextensively <strong>in</strong> the rest of the book.Eigenvalues play important role <strong>in</strong> analysis of stability of equilibria. To f<strong>in</strong>d theeigenvalues of a 2×2-matrix L, one solves the characteristic equation( )a − λ bdet= 0 .c d − λThis equation can be written <strong>in</strong> the polynomial form (a − λ)(d − λ) − bc = 0 orwhereλ 2 − τλ + ∆ = 0 ,τ = tr L = a + d and ∆ = det L = ad − bcare the trace and the determ<strong>in</strong>ant of the matrix L, respectively.polynomial has two solutions of the form)Such a quadraticλ 1 = τ + √ τ 2 − 4∆2andλ 2 = τ − √ τ 2 − 4∆2(4.10)and they are either real (when τ 2 −4∆ ≥ 0) or complex-conjugate (when τ 2 −4∆ < 0).What can you say about the case τ 2 = 4∆?


108 Two-Dimensional <strong>Systems</strong>In general, 2 × 2-matrices have two eigenvalues with dist<strong>in</strong>ct (<strong>in</strong>dependent) eigenvectors.In this case a general solution of the l<strong>in</strong>ear system has the form( ) u(t)= cw(t) 1 e λ1t v 1 + c 2 e λ2t v 2 ,where c 1 and c 2 are constants that depend on the <strong>in</strong>itial condition. This formula is validfor real and complex-conjugate eigenvalues. When both eigenvalues are negative (orhave negative real parts), u(t) → 0 and w(t) → 0, mean<strong>in</strong>g x(t) → x 0 and y(t) → y 0 ,so that the equilibrium (x 0 , y 0 ) is exponentially (and hence asymptotically) stable. It isunstable when at least one eigenvalue is positive or has a positive real part. We denotestable equilibria as filled circles • and unstable equilibria as open circles ◦ throughoutthe book.4.2.4 Local equivalenceAn equilibrium whose Jacobian matrix does not have zero eigenvalues or eigenvalueswith zero real part is called hyperbolic. Such an equilibrium can be stable or unstable.The Hartman-Grobman theorem states that the vector-field and hence the dynamics ofa nonl<strong>in</strong>ear system, e.g., (4.5, 4.6) near such a hyperbolic equilibrium is topologicallyequivalent to that of its l<strong>in</strong>earization (4.7, 4.8). That is, the higher-order terms that areneglected when (4.5, 4.6) is substituted by (4.7, 4.8) do not play any qualitative role.Thus, understand<strong>in</strong>g and classify<strong>in</strong>g the geometry of vector-fields of l<strong>in</strong>ear systemsprovides an exhaustive description of all possible behaviors of nonl<strong>in</strong>ear systems nearhyperbolic equilibria.A zero eigenvalue (or eigenvalues with zero real parts) arise when the equilibriumundergoes a bifurcation, e.g., as <strong>in</strong> Fig. 4.14b, and such equilibria are called nonhyperbolic.L<strong>in</strong>ear analysis cannot answer the question of stability of a nonl<strong>in</strong>earsystem <strong>in</strong> this case, s<strong>in</strong>ce small nonl<strong>in</strong>ear (high-order) terms play a crucial role here.We denote equilibria undergo<strong>in</strong>g a bifurcation as half-filled circles, e.g., .4.2.5 Classification of equilibriaBesides def<strong>in</strong><strong>in</strong>g the stability of an equilibrium, the eigenvalues also def<strong>in</strong>e the geometryof the vector field near the equilibrium, as we illustrate <strong>in</strong> Fig. 4.15 and ask the readerto prove <strong>in</strong> Ex. 4. (The proof is a straightforward consequence of (4.10)). There arethree major types of equilibria:Node (Fig. 4.16): The eigenvalues are real and of the same sign. The node is stablewhen the eigenvalues are negative, and unstable when they are positive. Thetrajectories tend to converge to or diverge from the node along the eigenvectorcorrespond<strong>in</strong>g to the eigenvalue hav<strong>in</strong>g the smallest absolute value.Saddle (Fig. 4.17): The eigenvalues are real and of opposite signs. Saddles are alwaysunstable, s<strong>in</strong>ce one of the eigenvalues is always positive. Most trajectories


Two-Dimensional <strong>Systems</strong> 109τ0eigenvaluessaddle(real eigenvalues, different signs)saddle-node bifurcation saddle-node bifurcation(real positive eigenvalues)unstable nodeunstable focus(complex eigenvalues,positive real part)Andronov-Hopf bifurcationstable focus(complex eigenvalues,negative real part)stable node(real negative eigenvalues)τ2 − 4∆ = 0τ2 − 4∆ = 0∆0Figure 4.15: Classification of equilibria accord<strong>in</strong>g to the trace (τ) and the determ<strong>in</strong>ant(∆) of the Jacobian matrix L. The shaded region corresponds to stable equilibria.approach the saddle equilibrium along the eigenvector correspond<strong>in</strong>g to the negative(stable) eigenvalue and then diverge from it along the eigenvector correspond<strong>in</strong>gto the positive (unstable) eigenvalue.Focus (Fig. 4.18): The eigenvalues are complex-conjugate. Foci are stable when theeigenvalues have negative real parts, and unstable when the eigenvalues have positivereal parts. The imag<strong>in</strong>ary part of the eigenvalues determ<strong>in</strong>es the frequencyof rotation of trajectories around the focus equilibrium.When the system undergoes a saddle-node bifurcation, one of the eigenvalues becomeszero and a mixed type of equilibrium occurs — saddle-node equilibrium, illustrated<strong>in</strong> Fig. 4.14b. There could be other types of mixed equilibria, such as saddle-focus,focus-node, etc., <strong>in</strong> dynamical systems hav<strong>in</strong>g dimension three and higher.Depend<strong>in</strong>g upon the value of the <strong>in</strong>jected current I, the I Na,p +I K -model (4.1, 4.2)with a low-threshold K + current has a stable focus (Fig. 4.8) or an unstable focus(Fig. 4.10) surrounded by a stable limit cycle. In Fig. 4.19 we depict the vector fieldand nullcl<strong>in</strong>es of the same model with a high-threshold K + current. As one expectsfrom the shape of the steady-state I-V curve <strong>in</strong> Fig. 4.1a, the model has three equilibria:a stable node, a saddle, and an unstable focus. Notice that the third equilibrium isunstable even though the I-V relation has a positive slope around it.Also notice that the y-axis starts at the negative value -0.1. However, the gat<strong>in</strong>gvariable n represents the proportion (probability) of the K + channels <strong>in</strong> the open state,hence a value less than zero has no physical mean<strong>in</strong>g. So while we can happily calculatethe nullcl<strong>in</strong>es for the negative n, and even start the trajectory with <strong>in</strong>itial condition


110 Two-Dimensional <strong>Systems</strong>v 2 v 2v 1v 1stable nodeunstable nodeFigure 4.16: Node equilibrium occurs when both eigenvalues are real and have the samesign, e.g., λ 1 = −1 and λ 2 = −3 (stable) or λ 1 = +1 and λ 2 = +3 (unstable). Mosttrajectories converge to or diverge from the node along the eigenvector v 1 correspond<strong>in</strong>gto the eigenvalue hav<strong>in</strong>g the smallest absolute value.v 2v 2v 1v 1saddlesaddleFigure 4.17: Saddle equilibrium occurs when two real eigenvalues have opposite signs,e.g., λ 1 = +1 and λ 2 = −1. Most trajectories diverge from the equilibrium along theeigenvector correspond<strong>in</strong>g to the positive eigenvalue (<strong>in</strong> this case v 1 ).stable focusunstable focusFigure 4.18: Focus equilibrium occurs when the eigenvalues are complex-conjugate,e.g., λ = −3 ± i (stable) or λ = +3 ± i (unstable). The imag<strong>in</strong>ary part (here 1)determ<strong>in</strong>es the frequency of rotation around the focus.


Two-Dimensional <strong>Systems</strong> 1110.60.5n-nullcl<strong>in</strong>e0.4K + activation variable, n0.30.2V-nullcl<strong>in</strong>e0.10-0.1-80 -70 -60 -50 -40 -30 -20 -10 0 10 20membrane voltage, V (mV)Figure 4.19: Phase portrait of the I Na,p +I K -model hav<strong>in</strong>g high-threshold K + current.n < 0, we cannot <strong>in</strong>terpret the result. As an exercise, prove that if all gat<strong>in</strong>g variablesof a model are <strong>in</strong>itially <strong>in</strong> the range [0, 1], then they stay <strong>in</strong> the range for all t ≥ 0.4.2.6 Example: FitzHugh-Nagumo modelThe FitzHugh-Nagumo model (FitzHugh 1961, Nagumo et al. 1962)˙V = V (a − V )(V − 1) − w + I , (4.11)ẇ = bV − cw , (4.12)imitates generation of action potentials by Hodgk<strong>in</strong>-Huxley-type models hav<strong>in</strong>g cubic(N-shaped) nullcl<strong>in</strong>es as <strong>in</strong> Fig. 4.4. Here V mimics the membrane voltage and the“recovery” variable w mimics activation of an outward current. Parameter I mimicsthe <strong>in</strong>jected current, and for the sake of simplicity we set I = 0 <strong>in</strong> our analysis below.Parameter a describes the shape of the cubic parabola V (a−V )(V −1), and parametersb > 0 and c ≥ 0 describe the k<strong>in</strong>etics of the recovery variable w. When b and c aresmall, the model may exhibit relaxation oscillations.The nullcl<strong>in</strong>es of the FitzHugh-Nagumo model have the cubic and l<strong>in</strong>ear formw = V (a − V )(V − 1) + Iw = b/c V(V -nullcl<strong>in</strong>e),(w-nullcl<strong>in</strong>e),


112 Two-Dimensional <strong>Systems</strong>V-nullcl<strong>in</strong>ew-nullcl<strong>in</strong>erecovery, w0.2w-nullcl<strong>in</strong>eV-nullcl<strong>in</strong>e00 0.5 1 0 0.5 1membrane voltage, Vmembrane voltage, VabFigure 4.20: Nullcl<strong>in</strong>es <strong>in</strong> the FitzHugh-Nagumo model (4.11, 4.12). Parameters: I =0, b = 0.01, c = 0.02, a = 0.1 (left) and a = −0.1 (right).?ctr L=-a-c = 0stability tr L < 0det L > 0=det L=ac+b0b0baFigure 4.21: Stability diagram of theequilibrium (0, 0) <strong>in</strong> the FitzHugh-Nagumo model (4.11,4.12).and they can <strong>in</strong>tersect <strong>in</strong> one, two, or three po<strong>in</strong>ts result<strong>in</strong>g <strong>in</strong> one, two, or threeequilibria, all of which may be unstable. Below we consider the simple case I = 0, sothat the orig<strong>in</strong>, (0, 0), is an equilibrium. Indeed, the nullcl<strong>in</strong>es of the model, depicted<strong>in</strong> Fig. 4.20, always <strong>in</strong>tersect at (0, 0) <strong>in</strong> this case. The <strong>in</strong>tersection may occur on theleft (Fig. 4.20a) or middle (Fig. 4.20b) branch of the cubic V -nullcl<strong>in</strong>e depend<strong>in</strong>g onthe sign of the parameter a. Let us determ<strong>in</strong>e how the stability of the equilibrium(0, 0) depends on the parameters a, b, and c.There is a common dogma that the equilibrium <strong>in</strong> Fig. 4.20a correspond<strong>in</strong>g toa > 0 is always stable, the equilibrium <strong>in</strong> Fig. 4.20b correspond<strong>in</strong>g to a < 0 is alwaysunstable, and the loss of stability occurs “exactly” at a = 0, i.e., at the bottom of theleft knee. Let us check that this is not necessarily true, at least when c ≠ 0. TheJacobian matrix of the FitzHugh-Nagumo model (4.11,4.12) at the equilibrium (0, 0)has the form( ) −a −1L =.b −c


Two-Dimensional <strong>Systems</strong> 113It is easy to check thatτ = tr L = −a − c and ∆ = det L = ac + b .Us<strong>in</strong>g Fig. 4.15 we conclude that the equilibrium is stable when tr L < 0 and det L >0, which corresponds to the shaded region <strong>in</strong> Fig. 4.21. Both conditions are alwayssatisfied when a > 0, hence the equilibrium <strong>in</strong> Fig. 4.20a is <strong>in</strong>deed stable. However,both conditions may also be satisfied for negative a, therefore, the equilibrium <strong>in</strong>Fig. 4.20b may also be stable. Thus, the equilibrium loses stability not at the leftknee, but slightly to the right of it, so that a part of the “unstable branch” of thecubic nullcl<strong>in</strong>e is actually stable. The part is small when b and c are small, i.e., when(4.11,4.12) is <strong>in</strong> a relaxation regime.4.3 Phase PortraitsAn important step <strong>in</strong> geometrical analysis of dynamical systems is sketch<strong>in</strong>g of theirphase portraits. The phase portrait of a two-dimensional system is a partition<strong>in</strong>g of thephase plane <strong>in</strong>to orbits or trajectories. Instead of depict<strong>in</strong>g all possible trajectories, itusually suffices to depict some representative trajectories. The phase portrait conta<strong>in</strong>sall important <strong>in</strong>formation about qualitative behavior of the dynamical system, suchas relative location and stability of equilibria, their attraction doma<strong>in</strong>s, separatrices,limit cycles, and other special trajectories that are discussed <strong>in</strong> this section.4.3.1 Bistability and attraction doma<strong>in</strong>sNon-l<strong>in</strong>ear two-dimensional systems can have many co-exist<strong>in</strong>g attractors. For example,the FitzHugh-Nagumo model (4.11,4.12) with nullcl<strong>in</strong>es depicted <strong>in</strong> Fig. 4.22 hastwo stable equilibria separated by an unstable equilibrium. Such a system is calledbistable (multi-stable when there are more than two attractors). Depend<strong>in</strong>g on the <strong>in</strong>itialconditions, the trajectory may approach the left or right equilibrium. The shadedarea denotes the attraction doma<strong>in</strong> of the right equilibrium; that is, the set of all<strong>in</strong>itial conditions that lead to this equilibrium. S<strong>in</strong>ce there are only two attractors,the complementary white area denotes the attraction doma<strong>in</strong> of the other equilibrium.The doma<strong>in</strong>s are separated not by equilibria as <strong>in</strong> one-dimensional case, but by specialtrajectories called separatrices, which we discuss <strong>in</strong> Sect. 4.3.2.Many neural models are bistable or can be made bistable when the parameters haveappropriate values. Often bistability results from the co-existence of an equilibriumattractor correspond<strong>in</strong>g to the rest state and a limit cycle attractor correspond<strong>in</strong>g tothe repetitive fir<strong>in</strong>g state. Fig. 4.23 depicts one of many possible cases. Here we usethe I Na,p +I K -model with a high-threshold fast K + current. The rest state exists due tothe balance of partially activated Na + and leak currents. The repetitive spik<strong>in</strong>g statepersists because the K + current deactivates too fast and cannot br<strong>in</strong>g the membranepotential <strong>in</strong>to the subthreshold voltage range. If the <strong>in</strong>itial state is <strong>in</strong> the shaded area,


114 Two-Dimensional <strong>Systems</strong>V-nullcl<strong>in</strong>e0.2recovery, wseparatrixw-nullcl<strong>in</strong>e0separatrixattraction doma<strong>in</strong>-0.4 -0.2 0 0.2 0.4 0.6 0.8 1membrane voltage, VFigure 4.22: Bistability of two equilibrium attractors (black circles) <strong>in</strong> the FitzHugh-Nagumo model (4.11,4.12). The shaded area — attraction doma<strong>in</strong> of the right equilibrium.Parameters: I = 0, b = 0.01, a = c = 0.1.which is the attraction doma<strong>in</strong> of the limit cycle attractor, the trajectory approachesthe limit cycle attractor and the neuron fires an <strong>in</strong>f<strong>in</strong>ite tra<strong>in</strong> of action potentials.4.3.2 Stable/unstable manifoldsIn contrast to one-dimensional systems, <strong>in</strong> two-dimensional systems unstable equilibriado not necessarily separate attraction doma<strong>in</strong>s. Nevertheless, they play an importantrole <strong>in</strong> def<strong>in</strong><strong>in</strong>g the boundary of attraction doma<strong>in</strong>s, as <strong>in</strong> Fig. 4.22 and Fig. 4.23.In both cases the attraction doma<strong>in</strong>s are separated by a pair of trajectories, calledseparatrices, which converge to the saddle equilibrium. Such trajectories form thestable manifold of a saddle po<strong>in</strong>t. Locally, the manifold is parallel to the eigenvectorcorrespond<strong>in</strong>g to the negative (stable) eigenvalue; see Fig. 4.24. Similarly, the unstablemanifold of a saddle is formed by the two trajectories that orig<strong>in</strong>ate exactly from thesaddle (or approach the saddle if the time is reversed). Locally, the unstable manifoldis parallel to the eigenvector correspond<strong>in</strong>g to the positive (unstable) eigenvalue.The stable manifold of the saddle <strong>in</strong> Fig. 4.23 plays the role of a threshold, s<strong>in</strong>ce itseparates rest and spik<strong>in</strong>g states. We illustrate this concept <strong>in</strong> Fig. 4.24: If the <strong>in</strong>itialstate of the system, denoted as A, is <strong>in</strong> the shaded area, the trajectory will convergeto the spik<strong>in</strong>g attractor (right) no matter how close the <strong>in</strong>itial condition to the stablemanifold is. In contrast, if the <strong>in</strong>itial condition, denoted as B, is <strong>in</strong> the white area, the


Two-Dimensional <strong>Systems</strong> 1150.60.5n-nullcl<strong>in</strong>eK + activation variable, n0.40.30.20.10restseparatrixattractiondoma<strong>in</strong>separatrixV-nullcl<strong>in</strong>e-80 -70 -60 -50 -40 -30 -20 -10 0 10 20membrane voltage, V (mV)membrane voltage, V (mV)-10restspik<strong>in</strong>g-700 5 10 15time, t (ms)Figure 4.23: Bistability of rest and spik<strong>in</strong>g states <strong>in</strong> the I Na,p +I K -model (4.1, 4.2) withhigh-threshold fast (τ(V ) = 0.152) K + current and I = 3. A brief strong pulse ofcurrent (arrow) br<strong>in</strong>gs the state vector of the system <strong>in</strong>to the attraction doma<strong>in</strong> of thestable limit cycle.to restunstable manifoldv 2stable manifold (separatrix)v 1AB C ACBto spik<strong>in</strong>gunstable manifoldmembrane voltageseparatrixto spik<strong>in</strong>gto saddletimeto restFigure 4.24: Stable and unstable manifolds to a saddle. The eigenvectors v 1 and v 2correspond to positive and negative eigenvalues, respectively.


116 Two-Dimensional <strong>Systems</strong>heterocl<strong>in</strong>ic orbithomocl<strong>in</strong>icorbitFigure 4.25: A heterocl<strong>in</strong>ic orbit starts andends at different equilibria. A homocl<strong>in</strong>ic orbitstarts and ends at the same equilibrium.trajectory will converge to the rest attractor (left). If the <strong>in</strong>itial condition is preciselyon the stable manifold (po<strong>in</strong>t C), the trajectory converges neither to rest nor to spik<strong>in</strong>gstate, but to the saddle equilibrium. Of course, this case is highly unstable and smallperturbations will certa<strong>in</strong>ly push the trajectory to one or the other side. The importantmessage <strong>in</strong> Fig. 4.24 is that a threshold is not a po<strong>in</strong>t, i.e., a s<strong>in</strong>gle voltage value, buta trajectory on the phase plane. (F<strong>in</strong>d an exceptional case when the threshold lookslike a s<strong>in</strong>gle voltage value. H<strong>in</strong>t: see Fig. 4.17.)4.3.3 Homocl<strong>in</strong>ic/heterocl<strong>in</strong>ic trajectoriesFig. 4.24 shows that trajectories form<strong>in</strong>g the unstable manifold orig<strong>in</strong>ate from thesaddle. Where do they go? Similarly, the trajectories form<strong>in</strong>g the stable manifoldterm<strong>in</strong>ate at the saddle. Where do they come from? We say that a trajectory isheterocl<strong>in</strong>ic if it orig<strong>in</strong>ates at one equilibrium and term<strong>in</strong>ates at another equilibrium,as <strong>in</strong> Fig. 4.25. A trajectory is homocl<strong>in</strong>ic if it orig<strong>in</strong>ates and term<strong>in</strong>ates at the sameequilibrium. These types of trajectories play an important role <strong>in</strong> geometrical analysisof dynamical systems.Heterocl<strong>in</strong>ic trajectories connect unstable and stable equilibria, as <strong>in</strong> Fig. 4.26, andthey are ubiquitous <strong>in</strong> dynamical systems hav<strong>in</strong>g two or more equilibrium po<strong>in</strong>ts. Infact, there are <strong>in</strong>f<strong>in</strong>itely many heterocl<strong>in</strong>ic trajectories <strong>in</strong> Fig. 4.26, s<strong>in</strong>ce all trajectories<strong>in</strong>side the bold loop orig<strong>in</strong>ate at the unstable focus and term<strong>in</strong>ate at the stable node.(F<strong>in</strong>d the exceptional trajectory that ends elsewhere.)In contrast, homocl<strong>in</strong>ic trajectories are rare. First, a homocl<strong>in</strong>ic trajectory divergesfrom an equilibrium, therefore the equilibrium must be unstable. Next, the trajectorymakes a loop and returns to the same equilibrium, as <strong>in</strong> Fig. 4.27. It needs to hitthe unstable equilibrium precisely, s<strong>in</strong>ce a small error would make it deviate from theunstable equilibrium. Though uncommon, homocl<strong>in</strong>ic trajectories <strong>in</strong>dicate that thesystem undergoes a bifurcation — appearance or disappearance of a limit cycle. Thehomocl<strong>in</strong>ic trajectory <strong>in</strong> Fig. 4.27 <strong>in</strong>dicates that the limit cycle <strong>in</strong> Fig. 4.23 is aboutto (dis)appear via saddle homocl<strong>in</strong>ic orbit bifurcation. The homocl<strong>in</strong>ic trajectory <strong>in</strong>Fig. 4.28 <strong>in</strong>dicates that a limit cycle is about to (dis)appear via saddle-node on <strong>in</strong>variantcircle bifurcation. We study these bifurcations <strong>in</strong> detail <strong>in</strong> Chap. 6.4.3.4 Saddle-node bifurcationIn Fig. 4.29 we simulate the <strong>in</strong>jection of a ramp current I <strong>in</strong>to the I Na,p +I K -modelhav<strong>in</strong>g high-threshold K + current. Our goal is to understand the transition from the


Two-Dimensional <strong>Systems</strong> 117rest state to repetitive spik<strong>in</strong>g. When I is small, the phase portrait of the model issimilar to the one depicted <strong>in</strong> Fig. 4.26 for I = 0. There are two equilibria <strong>in</strong> thelow-voltage range — a stable node correspond<strong>in</strong>g to the rest state and a saddle. Theequilibria are the <strong>in</strong>tersections of the cubic V -nullcl<strong>in</strong>e and the n-nullcl<strong>in</strong>e. Increas<strong>in</strong>gthe parameter I changes the shape of the cubic nullcl<strong>in</strong>e and shifts it upward, but doesnot change the n-nullcl<strong>in</strong>e. As a result, the distance between the equilibria decreases,until they coalesce as <strong>in</strong> Fig. 4.28 so that the nullcl<strong>in</strong>es only touch each other <strong>in</strong> the lowvoltagerange. Further <strong>in</strong>crease of I results <strong>in</strong> the disappearance of the saddle and nodeequilibrium, and hence <strong>in</strong> the disappearance of the rest state. The new phase portraitis depicted <strong>in</strong> Fig. 4.30; it has only a limit cycle attractor correspond<strong>in</strong>g to repetitivefir<strong>in</strong>g. We see that <strong>in</strong>creas<strong>in</strong>g I past the value I = 4.51 results <strong>in</strong> transition from restto periodic spik<strong>in</strong>g dynamics. What k<strong>in</strong>d of a bifurcation occurs when I = 4.51?Those readers who did not skip Sect. 3.3.3 <strong>in</strong> the previous chapter will immediatelyrecognize the saddle-node bifurcation, whose major stages are summarized <strong>in</strong> Fig. 4.31.As a bifurcation parameter changes, the saddle and the node equilibrium approach eachother, coalesce, and then annihilate each other so there are no equilibria left. Whencoalescent, the jo<strong>in</strong>t equilibrium is neither a saddle nor a node, but a saddle-node. Itsmajor feature is that it has precisely one zero eigenvalue, and it is stable on one side ofthe neighborhood and unstable on the other side. In Chap. 6 we will provide an exactdef<strong>in</strong>ition of a saddle-node bifurcation <strong>in</strong> a multi-dimensional system, and we will showthat there are two important subtypes of this bifurcation, result<strong>in</strong>g <strong>in</strong> slightly different0.60.5n-nullcl<strong>in</strong>e0.4K + activation variable, n0.30.2V-nullcl<strong>in</strong>e0.10heterocl<strong>in</strong>ic orbitheterocl<strong>in</strong>ic orbit-80 -70 -60 -50 -40 -30 -20 -10 0 10 20membrane voltage, V (mV)Figure 4.26: Two heterocl<strong>in</strong>ic orbits (bold curves connect<strong>in</strong>g stable and unstable equilibria)<strong>in</strong> the I Na,p +I K -model with high-threshold K + current.


118 Two-Dimensional <strong>Systems</strong>0.60.5n-nullcl<strong>in</strong>eK + activation variable, n0.40.30.20.1homocl<strong>in</strong>ic orbitV-nullcl<strong>in</strong>e0-80 -70 -60 -50 -40 -30 -20 -10 0 10 20membrane voltage, V (mV)Figure 4.27: Homocl<strong>in</strong>ic orbit (bold) <strong>in</strong> the I Na,p +I K -model with high-threshold fast(τ(V ) = 0.152) K + current.0.60.5n-nullcl<strong>in</strong>e0.4K + activation variable, n0.30.2V-nullcl<strong>in</strong>e0.1homocl<strong>in</strong>ic orbit0-80 -70 -60 -50 -40 -30 -20 -10 0 10 20membrane voltage, V (mV)Figure 4.28: Homocl<strong>in</strong>ic orbit (bold) to saddle-node equilibrium <strong>in</strong> the I Na,p +I K -modelwith high-threshold K + current and I = 4.51.


Two-Dimensional <strong>Systems</strong> 119membrane voltage, V (mV)<strong>in</strong>jected current, I200-20-40-60-80105resttransition(bifurcation)I=4.51spik<strong>in</strong>g00 50 100 150 200 250 300time (ms)Figure 4.29: Transition from rest state to repetitive spik<strong>in</strong>g <strong>in</strong> the I Na,p +I K -model with<strong>in</strong>jected ramp current I; see also Fig. 4.26, Fig. 4.28, and Fig. 4.30. Notice that thefrequency of spik<strong>in</strong>g is <strong>in</strong>itially small, and then it <strong>in</strong>creases as the amplitude of the<strong>in</strong>jected current <strong>in</strong>creases.0.60.5n-nullcl<strong>in</strong>e0.4K + activation variable, n0.30.2V-nullcl<strong>in</strong>e0.10limit cycle attractor-80 -70 -60 -50 -40 -30 -20 -10 0 10 20membrane voltage, V (mV)Figure 4.30: Limit cycle attractor (bold) <strong>in</strong> the I Na,p +I K -model when I = 10 (comparewith Fig. 4.26 and Fig. 4.28).


120 Two-Dimensional <strong>Systems</strong>eigenvalues:λ 1


Two-Dimensional <strong>Systems</strong> 121-20-50-30sa dle (threshold state)membrane voltage, V (mV)-40-50-60-70membrane voltage, V (mV)-60-70stable node (rest state)saddle-nodebifurcation-100 -50 0 50 100<strong>in</strong>jected dc-current I4.51-10 -5 0 5 10<strong>in</strong>jected dc-current IFigure 4.32: Saddle-node bifurcation diagram of the I Na,p +I K -model.given by the equation (4.13).The curve ismembrane voltage, V (mV)<strong>in</strong>jected current, I0-20-40-60-80302010resttransition(bifurcation)I=12spik<strong>in</strong>g00 10 20 30 40 50 60 70 80 90 100time (ms)Figure 4.33: Transition from rest state to repetitive spik<strong>in</strong>g <strong>in</strong> the I Na,p +I K -modelwith ramp <strong>in</strong>jected current I; see also Fig. 4.34 (small-amplitude noise is added tothe model to mask the slow passage effect). Notice that the frequency of spik<strong>in</strong>g isrelatively constant for a wide range of <strong>in</strong>jected current.


122 Two-Dimensional <strong>Systems</strong>K + activation variable, n0.70.60.50.40.30.20.1n-nullcl<strong>in</strong>e0.7I=00.6I=12V-nullcl<strong>in</strong>eK + activation variable, n0.50.40.30.20.1K + activation variable, n00.70.60.50.40.30.20.10-80 -60 -40 -20 0 20membrane voltage, V (mV)0.50.40.30.20.1-80 -60 -40 -20 0 20membrane voltage, V (mV)0.7I=20 I=400.6K + activation variable, nlimit cycle0-80 -60 -40 -20 0 20membrane voltage, V (mV)0-80 -60 -40 -20 0 20membrane voltage, V (mV)Figure 4.34: Supercritical Andronov-Hopf bifurcation <strong>in</strong> the I Na,p +I K -model (4.1, 4.2)with low-threshold K + current when I = 12; see also Fig. 4.33.4.3.5 Andronov-Hopf bifurcationIn Fig. 4.33 we repeat the current ramp experiment us<strong>in</strong>g the I Na,p +I K -model with lowthresholdK + current. The phase portrait of such a model is simple — it has a uniqueequilibrium, as we illustrate <strong>in</strong> Fig. 4.34. When I is small, the equilibrium is a stablefocus correspond<strong>in</strong>g to the rest state. When I <strong>in</strong>creases past I = 12, the focus losesstability and gives birth to a small-amplitude limit cycle attractor. The amplitude ofthe limit cycle grows as I <strong>in</strong>creases. We see that <strong>in</strong>creas<strong>in</strong>g I beyond I = 12 results <strong>in</strong>the transition from rest to spik<strong>in</strong>g behavior. What k<strong>in</strong>d of a bifurcation occurs there?Recall that stable foci have a pair of complex-conjugate eigenvalues with negativereal part. When I <strong>in</strong>creases, the real part of the eigenvalues also <strong>in</strong>creases until itbecomes zero (at I = 12) and then positive (when I > 12) mean<strong>in</strong>g that the focus isno longer stable. The transition from stable to unstable focus described above is calledAndronov-Hopf bifurcation. It occurs when the eigenvalues become purely imag<strong>in</strong>ary,as it happens when I = 12. We will study Andronov-Hopf bifurcations <strong>in</strong> detail <strong>in</strong>Chap. 6, where we will show that they can be supercritical or subcritical. The formercorrespond to birth of a small-amplitude limit cycle attractor, as <strong>in</strong> Fig. 4.34. Thelatter correspond to a death of an unstable limit cycle.


Two-Dimensional <strong>Systems</strong> 1230-100-10max V(t)membrane voltage, V-20-30-40-50-60-70stableAndronov-Hopf bifurcationis somewhere hereunstableI (V)membrane voltage, V-20-30-40-50-60-70supercriticalAndronov-Hopfbifurcationrestm<strong>in</strong> V(t)periodic orbitsunstable equilibrium-800 20 40 60 80 100<strong>in</strong>jected dc-current, I-800 20 40 60 80 100<strong>in</strong>jected dc-current, IabFigure 4.35: Andronov-Hopf bifurcation diagram <strong>in</strong> the I Na,p +I K -model with lowthresholdK + current. a. Equilibria of the model (solution of (4.13)). b. Equilibriaand limit cycles of the model.In Fig. 4.35a we plot the solution of (4.13) as an attempt to determ<strong>in</strong>e the bifurcationdiagram for the Andronov-Hopf bifurcation <strong>in</strong> the I Na,p +I K -model. However,all we can see is that the equilibrium persists as I <strong>in</strong>creases, but there is no <strong>in</strong>formationon its stability or on the existence of a limit cycle attractor. To study the limitcycle attractor, we need to simulate the model with various values of parameter I.For each I, we disregard the transient period and plot m<strong>in</strong> V (t) and max V (t) on the(I, V )-plane, as <strong>in</strong> Fig. 4.35b. When I is small, the solutions converge to the stableequilibrium, and both m<strong>in</strong> V (t) and max V (t) are equal to the rest<strong>in</strong>g voltage. WhenI <strong>in</strong>creases past I = 12, the m<strong>in</strong> V (t) and max V (t) values start to diverge, mean<strong>in</strong>gthat there is a limit cycle attractor whose amplitude <strong>in</strong>creases as I does. This methodis appropriate for analysis of supercritical Andronov-Hopf bifurcations but it fails forsubcritical Andronov-Hopf bifurcations. Why?Figure 4.36 depicts an <strong>in</strong>terest<strong>in</strong>g phenomenon observed <strong>in</strong> many biological neurons— excitation block. Spik<strong>in</strong>g activity of the layer 5 pyramidal neuron of rat’s visualcortex is blocked by strong excitation, i.e., <strong>in</strong>jection of strong depolariz<strong>in</strong>g current. Thegeometry of this phenomenon is illustrated <strong>in</strong> Fig. 4.37, bottom. As the magnitude ofthe <strong>in</strong>jected current <strong>in</strong>creases, the unstable equilibrium, which is the <strong>in</strong>tersection po<strong>in</strong>tof the nullcl<strong>in</strong>es, moves to right branch of the cubic V -nullcl<strong>in</strong>e and becomes stable.The limit cycle shr<strong>in</strong>ks and the spik<strong>in</strong>g activity disappears, typically but not necessarilyvia supercritical Andronov-Hopf type. Thus, the I Na,p +I K -model with low-thresholdK + current can exhibit two such bifurcations <strong>in</strong> response to ramp<strong>in</strong>g up of the <strong>in</strong>jectedcurrent, one lead<strong>in</strong>g to the appearance of periodic spik<strong>in</strong>g activity (Fig. 4.34), and thenone lead<strong>in</strong>g to its disappearance (Fig. 4.37).


124 Two-Dimensional <strong>Systems</strong>excitation blocksupercriticalAndronov-Hopfbifurcation20 mV100 ms-60 mV<strong>in</strong>jected current-10 mV180 pAFigure 4.36: Excitation block <strong>in</strong> layer 5 pyramidal neuron of rat’s visual cortex as theamplitude of the <strong>in</strong>jected current ramps up.Supercritical and subcritical Andronov-Hopf bifurcations <strong>in</strong> neurons result <strong>in</strong> slightlydifferent neuro-computational properties. In contrast, the saddle-node and Andronov-Hopf bifurcations result <strong>in</strong> dramatically different neuro-computational properties. Inparticular, neurons near saddle-node bifurcation act as <strong>in</strong>tegrators — they prefer highfrequency<strong>in</strong>put: The higher the frequency of the <strong>in</strong>put, the sooner they fire. Incontrast, neural systems near Andronov-Hopf bifurcation have damped oscillatory potentialsand they act as resonators — they prefer oscillatory <strong>in</strong>put with the samefrequency as that of damped oscillations. Increas<strong>in</strong>g the frequency may delay or eventerm<strong>in</strong>ate their response. We discuss this and other neuro-computational properties <strong>in</strong>Chap. 7.


Two-Dimensional <strong>Systems</strong> 125membrane potential, V (mV)0-10-20-30-40-50-60-70excitation blocksupercriticalAndronov-Hopfbifurcation<strong>in</strong>jectedcurrentI=40I=150I=300I=4000 10 20 30 40 50 60 70 80 90 100time, msK + activation variable, n10.80.60.40.2limit cycleI=40 I=15010.80.60.40.2K + activation variable, n0-80 -60 -40 -20 0 20membrane potential, V (mV)0-80 -60 -40 -20 0 20membrane potential, V (mV)K + activation variable, n10.80.60.40.2I=300 I=40010.80.60.40.2K + activation variable, nexcitation block0-80 -60 -40 -20 0 20membrane potential, V (mV)0-80 -60 -40 -20 0 20membrane potential, V (mV)Figure 4.37: Excitation block <strong>in</strong> the I Na,p +I K -model: As the magnitude of the <strong>in</strong>jectedcurrent I ramps up, the spik<strong>in</strong>g stops.


126 Two-Dimensional <strong>Systems</strong>Review of Important Concepts• A two-dimensional system of differential equationsẋ = f(x, y)ẏ = g(x, y) ,describes jo<strong>in</strong>t evolution of state variables x and y, which often are the membranevoltage and a “recovery” variable.• Solutions of the system are trajectories on the phase plane R 2 that are tangentto the vector field (f, g).• The sets given by the equations f(x, y) = 0 and g(x, y) = 0 are the x- andy-nullcl<strong>in</strong>es, respectively, where trajectories change their x and y directions.• Intersections of the nullcl<strong>in</strong>es are equilibria of the system.• Periodic dynamics correspond to closed loop trajectories.• Some special trajectories, e.g., separatrices, def<strong>in</strong>e thresholds and separateattraction doma<strong>in</strong>s.• An equilibrium or a periodic trajectory is stable if all nearby trajectories areattracted to it.• To determ<strong>in</strong>e the stability of an equilibrium, one needs to consider the Jacobianmatrix of partial derivatives( )fx fL =y.g x g y• The equilibrium is stable when both eigenvalues of L are negative or havenegative real parts.• The equilibrium is a saddle, a node, or a focus, when L has real eigenvaluesof opposite signs, of the same signs, or complex-conjugate eigenvalues,respectively.• When the equilibrium undergoes a saddle-node bifurcation, one of the eigenvaluesbecomes zero.• When the equilibrium undergoes an Andronov-Hopf bifurcation (birth or deathof a small periodic trajectory) the complex-conjugate eigenvalues becomepurely imag<strong>in</strong>ary.• The saddle-node and Andronov-Hopf bifurcations are ubiquitous <strong>in</strong> neuralmodels, and they result <strong>in</strong> different neuro-computational properties.


Two-Dimensional <strong>Systems</strong> 127Bibliographical NotesAmong many textbooks on the mathematical theory of dynamical systems we recommendthe follow<strong>in</strong>g three:• Nonl<strong>in</strong>ear Dynamics and Chaos by Strogatz (1994) is suitable as an <strong>in</strong>troductorybook for undergraduate math or physics majors or graduate students <strong>in</strong> lifesciences. It conta<strong>in</strong>s many exercises and worked out examples.• Differential Equations and <strong>Dynamical</strong> <strong>Systems</strong> by Perko (1996, Third edition<strong>in</strong> 2000) is suitable for math and physics graduate students, but may be tootechnical for life scientists. Nevertheless, it should be a standard textbook forcomputational neuroscientists.• Elements of Applied Bifurcation Theory by Kuznetsov (1995, Third edition <strong>in</strong>2004) is suitable for advanced graduate students <strong>in</strong> mathematics or physics andfor computational neuroscientists who want to pursue bifurcation analysis of neuralmodels.The second edition of The Geometry of Biological Time by W<strong>in</strong>free (2001) is a good<strong>in</strong>troductory book <strong>in</strong>to oscillations, limit cycles, and synchronization <strong>in</strong> biology. Itrequires little background <strong>in</strong> mathematics and can be suitable even for undergraduatelife science majors. Mathematical Biology by Murray (1993, Third edition <strong>in</strong> 2003) is anexcellent example how dynamical system theory can solve many problems <strong>in</strong> populationbiology and shed light on pattern formation <strong>in</strong> biological systems. Most of this book issuitable for advance undergraduate or graduate students <strong>in</strong> mathematics and physics.Mathematical Physiology by Keener and Sneyd (1998) is similar to Murray’s book, butis more focused on neural systems. Spikes, Decisions, and Actions by Wilson (1999)is a short <strong>in</strong>troduction to dynamical systems with many neuroscience examples.Exercises1. Use pencil (as <strong>in</strong> Fig. 4.39) to sketch the nullcl<strong>in</strong>es of the vector fields depicted<strong>in</strong> figures 4.40 through 4.44.2. Assume that the cont<strong>in</strong>uous curve is the x-nullcl<strong>in</strong>e and the dashed curve is they-nullcl<strong>in</strong>e <strong>in</strong> Fig. 4.38, and that ẋ or ẏ changes sign when (x, y) passes throughthe correspond<strong>in</strong>g nullcl<strong>in</strong>e. The arrow <strong>in</strong>dicates the direction of the vector field<strong>in</strong> one region. Determ<strong>in</strong>e the approximate directions of the vector field <strong>in</strong> theother regions of the phase plane.3. Use pencil (as <strong>in</strong> Fig. 4.39) to sketch phase portraits of the vector fields depicted<strong>in</strong> figures 4.40 through 4.44. Clearly mark all equilibria, their stability, andattraction doma<strong>in</strong>s. Show directions of all homocl<strong>in</strong>ic, heterocl<strong>in</strong>ic and periodic


128 Two-Dimensional <strong>Systems</strong>abcdFigure 4.38: Determ<strong>in</strong>e the approximate direction of the vector field <strong>in</strong> each regionbetween the nullcl<strong>in</strong>es. Cont<strong>in</strong>uous (dashed) curve is the x-nullcl<strong>in</strong>e (y-nullcl<strong>in</strong>e), andthe direction of vector field <strong>in</strong> one region is <strong>in</strong>dicated by the arrow.trajectories, as well as other representative trajectories. Estimate the signs ofeigenvalues at each equilibrium.4. Prove the classification diagram <strong>in</strong> Fig. 4.15.5. (van der Pol oscillator) Determ<strong>in</strong>e nullcl<strong>in</strong>es and draw phase portrait of the vander Pol oscillator (given <strong>in</strong> the Liénard (1928) form)where b > 0 is a parameter.ẋ = x − x 3 /3 − y ,ẏ = bx ,6. (Bonhoeffer–van der Pol oscillator) Determ<strong>in</strong>e the nullcl<strong>in</strong>es and sketch representativephase portraits of the Bonhoeffer–van der Pol oscillatorẋ = x − x 3 /3 − y ,ẏ = b(x − a) − cy ,<strong>in</strong> the case of c = 0. Treat a and b > 0 as parameters.7. (H<strong>in</strong>dmarsh-Rose spik<strong>in</strong>g neuron) The follow<strong>in</strong>g system is a generalization of theFitzHugh-Nagumo model (H<strong>in</strong>dmarsh and Rose 1982)ẋ = f(x) − y + I ,ẏ = g(x) − y ,


Two-Dimensional <strong>Systems</strong> 129Figure 4.39: Phase portrait of a vector field. Use pencil to draw phase portraits <strong>in</strong>figures 4.40 through 4.44.Figure 4.40: Use pencil to draw a phase portrait as <strong>in</strong> Fig. 4.39.


130 Two-Dimensional <strong>Systems</strong>Figure 4.41: Use pencil to draw a phase portrait as <strong>in</strong> Fig. 4.39.Figure 4.42: Use pencil to draw a phase portrait as <strong>in</strong> Fig. 4.39.Figure 4.43: Use pencil to draw a phase portrait as <strong>in</strong> Fig. 4.39.


Two-Dimensional <strong>Systems</strong> 131Figure 4.44: Use pencil to draw a phase portrait as <strong>in</strong> Fig. 4.39.where f(x) = −ax 3 + bx 2 , g(x) = −c + dx 2 , and a, b, c, d, and I are parameters.Suppose (¯x, ȳ) is an equilibrium. Determ<strong>in</strong>e its type and stability as a functionof f ′ = f ′ (¯x) and g ′ = g(¯x); that is, plot a diagram similar to the one <strong>in</strong> Fig. 4.15with f ′ and g ′ as coord<strong>in</strong>ates.8. (I K -model) Show that the unique equilibrium <strong>in</strong> the I K -modelC ˙V = −g L (V − E L ) − g K m 4 (V − E K ) , (4.14)ṁ = (m ∞ (V ) − m)/τ(V ) . (4.15)discussed <strong>in</strong> the previous chapter (see Fig. 3.40) is always stable, at least whenE L > E K . (H<strong>in</strong>t: look at the signs of the trace and determ<strong>in</strong>ant of the Jacobianmatrix).9. (I h -model) Show that the unique equilibrium <strong>in</strong> the full I h -modelC ˙V = −g L (V − E L ) − g h h(V − E h ) ,ḣ = (h ∞ (V ) − h)/τ(V ) ,discussed <strong>in</strong> the previous chapter is always stable.10. (Bendixson’s criterion) If the divergence of the vector field∂f(x, y)∂x+∂g(x, y)∂yof a two-dimensional dynamical system is not identically zero and does not changesign on the plane, then the dynamical system cannot have limit cycles. Use thiscriterion to show that the I K -model and the I h -model cannot oscillate.11. Determ<strong>in</strong>e stability of equilibria <strong>in</strong> the modelẋ = a + x 2 − y ,ẏ = bx − cy ,where a ∈ R, b ≥ 0 and c > 0 are some parameters.


132 Two-Dimensional <strong>Systems</strong>


Chapter 5Conductance-Based Models andTheir ReductionsIn this chapter we present examples of geometrical phase plane analysis of varioustwo-dimensional neural models. In particular, we consider m<strong>in</strong>imal models, i.e., thosehav<strong>in</strong>g m<strong>in</strong>imal sets of currents that enable the models to generate action potentials.The remarkable fact is that all these models can be reduced to planar systems hav<strong>in</strong>gN-shaped V -nullcl<strong>in</strong>es. We will see that the behavior of the models depends not somuch on the ionic currents as on the relationship between (<strong>in</strong>)activation curves and thetime constants. That is, models <strong>in</strong>volv<strong>in</strong>g completely different currents can have identicaldynamics, and conversely, models <strong>in</strong>volv<strong>in</strong>g similar currents can have completelydifferent dynamics.5.1 M<strong>in</strong>imal ModelsThere are a few dozens of known voltage- and Ca 2+ -gated currents hav<strong>in</strong>g diverseactivation and <strong>in</strong>activation dynamics, and this number grows every year. Some ofthem are summarized <strong>in</strong> Sect. 2.3.5. Almost any comb<strong>in</strong>ation of the currents wouldresult <strong>in</strong> <strong>in</strong>terest<strong>in</strong>g non-l<strong>in</strong>ear behavior, such as excitability. Therefore, there arebillions (more than 2 30 ) of different electrophysiological models of neurons. Here wesay that two models are “different” if for example one has the h-current I h and theother does not, without even consider<strong>in</strong>g how much of the I h is there. How can weclassify all such models?Let us do the follow<strong>in</strong>g thought experiment: Consider a conductance-based modelcapable of exhibit<strong>in</strong>g periodic spik<strong>in</strong>g, i.e., hav<strong>in</strong>g a limit cycle attractor. Let usremove completely a current or one of its gat<strong>in</strong>g variables, and ask the question “Doesthe reduced model have a limit cycle attractor, at least for some values of parameters?”If it does, we remove one more gat<strong>in</strong>g variable or current, and proceed until we arriveat a model that satisfies the follow<strong>in</strong>g two properties:• It has a limit cycle attractor, at least for some values of parameters.133


134 Conductance-Based Models• If one removes any current or gat<strong>in</strong>g variable, the model has only equilibriumattractors for any values of parameters.We refer to such a model as be<strong>in</strong>g m<strong>in</strong>imal or irreducible for spik<strong>in</strong>g. Thus, m<strong>in</strong>imalmodels can exhibit periodic activity, even if of small amplitude, but their reductionscannot. Accord<strong>in</strong>g to this def<strong>in</strong>ition, any space-clamped conductance-based model iseither a m<strong>in</strong>imal model, or could be reduced to a m<strong>in</strong>imal model or models by remov<strong>in</strong>ggat<strong>in</strong>g variables. This will be the basis for our classification of electrophysiologicalmechanisms <strong>in</strong> neurons.For example, the Hodgk<strong>in</strong>-Huxley model considered <strong>in</strong> Sect. 2.3 is not m<strong>in</strong>imal forspik<strong>in</strong>g. Recall that the model consists of three currents: leakage I L , transient sodiumI Na,t (gat<strong>in</strong>g variables m and h) and persistent potassium I K (gat<strong>in</strong>g variable n); seeFig. 5.1. Which of these currents are responsible for excitability and spik<strong>in</strong>g?We can remove the leakage current and the gat<strong>in</strong>g variable, h, for the <strong>in</strong>activationof the sodium current: The result<strong>in</strong>g I Na,p +I K -modelC ˙V = I −I K{ }} {g K n 4 (V − E K ) −ṅ = (n ∞ (V ) − n)/τ n (V ) ,ṁ = (m ∞ (V ) − m)/τ m (V ) ,I Na,p{ }} {g Na m 3 (V − E Na ) ,was considered <strong>in</strong> the previous chapter where we have shown that it could oscillate dueto the <strong>in</strong>terplay between the activation of persistent sodium and potassium currents.Alternatively, we can remove the K + current from the Hodgk<strong>in</strong>-Huxley model, yet thenew I Na,t -modelC ˙V = I −I Na,t{ }} {g Na m 3 h(V − E Na ) −ṁ = (m ∞ (V ) − m)/τ m (V ) ,ḣ = (h ∞ (V ) − h)/τ h (V ) ,I L{ }} {g L (V − E L ) ,can still oscillate via the <strong>in</strong>terplay between activation and <strong>in</strong>activation of the Na +current, as we will see later <strong>in</strong> this chapter. Both models are m<strong>in</strong>imal, because removalof any other gat<strong>in</strong>g variable results <strong>in</strong> either the I Na,p -, I K -, or I h -models, neither ofwhich can have a limit cycle attractor, as the reader is asked to prove at the end of theprevious chapter.We see that the Hodgk<strong>in</strong>-Huxley model is not m<strong>in</strong>imal, but it is a comb<strong>in</strong>ation oftwo m<strong>in</strong>imal models. M<strong>in</strong>imal models are appeal<strong>in</strong>g because they are relatively simple;each <strong>in</strong>dividual variable has an established electrophysiological mean<strong>in</strong>g, and its role<strong>in</strong> dynamics can be easily identified. As we show below, many m<strong>in</strong>imal models can


Conductance-Based Models 135Hodgk<strong>in</strong>-HuxleyTransient Na Current (m,h)Persistent K Current (n)Leak CurrentRemove nRemove hand LeakTransient Na Current (m,h)Leak CurrentPersistent Na Current (m)Persistent K Current (n)M<strong>in</strong>imal ModelshGat<strong>in</strong>g forInactivationof Na CurrentmGat<strong>in</strong>g forActivationof Na CurrentnGat<strong>in</strong>g forActivationof K CurrentLeak CurrentGat<strong>in</strong>g variablesFigure 5.1: The Hodgk<strong>in</strong>-Huxley model (top box) is a comb<strong>in</strong>ation of m<strong>in</strong>imal models(shaded boxes on second level). Each m<strong>in</strong>imal model can oscillate at least for somevalues of its parameters.be reduced to planar systems, which are amenable to analysis us<strong>in</strong>g geometrical phaseplane methods. In Sect. 5.2 we discuss other methods of reduction of multi-dimensionalmodels, e.g., the Hodgk<strong>in</strong>-Huxley model, to planar systems.There are only few m<strong>in</strong>imal models, and understand<strong>in</strong>g their dynamics can shedlight on dynamics of more complicated electrophysiological models. However, thereader should be aware of limitations of such an approach: Understand<strong>in</strong>g m<strong>in</strong>imalmodels cannot provide exhaustive <strong>in</strong>formation about all electrophysiological models(the same way as understand<strong>in</strong>g the zeros of the equations y = x, and y = x 2 does notprovide complete <strong>in</strong>formation about the zeros of the equation y = x + x 2 ).5.1.1 Amplify<strong>in</strong>g and resonant gat<strong>in</strong>g variablesThe def<strong>in</strong>ition of the m<strong>in</strong>imal models <strong>in</strong>volves a top-down approach: Take a complicatedmodel and strip it down to m<strong>in</strong>imal ones. It is unlikely that this could be done for all2 30 or so electrophysiological models. Instead, we employ here a bottom-up approach,which is based on the follow<strong>in</strong>g rule of thumb: A mixture of one amplify<strong>in</strong>g and oneresonant (recovery) gat<strong>in</strong>g variable (plus an Ohmic leak current) results <strong>in</strong> a m<strong>in</strong>imalmodel. Indeed, neither of the variables alone can produce oscillation, but together they


136 Conductance-Based Models<strong>in</strong>ward(Na, Ca)currentsoutward(K, Cl)amplify<strong>in</strong>gresonantactivation, mreversepotentialreversepotentialgat<strong>in</strong>g<strong>in</strong>activation, hresonantreversepotentialreversepotentialamplify<strong>in</strong>gFigure 5.2: Gat<strong>in</strong>g variables maybe amplify<strong>in</strong>g or resonant depend<strong>in</strong>gon whether they represent activation/<strong>in</strong>activationof <strong>in</strong>ward/outwardcurrents (see also Fig. 3.3 andFig. 3.4).can, as we will see below.The amplify<strong>in</strong>g gat<strong>in</strong>g variable is the activation variable m for voltage-gated <strong>in</strong>wardcurrent or <strong>in</strong>activation variable h for voltage-gated outward current, as <strong>in</strong> Fig. 5.2.These variables amplify voltage changes via a positive feedback loop. Indeed, a smalldepolarization <strong>in</strong>creases m and decreases h, which <strong>in</strong> turn <strong>in</strong>crease <strong>in</strong>ward and decreaseoutward currents and <strong>in</strong>crease depolarization. Similarly, a small hyperpolarizationdecreases m and <strong>in</strong>creases h, result<strong>in</strong>g <strong>in</strong> less <strong>in</strong>ward and more outward current, andhence <strong>in</strong> more hyperpolarization.The resonant gat<strong>in</strong>g variable is the <strong>in</strong>activation variable h for an <strong>in</strong>ward current oractivation variable n for an outward current. These variables resist voltage changes viaa negative feedback loop. A small depolarization decreases h and <strong>in</strong>creases n, which <strong>in</strong>turn decrease <strong>in</strong>ward and <strong>in</strong>crease outward currents and produce a net outward currentthat resists the depolarization. Similarly, a small hyperpolarization produces <strong>in</strong>wardcurrent and possibly rebound depolarization.Currents with amplify<strong>in</strong>g gat<strong>in</strong>g variables can result <strong>in</strong> bistability, and they behaveessentially like the I Na,p -model or I Kir -model considered <strong>in</strong> Chap. 3. Currents withresonant gat<strong>in</strong>g variables have one stable equilibrium with possibly damped oscillation,and they behave essentially like the I K -model or I h model (compare Fig. 5.2 withFig. 3.3). A typical neuronal model consists of at least one amplify<strong>in</strong>g and at least oneresonant gat<strong>in</strong>g variable. Amplify<strong>in</strong>g and resonant gat<strong>in</strong>g variables for Ca 2+ -sensitivecurrents are discussed at the end of this chapter.To get spikes <strong>in</strong> a m<strong>in</strong>imal model, we need a fast positive feedback and a slowernegative feedback. Indeed, if an amplify<strong>in</strong>g gat<strong>in</strong>g variable has a long time constant,it would act more as a low-pass filter, hardly affect<strong>in</strong>g fast fluctuations, and onlyamplify<strong>in</strong>g slow fluctuations. If a resonant gat<strong>in</strong>g variable has a fast time constant,it would act to damp <strong>in</strong>put fluctuations (faster than they could be amplified by theamplify<strong>in</strong>g variable), result<strong>in</strong>g <strong>in</strong> stability of the rest<strong>in</strong>g state. Instead, the resonant


Conductance-Based Models 137resonant gat<strong>in</strong>g variables<strong>in</strong>activation of<strong>in</strong>ward currentactivation ofoutward currentamplify<strong>in</strong>g gat<strong>in</strong>g variablesactivation of<strong>in</strong>ward current<strong>in</strong>activation ofoutward currentI Na,t -modelI Na,p +I h -modelI Kir +I h -modelI Na,p +I K -modelI Kir +I K -modelI A -modelFigure 5.3: Any comb<strong>in</strong>ation of oneamplify<strong>in</strong>g and one resonant gat<strong>in</strong>gvariables results <strong>in</strong> a spik<strong>in</strong>g model.variable acts as a band-pass filter; It has no effect on oscillations with a period muchsmaller than its time constant. It damps oscillations hav<strong>in</strong>g period much larger than itstime constant, because the variable oscillates <strong>in</strong> phase with the voltage fluctuations; Itamplifies oscillations with a period that is about the same as its time constant becausethe variable lags the voltage fluctuations.S<strong>in</strong>ce the amplify<strong>in</strong>g gat<strong>in</strong>g variable, say m, has relatively fast k<strong>in</strong>etics, it can bereplaced by its equilibrium (steady-state) value m ∞ (V ). This allows to reduce thedimension of the m<strong>in</strong>imal models from three (say V , m, n) to two (V and n).Two amplify<strong>in</strong>g and two resonant gat<strong>in</strong>g variables produce four different comb<strong>in</strong>ations,depicted <strong>in</strong> Fig. 5.3. However, the number of m<strong>in</strong>imal models is not four, but six.The additional models arise due to the fact that a pair of gat<strong>in</strong>g variables may describeactivation/<strong>in</strong>activation properties of the same current or of two different currents. Forexample, the activation and <strong>in</strong>activation gat<strong>in</strong>g variables m and h may describe thedynamics of a transient <strong>in</strong>ward current, such as I Na,t , or the dynamics of a comb<strong>in</strong>ationof one persistent <strong>in</strong>ward current, such as I Na,p , and one “hyperpolarization-activated”<strong>in</strong>ward current, such as I h . Hence this pair results <strong>in</strong> two models, I Na,t - and I Na,p +I h -model. Similarly, the pair of activation and <strong>in</strong>activation variables of an outward currentmay describe the dynamics of the same transient current, such as I A , or the dynamicsof two different outward currents, hence the two models, I A - and I Kir +I K -model.Below we present the geometrical analysis of the six m<strong>in</strong>imal voltage-gated modelsshown <strong>in</strong> Fig. 5.3. Though based on different ionic currents, the models have manysimilarities from the dynamical systems po<strong>in</strong>t of view. In particular, all can exhibitsaddle-node and Andronov-Hopf bifurcations. For each model we first provide a worddescription of the mechanism of generation of susta<strong>in</strong>ed oscillations, and then usephase plane analysis to provide a geometrical description. The first two, I Na,p +I K -and I Na,t -models are common; they describe the mechanism of generation of actionpotentials or subthreshold oscillations by many cells. The other four models are rare;they even might be classified as weird or bizarre by biologists, s<strong>in</strong>ce these models reveal


138 Conductance-Based Models1I=01I=0K + activation, n0.80.60.40 15n-nullcl<strong>in</strong>eK + activation, n0.80.60.4n-nullcl<strong>in</strong>e0 15V-nullcl<strong>in</strong>eV-nullcl<strong>in</strong>e0.20.200-60 -20 0 20membrane voltage, V-60 20membrane voltage, V1I=101I=40K + activation, n0.80.60.40 15n-nullcl<strong>in</strong>eK + activation, n0.80.60.4n-nullcl<strong>in</strong>e0 15V-nullcl<strong>in</strong>eV-nullcl<strong>in</strong>e0.20.200-60 -20 0 20membrane voltage, Va-60 -20 0 20membrane voltage, VbFigure 5.4: Possible forms of nullcl<strong>in</strong>es <strong>in</strong> the I Na,p +I K -model. Parameters as <strong>in</strong> theprevious chapter (see caption to Fig. 4.1).rather unexpected mechanisms for voltage oscillations. Nevertheless, it is educationalto consider all 6 models to see how the theory of dynamical systems works where<strong>in</strong>tuition and common sense fail.5.1.2 I Na,p +I K -modelOne of the most fundamental models <strong>in</strong> computational neuroscience is the I Na,p +I K -model consist<strong>in</strong>g of a fast Na + current and a relatively slower K + current:I Na,pleak I LIC ˙V{ }} { { }} { { }} K{= I − g L (V −E L ) − g Na m(V − E Na ) − g K n(V − E K ) ,ṁ = (m ∞ (V ) − m)/τ m (V ) ,ṅ = (n ∞ (V ) − n)/τ n (V ) .


Conductance-Based Models 139This model is <strong>in</strong> many respects equivalent to the I Ca +I K -model proposed by Morrisand Lecar (1981) to describe voltage oscillations <strong>in</strong> the barnacle giant muscle fiber.A reasonable assumption based on experimental observations is that the Na + gat<strong>in</strong>gvariable m(t) is much faster than the voltage variable V (t), so that m approaches theasymptotic value m ∞ (V ) practically <strong>in</strong>stantaneously. In this case we can substitutem = m ∞ (V ) <strong>in</strong> the voltage equation and reduce the three-dimensional system aboveto a planar systemleak I L<strong>in</strong>stantaneous I Na,p IC ˙V{ }} { { }} { { }} K{= I − g L (V −E L ) − g Na m ∞ (V ) (V −E Na ) − g K n (V −E K ) (5.1)ṅ = (n ∞ (V ) − n)/τ(V ) , (5.2)which was considered <strong>in</strong> detail <strong>in</strong> the previous chapter. In Fig. 5.4 we summarize itsdynamic repertoire. A strik<strong>in</strong>g observation is that the other m<strong>in</strong>imal models can havesimilar nullcl<strong>in</strong>es and similar dynamic repertoire, even though they consist of quitedifferent ionic currents.5.1.3 I Na,t -modelAn <strong>in</strong>terest<strong>in</strong>g example of a spik<strong>in</strong>g mechanism, implicitly present <strong>in</strong> practically everybiological neuron, is given by the I Na,t -modelC ˙V = I −leak I L{ }} {g L (V − E L ) −ṁ = (m ∞ (V ) − m)/τ m (V ) ,ḣ = (h ∞ (V ) − h)/τ h (V ) ,I Na,t{ }} {g Na m 3 h(V − E Na ) ,consist<strong>in</strong>g only of an Ohmic leak current and a transient voltage-gated <strong>in</strong>ward Na +current. How could such a model generate action potentials? The upstroke of an actionpotential is generated because of the regenerative process <strong>in</strong>volv<strong>in</strong>g the activation gatem. This mechanism is similar to the one <strong>in</strong> the Hodgk<strong>in</strong>-Huxley model or <strong>in</strong> theI Na,p +I K -model: Increase of m results <strong>in</strong> <strong>in</strong>crease of the <strong>in</strong>ward current, hence moredepolarization and further <strong>in</strong>crease of m until the excited state is achieved. At theexcited state there is a balance of the Na + <strong>in</strong>ward current and the leak outward current.S<strong>in</strong>ce there is no I K , the downstroke from the excited state occurs via a differentmechanism: While <strong>in</strong> the excited state, the Na + current <strong>in</strong>activates (turns off) and theOhmic leak current slowly repolarizes the membrane potential toward the leak reversepotential E L , which determ<strong>in</strong>es the rest<strong>in</strong>g state. While at rest, the Na + currentde<strong>in</strong>activates; i.e., becomes available, and the neuron is ready to generate anotheraction potential. This mechanism is summarized <strong>in</strong> Fig. 5.5.To study the dynamics of the I Na,t -model, we first reduce it to a planar system.Assum<strong>in</strong>g that activation dynamics is <strong>in</strong>stantaneous, we use m = m ∞ (V ) <strong>in</strong> the voltage


140 Conductance-Based Modelsleak currentrest<strong>in</strong>g potentialde<strong>in</strong>activationof I Na<strong>in</strong>jecteddc-current<strong>in</strong>activationof I Naactivationof I NadepolarizationFigure 5.5: Mechanism of generation of susta<strong>in</strong>edoscillations <strong>in</strong> the I Na,t -model.equation and obta<strong>in</strong>C ˙V = I −leak I L{ }} {g L (V − E L ) −ḣ = (h ∞ (V ) − h)/τ h (V ) .I Na,t , <strong>in</strong>st. activation{ }} {g Na m 3 ∞(V )h(V − E Na ) ,One can easily f<strong>in</strong>d the nullcl<strong>in</strong>esh =I − g L (V −E L )g Na m 3 ∞(V )(V − E Na )(V -nullcl<strong>in</strong>e)andh = h ∞ (V ) (h-nullcl<strong>in</strong>e) .The V -nullcl<strong>in</strong>e looks like a cubic parabola (flipped N shape), and the h-nullcl<strong>in</strong>e hasa sigmoid shape. In Fig. 5.6 we depict two typical cases (we <strong>in</strong>vert the h-axis so thatthe vector field is directed counterclockwise and this phase portrait is consistent withthe other phase portraits <strong>in</strong> this book).When the <strong>in</strong>activation curve h ∞ (V ) has a high threshold (i.e., I Na,t is a w<strong>in</strong>dowcurrent), there are three <strong>in</strong>tersections of the nullcl<strong>in</strong>es, hence three equilibria, as <strong>in</strong>Fig. 5.6a. A stable node (filled circle) corresponds to the rest state, and a nearbysaddle corresponds to the threshold state. Another equilibrium, an unstable focusdenoted by white circle at the top of the figure, determ<strong>in</strong>es the shape of the actionpotential s<strong>in</strong>ce all “spik<strong>in</strong>g” trajectories have to go around it. Because of the highthreshold of <strong>in</strong>activation, the I Na,t current is de<strong>in</strong>activated at rest. Moreover, smallfluctuations of V do not produce significant changes of the <strong>in</strong>activation variable hbecause the h-nullcl<strong>in</strong>e is nearly horizontal at rest. Such a system does not performdamped oscillations, and the nonl<strong>in</strong>ear dynamics of V near rest state can be describedby the one-dimensional system (where h = h ∞ (V ))C ˙V = I − g L (V − E L ) − g Na m 3 ∞(V )h ∞ (V )(V − E Na )studied <strong>in</strong> Chap. 3. When I <strong>in</strong>creases, the stable node and the saddle approach,coalesce, and annihilate each other via saddle-node bifurcation. When I = 0.5, there is


V-nullcl<strong>in</strong>eConductance-Based Models 1410I=00I=0V-nullcl<strong>in</strong>eNa + <strong>in</strong>activation, h Na + <strong>in</strong>activation, h0.20.40.60.8100.20.40.60.8h-nullcl<strong>in</strong>e0 100-50 0 50membrane voltage, Vh-nullcl<strong>in</strong>eNa + <strong>in</strong>activation, h0.20.40.60.80 1001-50 0 50membrane voltage, V0I=0.5 I=40.2Na + <strong>in</strong>activation, h0.40.6h-nullcl<strong>in</strong>e0 3000.80 100h-nullcl<strong>in</strong>eV-nullcl<strong>in</strong>eV-nullcl<strong>in</strong>e1-50 0 50membrane voltage, Va1-50 0 50membrane voltage, VbFigure 5.6: Possible forms of nullcl<strong>in</strong>es <strong>in</strong> the I Na,t -model. Notice that the h-axis is<strong>in</strong>verted. Parameters for I Na,t are as <strong>in</strong> Hodgk<strong>in</strong>-Huxley model except τ h (V ) = 5 ms.E Na = 60 mV, E L = −70 mV, g L = 1, g Na = 15 (<strong>in</strong> b) and g Na = 10 and V 1/2 = −42mV (<strong>in</strong> a).


142 Conductance-Based Modelshyperpolarizationdeactivationof I Na,pde<strong>in</strong>activationof I hleak currentor <strong>in</strong>jecteddc-current<strong>in</strong>activationof I hactivationof I Na,pdepolarizationFigure 5.7: Mechanism of generation of susta<strong>in</strong>edvoltage oscillations <strong>in</strong> the I Na,p +I h -model.a periodic trajectory with a long period (compare the time scales <strong>in</strong> the bottom <strong>in</strong>sets<strong>in</strong> Fig. 5.6a and b).When the Na + <strong>in</strong>activation curve h ∞ (V ) has a low threshold, the nullcl<strong>in</strong>es haveonly one <strong>in</strong>tersection, hence there is only one equilibrium, as <strong>in</strong> Fig. 5.6b. WhenI = 0, the equilibrium (filled circle) is stable and all trajectories converge to it. Thereare damped oscillations near the equilibrium, though they can hardly be seen <strong>in</strong> thefigure. The oscillations occur because the I Na,t current is partially <strong>in</strong>activated at rest.An <strong>in</strong>crease of V leads to more <strong>in</strong>activation, less <strong>in</strong>ward current, and hence rebounddecrease of V , which <strong>in</strong> turn leads to partial de<strong>in</strong>activation, more <strong>in</strong>ward current, andrebound <strong>in</strong>crease of V . When the applied dc-current I <strong>in</strong>creases, the equilibrium losesstability via Andronov-Hopf bifurcation. When I = 4, the equilibrium is an unstablefocus (white circle <strong>in</strong> the figure), and there is a stable limit cycle attractor around itcorrespond<strong>in</strong>g to periodic spik<strong>in</strong>g.We see that the I Na,t -model exhibits essentially the same dynamic repertoire as theI Na,p +I K -model, even though the models are quite different from the electrophysiologicalpo<strong>in</strong>t of view.5.1.4 I Na,p +I h -modelThe systemI Na,pleak IC ˙V{ }} L{ { }} { { }} {= I − g L (V −E L ) − g Na m(V − E Na ) − g h h(V − E h ) ,ṁ = (m ∞ (V ) − m)/τ m (V ) ,ḣ = (h ∞ (V ) − h)/τ h (V ) ,I his believed to describe the essence of the mechanism of slow subthreshold voltage oscillations<strong>in</strong> some cortical, thalamic, and hippocampal neurons, which we summarize<strong>in</strong> Fig. 5.7. As any other m<strong>in</strong>imal model <strong>in</strong> this section, it consists of one amplify<strong>in</strong>g(I Na,p ) and one resonant (I h ) current. Both currents may be partially active at rest<strong>in</strong>gvoltage. Recall that we treat the h-current as be<strong>in</strong>g an <strong>in</strong>ward current that is alwaysactivated (its activation variable m = 1 all the time), but can be <strong>in</strong>activated (turned


Conductance-Based Models 143off) by depolarization and de<strong>in</strong>activated (turned on) by hyperpolarization. At rest<strong>in</strong>gvoltage this current is usually <strong>in</strong>activated (turned off). A sufficient hyperpolarizationof V de<strong>in</strong>activates (turns on) the h-current result<strong>in</strong>g <strong>in</strong> rebound depolarization. Whiledepolarized, the h-current <strong>in</strong>activates (turns off), and the leak current repolarizes themembrane potential toward rest<strong>in</strong>g state. Without the persistent Na + current, or someother amplify<strong>in</strong>g current, these oscillations always subside, as the reader was asked toprove <strong>in</strong> Chap. 4, Ex. 10. However, they may become susta<strong>in</strong>ed when I Na,p is <strong>in</strong>volved.To study dynamics of the I Na,p +I h -model we assume that the activation k<strong>in</strong>eticsof the Na + current is <strong>in</strong>stantaneous, and use m = m ∞ (V ) <strong>in</strong> the voltage equation toobta<strong>in</strong> a two-dimensional systemleak I L<strong>in</strong>stantaneous I Na,p IC ˙V{ }} { { }} { { }} h{= I − g L (V −E L ) − g Na m ∞ (V )(V − E Na ) − g h h(V − E h ) ,ḣ = (h ∞ (V ) − h)/τ h (V ) .The nullcl<strong>in</strong>es of this systemh = I − g L(V −E L ) − g Na m ∞ (V )(V − E Na )g h (V − E h )(V -nullcl<strong>in</strong>e)andh = h ∞ (V )(h-nullcl<strong>in</strong>e)have the familiar N and sigmoid shapes depicted <strong>in</strong> Fig. 5.8. We take the parameters forboth currents from the experimental studies of thalamic relay neurons (see Sect. 2.3.5).This choice results <strong>in</strong> one <strong>in</strong>tersection of the nullcl<strong>in</strong>es <strong>in</strong> the relevant voltage range,which corresponds to only one equilibrium. This equilibrium is a stable rest<strong>in</strong>g statewhen no current is <strong>in</strong>jected, i.e., when I = 0. In Fig. 5.8, top, one can clearly seethat h ≈ 0; that is, the h-current is <strong>in</strong>activated (turned off). The rest state is dueto the balance of <strong>in</strong>ward persistent Na + current and the Ohmic leak current. A smallhyperpolarization deactivates the fast Na + current and shifts the balance toward theleak current, which br<strong>in</strong>gs V closer to E leak . This, <strong>in</strong> turn, results <strong>in</strong> slow de<strong>in</strong>activation(turn<strong>in</strong>g on) of the h-current, which produces strong <strong>in</strong>ward current and br<strong>in</strong>gs themembrane voltage back to the rest<strong>in</strong>g state.Negative <strong>in</strong>jected current (case I = −1 <strong>in</strong> Fig. 5.8) destroys the balance of <strong>in</strong>ward(I Na,p ) and outward (I leak ) currents at rest, and makes the rest<strong>in</strong>g state unstable. As aresult, the model exhibits susta<strong>in</strong>ed subthreshold oscillations of membrane potential.Indeed, prolonged hyperpolarization turns on strong h-current and produces prolongeddepolarization. Such a depolarization turns off the h-current, and the negative <strong>in</strong>jectedcurrent hyperpolarizes the membrane potential aga<strong>in</strong>. As a result, the model exhibitssusta<strong>in</strong>ed oscillations <strong>in</strong> the voltage range of −55 mV to −65 mV. The frequency of suchoscillations depends on the parameters of the voltage equation and the time constantτ(V ) of the h-current, and it is near 4 Hz <strong>in</strong> Fig. 5.8.


144 Conductance-Based Models0I=0<strong>in</strong>activation, h0.20.40.6-50-60-700 time, ms 1000h-nullcl<strong>in</strong>e0V-nullcl<strong>in</strong>e0.80.06-70 -60 -5010I=-1<strong>in</strong>activation, h0.20.40.6-50-60-700 time, ms 1000h-nullcl<strong>in</strong>e0V-nullcl<strong>in</strong>e0.80.06-70 -60 -501-100 -90 -80 -70 -60 -50membrane voltage, VFigure 5.8: Rest and susta<strong>in</strong>ed subthreshold oscillations <strong>in</strong> the I Na,p +I h -model. Parametersfor currents as <strong>in</strong> thalamocortical neurons, E Na = 20 mV, E h = −43 mV,E L = −80 mV, g L = 1.3, g Na = 0.9, and g h = 3.5.1.5 I h +I Kir -modelThe persistent Na + current, which amplifies the damped oscillations <strong>in</strong> the I Na,p +I h -model, can be substituted by the K + <strong>in</strong>wardly rectify<strong>in</strong>g current I Kir to achieve thesame amplify<strong>in</strong>g effect. The result<strong>in</strong>g I h +I Kir -modelleak I L IC ˙V{ }} {KirI{ }} { { }} h{= I − g L (V −E L ) − g Kir h Kir (V − E K ) − g h h(V − E h ) ,ḣ Kir = (h Kir,∞ (V ) − h Kir )/τ Kir (V ) ,ḣ = (h ∞ (V ) − h)/τ h (V ) ,can exhibit susta<strong>in</strong>ed subthreshold oscillations of membrane voltage via a rather weirdmechanism illustrated <strong>in</strong> Fig. 5.9. The <strong>in</strong>wardly rectify<strong>in</strong>g K + current I Kir behavessimilarly to I h , except that it is an outward current. A brief hyperpolarization de<strong>in</strong>activates(turns on) the fast outward current I Kir and produces more hyperpolarization


Conductance-Based Models 145leak currenthyperpolarizationde<strong>in</strong>activationof I Kirde<strong>in</strong>activationof I h<strong>in</strong>activationof I h<strong>in</strong>activationof I KirdepolarizationFigure 5.9: Mechanism of generation of susta<strong>in</strong>edvoltage oscillations <strong>in</strong> the I h +I Kir -model.via a positive feedback loop. Such a regenerative process results <strong>in</strong> a prolonged hyperpolarizationthat de<strong>in</strong>activates (turns on) the slower <strong>in</strong>ward current I h and produces arebound depolarization. This depolarization is enhanced by <strong>in</strong>activation (turn<strong>in</strong>g off)of the fast I Kir . However, the membrane potential cannot hold long <strong>in</strong> the depolarizedstate because of the slow <strong>in</strong>activation of I h , and the leak current repolarizes themembrane potential. The repolarization is enhanced by the de<strong>in</strong>activation of I Kir andbecomes a hyperpolarization aga<strong>in</strong>, lead<strong>in</strong>g to the oscillations summarized <strong>in</strong> Fig. 5.9.S<strong>in</strong>ce the k<strong>in</strong>etics of I Kir is practically <strong>in</strong>stantaneous, we can use h Kir = h Kir,∞ (V )<strong>in</strong> the voltage equation above and consider the two-dimensional systemleak I L <strong>in</strong>stantaneous IC ˙V{ }} {Kir I{ }} { { }} h{= I − g L (V −E L ) − g Kir h Kir,∞ (V )(V − E K ) − g h h(V − E h ) ,ḣ = (h ∞ (V ) − h)/τ h (V ) .One can easily f<strong>in</strong>d the nullcl<strong>in</strong>es of this systemandh = I − g L(V −E L ) − g Kir h Kir,∞ (V )(V − E K )g h (V − E h )h = h ∞ (V ) (h-nullcl<strong>in</strong>e) ,(V -nullcl<strong>in</strong>e)which have the familiar form depicted <strong>in</strong> Fig. 5.10. Most values of the parametersresult <strong>in</strong> a phase portrait similar to the one depicted <strong>in</strong> Fig. 5.10, left. The V -nullcl<strong>in</strong>eis a monotonic curve that <strong>in</strong>tersects the h-nullcl<strong>in</strong>e <strong>in</strong> one po<strong>in</strong>t correspond<strong>in</strong>g to astable rest<strong>in</strong>g state. An <strong>in</strong>jected dc-current I shifts the rest<strong>in</strong>g state, but does notchange its stability: Voltage perturbations always subside, result<strong>in</strong>g only <strong>in</strong> dampedoscillations. There is, however, a narrow region <strong>in</strong> parameter space (it took the authora few hours to f<strong>in</strong>d that region) that produces just the right relationship between<strong>in</strong>activation curves and conductances so that the V -nullcl<strong>in</strong>e becomes N-shaped, andthe subthreshold oscillations become susta<strong>in</strong>ed, as <strong>in</strong> Fig. 5.10, right.5.1.6 I K +I Kir -modelThe last two m<strong>in</strong>imal models consist exclusively of outward K + currents, yet they canexhibit susta<strong>in</strong>ed oscillations of membrane voltage. The models defy imag<strong>in</strong>ation of


146 Conductance-Based Models0I=9.750I=100.10.1<strong>in</strong>activation, h0.20.3h-nullcl<strong>in</strong>e-50-600 time, s 2-70 -65 -60 -55 -50 -45membrane voltage, VV-nullcl<strong>in</strong>e<strong>in</strong>activation, h0.20.3h-nullcl<strong>in</strong>e0 time, s 2-70 -65 -60 -55 -50 -45membrane voltage, VFigure 5.10: Rest and susta<strong>in</strong>ed subthreshold oscillations <strong>in</strong> the I h +I Kir -model. Parameters:E K = −80 mV, E h = −43 mV, E L = −50 mV, g Kir = 4, g h = 0.5, g L = 0.44.h-current is the same as <strong>in</strong> the previous section, except V 1/2 = −65 mV. InstantaneousI Kir has V 1/2 = −76 mV and k = −11.-50-60V-nullcl<strong>in</strong>ehyperpolarizationdeactivationof I K<strong>in</strong>jecteddc-currentactivationof I K<strong>in</strong>activationof I KirdepolarizationFigure 5.11: Mechanism of generation of susta<strong>in</strong>edvoltage oscillations <strong>in</strong> the I K +I Kir -model.many biologists: How can a neuron with only outward K + currents and no <strong>in</strong>ward Na +or Ca 2+ currents fire action potentials?In the I K +I Kir -modelI KirIC ˙V{ }} { { }} K{= I − g Kir h(V − E K ) − g K n(V − E K ) ,ṅ = (n ∞ (V ) − n)/τ n (V ) ,ḣ = (h ∞ (V ) − h)/τ h (V ) ,the amplify<strong>in</strong>g current is I Kir with <strong>in</strong>activation gat<strong>in</strong>g variable h, and the resonantcurrent is I K with activation variable n. The mechanism of generation of action potentialsis summarized <strong>in</strong> Fig. 5.11. A strong <strong>in</strong>jected current depolarizes the membranepotential and <strong>in</strong>activates (turns off) I Kir , which amplifies the depolarization. Whiledepolarized, the slower K + current I K activates and br<strong>in</strong>gs the potential down withpossible hyperpolarization, which is amplified by the de<strong>in</strong>activation of I Kir . While themembrane potential is hyperpolarized, I K deactivates and the strong <strong>in</strong>jected current


Conductance-Based Models 147br<strong>in</strong>gs the potential up aga<strong>in</strong>. Thus, the upstroke of the action potential is due exclusivelyto the <strong>in</strong>jected dc-current I, while the downstroke is due to the persistentoutward K + current.To perform the geometrical phase plane analysis of the model we take advantage ofthe same observation as before: The k<strong>in</strong>etics of the amplify<strong>in</strong>g current I Kir is relativelyfast, so that h = h ∞ (V ) can be used <strong>in</strong> the voltage equation to reduce the threedimensionalsystem above to the two-dimensional system<strong>in</strong>stantaneous I Kir IC ˙V{ }} { { }} K{= I − g Kir h ∞ (V )(V − E K ) − g K n(V − E K ) ,ṅ = (n ∞ (V ) − n)/τ n (V ) .It is an easy exercise to f<strong>in</strong>d the nullcl<strong>in</strong>esn = I/{g K (V − E K )} − g Kir h ∞ (V )/g K(V -nullcl<strong>in</strong>e)andn = n ∞ (V ) (n-nullcl<strong>in</strong>e) ,which we depict <strong>in</strong> Fig. 5.12. There are two <strong>in</strong>terest<strong>in</strong>g cases correspond<strong>in</strong>g to highthreshold(Fig. 5.12a) and low-threshold (Fig. 5.12b) K + current I K .When the I K has low threshold, it is partially activated at rest<strong>in</strong>g potential. In thiscase, rest state corresponds to the balance of partially activated I K , partially <strong>in</strong>activatedI Kir , and a strong <strong>in</strong>jected dc-current I (without the dc-current the membrane voltagewould converge to E K = −80 mV and stay there forever). A small depolarizationpartially <strong>in</strong>activates fast I Kir but leaves slower I K relatively unchanged. This results<strong>in</strong> an imbalance of the <strong>in</strong>ward dc-current I and all outward currents, and the net<strong>in</strong>ward current further depolarizes the membrane voltage. Depend<strong>in</strong>g on the size of thedepolarization, the model may generate a subthreshold response or an action potential,as one can see <strong>in</strong> Fig. 5.12b, top. Dur<strong>in</strong>g the generation of the action potential, thepersistent K + current activates and causes after-hyperpolarization. Dur<strong>in</strong>g the afterhyperpolarization,the persistent K + current deactivates below the rest<strong>in</strong>g level. Thislets the <strong>in</strong>jected dc-current I depolarize the membrane potential aga<strong>in</strong>, provided thatI is strong enough, as <strong>in</strong> Fig. 5.12b, bottom.In Fig. 5.12a we leave all parameters unchanged except that we <strong>in</strong>crease the halfvoltageactivation V 1/2 of I K by 15 mV and decrease I to compensate for the deficitof outward current. Now, the rest<strong>in</strong>g state corresponds to the balance of the I Kir andI, because the high-threshold persistent K + current is completely deactivated <strong>in</strong> thisvoltage range. The behavior near the rest state is determ<strong>in</strong>ed by the <strong>in</strong>terplay between<strong>in</strong>stantaneous I Kir and I, and it was studied <strong>in</strong> Chap. 3 (see I Kir -model). There are twoequilibria: a stable node correspond<strong>in</strong>g to the rest<strong>in</strong>g state, and a saddle correspond<strong>in</strong>gto the threshold state. A sufficiently strong perturbation can push V beyond the saddleequilibrium, as <strong>in</strong> Fig. 5.12a, top, and can cause the m<strong>in</strong>imal model to fire an action


148 Conductance-Based Models1I=651I=700.80.8V-nullcl<strong>in</strong>eV-nullcl<strong>in</strong>eK + activation, n0.60.40.2n-nullcl<strong>in</strong>e0 50K + activation, n0.60.40.2n-nullcl<strong>in</strong>e000 50-80 -60 -40 -20 0 20membrane voltage, V-80 -60 -40 -20 0 20membrane voltage, V1I=681I=73V-nullcl<strong>in</strong>e0.80.8V-nullcl<strong>in</strong>eK + activation, n0.60.40.2n-nullcl<strong>in</strong>e0 100K + activation, n0.60.40.2n-nullcl<strong>in</strong>e000 100-80 -60 -40 -20 0 20membrane voltage, Va-80 -60 -40 -20 0 20membrane voltage, VbFigure 5.12: Possible <strong>in</strong>tersections of nullcl<strong>in</strong>es <strong>in</strong> the I K +I Kir -model. Parameters:E K = −80 mV, g Kir = 20, g K = 2. Instantaneous I Kir with V 1/2 = −80 mV andk = −12. Slower I K with k = 5, τ(V ) = 5 ms, and V 1/2 = −40 mV (<strong>in</strong> a) orV 1/2 = −55 mV (<strong>in</strong> b).potential. If we <strong>in</strong>crease I, the node and the saddle approach, coalesce, and annihilateeach other via a saddle-node bifurcation, and the model start to fire action potentialsperiodically.We see that I K +I Kir -model has essentially the same dynamic repertoire as the moreconventional I Na,p +I K -model or I Na,t -model, despite the fact that it is based on a ratherbizarre ionic mechanism for excitability and spik<strong>in</strong>g.5.1.7 I A -modelThe last m<strong>in</strong>imal voltage-gated model has only one transient K + current, often referredto as be<strong>in</strong>g A-current I A , yet it can also generate susta<strong>in</strong>ed oscillations. In some sense,the model is similar to the I Na,t -model. Indeed, each consists of only one transient


Conductance-Based Models 149de-<strong>in</strong>activationof I Ahyperpolarizationdeactivationof I A<strong>in</strong>jecteddc-currentactivation +<strong>in</strong>activationof I Arest<strong>in</strong>g potentialFigure 5.13: Mechanism of generation of susta<strong>in</strong>edvoltage oscillations <strong>in</strong> the I A -model.current and an Ohmic leak current. The only difference is that A-current is outward,and as a result, the action potentials are fired downward; see Fig. 5.14 and Fig. 5.15below.The A-current has activation and <strong>in</strong>activation variables m and h, respectively, andthe model has the formC ˙V = I −leak I L{ }} {g L (V − E L ) −ṁ = (m ∞ (V ) − m)/τ m (V )ḣ = (h ∞ (V ) − h)/τ h (V ) .I A{ }} {g A mh(V − E K )The mechanism of generation of downward action potentials is summarized <strong>in</strong> Fig. 5.13.Due to a strong <strong>in</strong>jected dc-current, the rest state is <strong>in</strong> the depolariz<strong>in</strong>g voltage range,and it corresponds to the balance of the partially activated, partially <strong>in</strong>activated A-current, leak outward current, and the <strong>in</strong>jected dc-current. A small hyperpolarizationcan simultaneously deactivate and de<strong>in</strong>activate the A-current, i.e., decrease variablem and <strong>in</strong>crease variable h. Depend<strong>in</strong>g on the relationship between the activation and<strong>in</strong>activation time constants, this may result <strong>in</strong> an <strong>in</strong>crease of the A-current conductance,which is proportional to the product mh. More outward current produces morehyperpolarization and even more outward current. As a result of this regenerativeprocess, the membrane voltage produces a sudden downstroke. While hyperpolarized,the A-current deactivates (variable m → 0), and the <strong>in</strong>jected dc-current slowly br<strong>in</strong>gsthe membrane potential toward the rest<strong>in</strong>g state, result<strong>in</strong>g <strong>in</strong> a slow upstroke. Fastdownstroke and a slower upstroke from a depolarized rest<strong>in</strong>g state look like an actionpotential po<strong>in</strong>t<strong>in</strong>g downwards.If activation k<strong>in</strong>etics is much faster than the <strong>in</strong>activation k<strong>in</strong>etics, we can substitutem = m ∞ (V ) <strong>in</strong>to the voltage equation above and reduce the I A -model to atwo-dimensional system, which hopefully would have the right k<strong>in</strong>d of nullcl<strong>in</strong>es anda limit cycle attractor. After all, this is what we have done with previous m<strong>in</strong>imalmodels and it always worked. As the reader is asked to prove <strong>in</strong> Ex. 1, the I A -modelcannot have a limit cycle attractor when the A-current activation k<strong>in</strong>etics is <strong>in</strong>stantaneous.Oscillations are possible only when the activation and <strong>in</strong>activation k<strong>in</strong>etics


150 Conductance-Based Modelshave comparable time constants or <strong>in</strong>activation is much faster than activation.Even though none of the experimentally measured A-currents show fast <strong>in</strong>activationand a relatively slower activation, this case is still <strong>in</strong>terest<strong>in</strong>g from the pure theoreticalpo<strong>in</strong>t of view, s<strong>in</strong>ce it shows how a s<strong>in</strong>gle K + current can give rise to oscillations.Assum<strong>in</strong>g <strong>in</strong>stantaneous <strong>in</strong>activation and us<strong>in</strong>g h = h ∞ (V ) <strong>in</strong> the voltage equation, weobta<strong>in</strong> a two-dimensional systemC ˙V = I −leak I L{ }} {g L (V − E L ) −ṁ = (m ∞ (V ) − m)/τ m (V ) ,I A , <strong>in</strong>st. <strong>in</strong>activation{ }} {g A m h ∞ (V )(V − E K ) ,whose nullcl<strong>in</strong>es can easily be found:m = I − g L(V −E L )g A h ∞ (V )(V − E K )(V -nullcl<strong>in</strong>e)andm = m ∞ (V ) (m-nullcl<strong>in</strong>e) .Two typical cases are depicted <strong>in</strong> Fig. 5.14a and b. We start with the simpler case <strong>in</strong>Fig. 5.14b.Figure 5.14b depicts nullcl<strong>in</strong>es when the A-current has low activation threshold.There is only one <strong>in</strong>tersection of the nullcl<strong>in</strong>es, hence there is only one equilibrium,which is a stable focus when the <strong>in</strong>jected dc-current I is not strong enough (Fig. 5.14b,top). Increas<strong>in</strong>g I makes the equilibrium lose stability via a supercritical Andronov-Hopf bifurcation that gives birth to a small amplitude limit cycle attractor (not shown<strong>in</strong> the figure). A further <strong>in</strong>crease of I <strong>in</strong>creases the amplitude of oscillations, and e.g.,when I = 10 (middle of Fig. 5.14b), the attractor corresponds to periodic fir<strong>in</strong>g of actionpotentials. When I = 10.5, the attractor disappears and the equilibrium becomesstable (via Andronov-Hopf bifurcation) aga<strong>in</strong>. The model, however, becomes excitable.A small hyperpolarization does not change significantly the A-current, and the voltagereturns back to rest result<strong>in</strong>g <strong>in</strong> a “subthreshold response”. A sufficiently largehyperpolarization de<strong>in</strong>activates enough I A to open the K + current and hyperpolarizethe membrane even further. This regenerative process produces the downstroke andbr<strong>in</strong>gs V close to E K . Dur<strong>in</strong>g the state of hyperpolarization, the A-current deactivates(m → 0), and the dc-current I br<strong>in</strong>gs V back to rest. Notice that the action potentialis directed downward.In Fig. 5.14a we consider the I A -model with exactly the same parameters exceptthat we shift the half-voltage activation V 1/2 of I A by 10 mV so that the A-currenthas higher activation threshold. This does not affect much the behavior of the systemwhen I is small. However, when I ≈ 10.7 the spik<strong>in</strong>g limit cycle attractor undergoesa new k<strong>in</strong>d of bifurcation — saddle-node bifurcation — result<strong>in</strong>g <strong>in</strong> the appearance oftwo new equilibria: a stable node and a saddle. If the reader looks at Fig. 5.14a upsidedown,he will notice that this figure resembles figures 5.4a, 5.6a, or 5.12a, with all the


Conductance-Based Models 151K + activation, m10.80.60.40.20 time, ms 200m-nullcl<strong>in</strong>eI=8K + activation, m10.80.60.40.2V-nullcl<strong>in</strong>e0 time, ms 200I=6m-nullcl<strong>in</strong>eV-nullcl<strong>in</strong>eK + activation, mK + activation, m0-80 -60 -40 -20 0membrane voltage, V10.80.60.40.20 time, ms 5000-80 -60 -40 -20 0membrane voltage, V10.80.60.40.2I=10.6m-nullcl<strong>in</strong>eI=10.8m-nullcl<strong>in</strong>e0 time, ms 2000-80 -60 -40 -20 0membrane voltage, VaV-nullcl<strong>in</strong>eV-nullcl<strong>in</strong>eK + activation, mK + activation, m0-80 -60 -40 -20 0membrane voltage, V10.80.60.40.20-80 -60 -40 -20 0membrane voltage, V10.80.60.40.20 time, ms 500m-nullcl<strong>in</strong>e0 time, ms 200m-nullcl<strong>in</strong>eI=10.50-80 -60 -40 -20 0membrane voltage, VbI=10V-nullcl<strong>in</strong>eV-nullcl<strong>in</strong>eFigure 5.14: Possible <strong>in</strong>tersections of nullcl<strong>in</strong>es <strong>in</strong> the I A -model. Parameters: E K =−80 mV, E L = −60 mV, g A = 5, g L = 0.2. Instantaneous <strong>in</strong>activation k<strong>in</strong>etics withV 1/2 = −66 mV and k = −10. Activation of the A-current with k = 10, τ(V ) = 20 ms,and V 1/2 = −45 mV (<strong>in</strong> a) or V 1/2 = −35 mV (<strong>in</strong> b).


152 Conductance-Based Models1 sec100 nA50 mV20 mV(a) (b) 1 secFigure 5.15: Anomalous (upside-down) spikes <strong>in</strong> (a) lobster muscle fibers (modifiedfrom Fig.2 of Reuben et al. 1961) and <strong>in</strong> (b) Ascaris Esophageal cells (modified fromFig.16 of del Castillo and Morales 1967; the cell is depolarized by <strong>in</strong>jected dc-current).The voltage axis is not <strong>in</strong>verted.consequences: The node corresponds to a rest<strong>in</strong>g state, and the saddle correspondsto the threshold state. The large-amplitude trajectory that starts at the saddle andterm<strong>in</strong>ates at the node corresponds to an action potential, though to a weird one.Thus, the behavior of this model is similar to the behavior of other models with theexception that the V -axis is reversed.The existence of “upside-down” K + spikes may (or better say does) look bizarre tomany researchers, even though “<strong>in</strong>verted” K + and Cl − spikes were reported <strong>in</strong> manypreparations, <strong>in</strong>clud<strong>in</strong>g frog and toad axons, squid axons, lobster muscle fibers, dogcardiac muscle, etc., as reviewed by Reuben et al. (1961) and Grundfest (1971). Twosuch cases are depicted <strong>in</strong> Fig. 5.15. Interest<strong>in</strong>gly, Reuben et al. (1961) postulated,albeit reluctantly, that the <strong>in</strong>verted spikes are caused by the <strong>in</strong>activation of K + current.The reluctance was due to the fact that transient K + I A was not known at that time.By now the reader must have conv<strong>in</strong>ced himself that quite different models canhave practically identical dynamics. Conversely, the same model could have quite differentbehavior if only one parameter, e.g., V 1/2 , is changed by as little as 10 mV.Such dramatic conclusions emphasize the importance of geometrical phase plane analysisof neuronal models, s<strong>in</strong>ce the conclusions can hardly be drawn from mere worddescriptions of the spik<strong>in</strong>g mechanisms.5.1.8 Ca 2+ -gated m<strong>in</strong>imal modelsSo far, we considered m<strong>in</strong>imal models consist<strong>in</strong>g of voltage-gated currents only. However,there are many ionic currents that depend not only on the membrane potential,but also on the concentration of <strong>in</strong>tracellular ions, mostly Ca 2+ . Such currents arereferred to as be<strong>in</strong>g Ca 2+ -gated, and they are summarized <strong>in</strong> Fig. 5.16. In addition,there are Cl − -gated, K + -gated, and Na + -gated currents, such as SLO gene family ofCl − -gated K + currents discovered <strong>in</strong> C. elegans, and related “slack and slick” familyof Na + -gated K + currents (Yuan et al. 2000, 2003). Consider<strong>in</strong>g m<strong>in</strong>imal models<strong>in</strong>volv<strong>in</strong>g these currents goes outside the scope of this book, but it could be a good


Conductance-Based Models 153Voltage-GatedCa 2+ -GatedCurrentsActivationInactivationActivationInactivationI leakInwardI Na,pI CafastI Na,tI Ca(T)fastfastI Ca(P)fastslowI Ca(L)fastslowI Ca(N)fastmediummediumI hmediumI CANslowOutwardI KfastI K(M)slowI AfastfastI K(D)mediumslowI K(Ca)fastfastI AHPslowI KirfastFigure 5.16: Some representative voltage- and Ca 2+ -gated ionic currents (Johnston andWu 1995, Hille 2001, Shepherd 2004).<strong>in</strong>tellectual exercise for an expert reader (see also Ex. 7 and 8).Ca 2+ -gated currents can also be divided <strong>in</strong>to amplify<strong>in</strong>g and resonant. Ca 2+ -activated <strong>in</strong>ward currents, such as the cation non-selective I CAN , act as amplify<strong>in</strong>gcurrents. Indeed, activation of such a current leads to an <strong>in</strong>flux of Ca 2+ ions and tomore activation. Similarly, a hypothetical outward current <strong>in</strong>activated by Ca 2+ , notpresent <strong>in</strong> the figure, might also act as an amplify<strong>in</strong>g current. Indeed, a depolarizationdue to the Ca 2+ <strong>in</strong>flux <strong>in</strong>activates such a hypothetical outward current, therebyproduc<strong>in</strong>g a net shift toward <strong>in</strong>ward currents and lead<strong>in</strong>g to more depolarization.In contrast, Ca 2+ -<strong>in</strong>activat<strong>in</strong>g <strong>in</strong>ward currents and Ca 2+ -activat<strong>in</strong>g outward currents,such as I Ca(L) and I AHP , respectively, act as resonant currents. Indeed, a depolarizationdue to the Ca 2+ <strong>in</strong>flux <strong>in</strong>activates the <strong>in</strong>ward current and activates theoutward current, and resists further depolarization.


154 Conductance-Based Modelsvoltage-gated<strong>in</strong>activationof <strong>in</strong>wardcurrentresonant gateactivationof outwardcurrent<strong>in</strong>activationof <strong>in</strong>wardcurrentCa 2+ -gatedactivationof outwardcurrentamplify<strong>in</strong>g gatevoltage-gatedCa 2+ -gatedactivationof <strong>in</strong>wardcurrent<strong>in</strong>activationof outwardcurrentactivationof <strong>in</strong>wardcurrent<strong>in</strong>activationof outwardcurrentINa,p+IhINa,tIKir+IhICAN+IhI1I4+I3INa,p+IKIKir+IKIAICAN+IKICa+I2ICa(L)IKir+I2ICAN+I2ICAN,tI4+I2ICa+IK(Ca)ICAN+IK(Ca)Figure 5.17: Voltage- and Ca 2+ -gated m<strong>in</strong>imal models. I 1 , . . . , I 4 are four hypotheticalcurrents that could exist theoretically, but has never been recorded experimentally; seetext.Any comb<strong>in</strong>ation of one voltage- or Ca 2+ -gated amplify<strong>in</strong>g current and one voltageorCa 2+ -gated resonant current leads to a m<strong>in</strong>imal model for spik<strong>in</strong>g. All such comb<strong>in</strong>ationsare depicted <strong>in</strong> Fig. 5.17. Here, I 1 denotes a hypothetical Ca 2+ -activatedvoltage-<strong>in</strong>activated transient <strong>in</strong>ward current. Though such a current is not currentlyknown, one can easily write a conductance-gated model for it. A biologist would treatsuch a current as hyperpolarization- and Ca 2+ -activated. I 2 is a hypothetical Ca 2+ currentthat is <strong>in</strong>activated by Ca 2+ . I 3 is a hypothetical voltage-<strong>in</strong>activated Ca 2+ current.I 4 is an outward Ca 2+ -<strong>in</strong>activated current.We see that there are many m<strong>in</strong>imal models <strong>in</strong> Fig. 5.17. Six of them are purelyvoltage-gated, and they have been <strong>in</strong>vestigated above. The others are mixed-mode orpurely Ca 2+ -gated models. An <strong>in</strong>terested reader can work out the details of their phaseportraits.


Conductance-Based Models 15510050V(t)010.5m(t)n(t)h(t)02n(t)+h(t)0.8400 10 20 30 40 50 60 70 80 90 100time, t (ms)Figure 5.18: The sum n(t) + h(t) ≈ 0.84 <strong>in</strong> the Hodgk<strong>in</strong>-Huxley model. Parameters as<strong>in</strong> Chap. 2 and I = 8.5.2 Reduction of multi-dimensional models5.2.1 Hodgk<strong>in</strong>-Huxley modelLet us consider aga<strong>in</strong> the Hodgk<strong>in</strong>-Huxley modelC ˙V = I −I K{ }} {g K n 4 (V − E K ) −ṅ = (n ∞ (V ) − n)/τ n (V ) ,ṁ = (m ∞ (V ) − m)/τ m (V ) ,ḣ = (h ∞ (V ) − h)/τ h (V ) ,I Na{ }} {g Na m 3 h(V − E Na ) −I L{ }} {g L (V − E L ) ,with the orig<strong>in</strong>al values of parameters presented <strong>in</strong> Chap. 2. How can we understandthe qualitative dynamics of this model? One way, discussed above, is to throw awayvariable h or n and to reduce this model to the I Na,p +I K -model or I Na,t -model, respectively.Although the reduced m<strong>in</strong>imal models can tell a lot about the behaviorof the orig<strong>in</strong>al model, they are not equivalent to the Hodgk<strong>in</strong>-Huxley model from theelectrophysiological or the dynamical systems po<strong>in</strong>t of view. Below we discuss anothermethod of reduction of multidimensional electrophysiological models to planar systems.The Hodgk<strong>in</strong>-Huxley model has four <strong>in</strong>dependent variables. Early computer simulationsby Kr<strong>in</strong>skii and Kokoz (1973) have shown that there is a relationship between


156 Conductance-Based Models1h=0.89-1.1n0.8h0.60.40.200 0.2 0.4 0.6 0.8 1nFigure 5.19: The relationship between n(t) andh(t) <strong>in</strong> the Hodgk<strong>in</strong>-Huxley model can better bedescribed by h = 0.89 − 1.1n.the gat<strong>in</strong>g variables n(t) and h(t), namely,n(t) + h(t) ≈ 0.84 ,as shown <strong>in</strong> Fig. 5.18. In fact, plott<strong>in</strong>g the variables on the (n, h) plane, as we do <strong>in</strong>Fig. 5.19, reveals that the orbit is near the straight l<strong>in</strong>eh = 0.89 − 1.1n .We can use this relationship <strong>in</strong> the voltage equation to reduce the Hodgk<strong>in</strong>-Huxleymodel to a three-dimensional system. If, <strong>in</strong> addition, we assume that the activationk<strong>in</strong>etics of the Na + current is <strong>in</strong>stantaneous, i.e., m = m ∞ (V ), then the Hodgk<strong>in</strong>-Huxley model can be reduced to the two-dimensional systemI K<strong>in</strong>stantaneous IC ˙V{ }} {NaI{ }} { { }} L{= I − g K n 4 (V −E K ) − g Na m 3 ∞(V )(0.89−1.1n)(V −E Na ) − g L (V −E L ) ,ṅ = (n ∞ (V ) − n)/τ n (V ) ,whose solutions reta<strong>in</strong> qualitative and some quantitative agreement with the orig<strong>in</strong>alfour-dimensional Hodgk<strong>in</strong>-Huxley system; see Fig. 5.20.The first step <strong>in</strong> the analysis of any two-dimensional system is to f<strong>in</strong>d its nullcl<strong>in</strong>es.The V -nullcl<strong>in</strong>e can be found by solv<strong>in</strong>g numerically the equationI − g K n 4 (V −E K ) − g Na m 3 ∞(V )(0.89−1.1n)(V −E Na ) − g L (V −E L ) = 0for n. The nullcl<strong>in</strong>e has the familiar N-shape depicted <strong>in</strong> Fig. 5.21. Notice that ithas only one <strong>in</strong>tersection with the n-nullcl<strong>in</strong>e n = n ∞ (V ), hence there is only oneequilibrium, which is stable when I = 0. When the parameter I <strong>in</strong>creases, the equilibriumloses stability via subcritical Andronov-Hopf bifurcation, as discussed <strong>in</strong> the nextchapter. When I is sufficiently large (e.g. I = 12 <strong>in</strong> Fig. 5.21), there is a limit cycleattractor correspond<strong>in</strong>g to periodic spik<strong>in</strong>g. In Ex. 2 we discuss what happens when Ibecomes very large.


Conductance-Based Models 157150orig<strong>in</strong>al Hodgk<strong>in</strong>-Huxley model100500150100reduced Hodgk<strong>in</strong>-Huxley model10 mV1 ms5000 10 20 30 40time (ms)Figure 5.20: Action potentials <strong>in</strong> the orig<strong>in</strong>al (top) and reduced (bottom) Hodgk<strong>in</strong>-Huxley model (I = 8).I=0 I=120.8V-nullcl<strong>in</strong>en-nullcl<strong>in</strong>eK + activation, n0.70.60.50.4n-nullcl<strong>in</strong>eV-nullcl<strong>in</strong>e0.30.20 20 40 60 80 100 0 20 40 60 80 100membrane voltage, Vmembrane voltage, VFigure 5.21: Reduction of the Hodgk<strong>in</strong>-Huxley model to the (V, n)-phase plane.


158 Conductance-Based Models5.2.2 Equivalent potentialsInspired by the reduction idea of Kr<strong>in</strong>skii and Kokoz (1973), Kepler et al. (1992) suggesteda systematic method of reduc<strong>in</strong>g the complexity of conductance-based Hodgk<strong>in</strong>-Huxley-type modelsC ˙V = I − I(V, x 1 , . . . , x n )ẋ i = (m i,∞ (V ) − x i )/τ i (V ) , i = 1, . . . , n ,where x 1 , . . . , x n is a set of gat<strong>in</strong>g variables. The goal is to f<strong>in</strong>d certa<strong>in</strong> patterns orcomb<strong>in</strong>ations of the gat<strong>in</strong>g variables that can be lumped to reduce the dimension ofthe system. For example, we want to comb<strong>in</strong>e all resonant variables operat<strong>in</strong>g on asimilar time scale <strong>in</strong>to a “master” recovery variable, then do the same for amplify<strong>in</strong>gvariables.Let us convert each variable x i (t) to the equivalent potential v i (t) that satisfiesx i = m i,∞ (v i ) .In other words, the equivalent potential is the voltage which, <strong>in</strong> a voltage clamp, wouldgive the value x i when the model is at an equilibrium. Apply<strong>in</strong>g the cha<strong>in</strong> rule tov i = m −1i,∞ (x i), we express the model above <strong>in</strong> terms of equivalent potentials:C ˙V = I − I(V, m 1,∞ (v 1 ), . . . , m n,∞ (v n )) ,˙v i = (m i,∞ (V ) − m i,∞ (v i ))/(τ i (V ) m ′ i,∞(v i )) .S<strong>in</strong>ce the Boltzmann functions m i,∞ (V ) are <strong>in</strong>vertible, the denom<strong>in</strong>ators do not vanish.No approximations have been made yet; the new model is entirely equivalent to theorig<strong>in</strong>al one, it is just expressed <strong>in</strong> a different coord<strong>in</strong>ate system. The new coord<strong>in</strong>ates,however, expose many patterns among the equivalent voltage variables that were notobvious <strong>in</strong> the orig<strong>in</strong>al, gat<strong>in</strong>g coord<strong>in</strong>ate system.Kepler et al. (1992) developed an algorithm that substitutes resonant and amplify<strong>in</strong>gvariables by their weighted averages. The weights are found us<strong>in</strong>g Lagrangemultipliers and strictly local criteria aimed at preserv<strong>in</strong>g the bifurcation structure ofthe model. There is also a set of tests that <strong>in</strong>forms the user when the method is likelyto fail. The method results <strong>in</strong> a lower-dimensional system that is easier to simulate,visualize, and understand.5.2.3 Nullcl<strong>in</strong>es and I-V record<strong>in</strong>gsWe saw that the form and the position of nullcl<strong>in</strong>es provided important <strong>in</strong>formationabout the neuron dynamics, i.e., the number of equilibria, their stability, the existenceof limit cycle attractors, etc. The same <strong>in</strong>formation, <strong>in</strong> pr<strong>in</strong>ciple, can also be obta<strong>in</strong>edfrom the analysis of the neuronal current-voltage (I-V) relations. This is not a co<strong>in</strong>cidence,s<strong>in</strong>ce there is a profound relationship between nullcl<strong>in</strong>es and experimentallymeasured I-V curves.


Conductance-Based Models 159currenthold<strong>in</strong>g voltage, (mV)250<strong>in</strong>ward outward0-250I (V)I 0 (V)0V-50E K-1000 10time, msa1000500currentslow conductance, g01086I-V relationsI (V)I 0 (V)b{I+I (V)-I 0 (V)}/(V-E K )V-nullcl<strong>in</strong>eE K 0{I-I 0 (V)}/(V-E K )42-100 -80 -60 -40 -20membrane voltage, V (mV)0 20cg-nullcl<strong>in</strong>eFigure 5.22: Voltage-clamp protocol to measure <strong>in</strong>stantaneous (peak) and steadystate current-voltage (I-V) relations. (Shown simulations of the I Na,p +I K -model fromFig. 5.4b.)Let us illustrate the relationship us<strong>in</strong>g the I Na,p +I K -model, which we write <strong>in</strong> theformwhereC ˙V = I − I 0 (V ) − g(V − E K ) , (5.3)ġ = f(V, g) , (5.4)I 0 (V ) = g L (V −E L ) + g Na m ∞ (V )(V − E Na )is the <strong>in</strong>stantaneous (peak) current, and g = g K n is the slow conductance. The functionf(V, g) describes the dynamics of g, and its form is not important here. The methoddescribed below is quite general, and it can be used <strong>in</strong> many circumstances when littleis known about the neuron’s electrophysiology.In Fig. 5.22 we describe a typical voltage-clamp experiment to measure the <strong>in</strong>stantaneous(peak) and the steady-state I-V relations, denoted here as I 0 (V ) and I ∞ (V ),respectively. The hold<strong>in</strong>g voltage (Fig. 5.22a, bottom) is kept at E K and then steppedto various values V . The recorded current (Fig. 5.22a, top) typically consists of a fast(peak) component I 0 (V ) that is due to the <strong>in</strong>stantaneous activation of Na + currents,leak current, and other fast currents, and then it relaxes to the asymptotic steady statevalue I ∞ (V ). Repeat<strong>in</strong>g this experiment for various V , one can measure the I-V functionsI 0 (V ) and I ∞ (V ) depicted <strong>in</strong> Fig. 5.22b. Notice that I 0 (V ) has the N-shape witha large region of negative slope. This region corresponds to the regenerative activation


160 Conductance-Based Models1recovery variable, u0.80.60.40.2u-nullcl<strong>in</strong>eV-nullcl<strong>in</strong>e0.20.10-80 -60 -40 -20 0 20membrane potential, V (mV)a0-70 -60 -50 -40membrane potential, V (mV)bFigure 5.23: Phase portrait (a) and its magnification (b) of a typical neuronal modelhav<strong>in</strong>g voltage variable V and a recovery variable u.of the Na + current, and it is responsible for the excitability property of the neuron. Itis also responsible for the N-shape of the V -nullcl<strong>in</strong>e, as we see next.Once the I-V relations are found, we can f<strong>in</strong>d the nullcl<strong>in</strong>es of the system (5.3, 5.4).From the equationI − I 0 (V ) − g(V − E K ) = 0we can easily f<strong>in</strong>d the V -nullcl<strong>in</strong>eg = {I − I 0 (V )}/(V − E K ) (V -nullcl<strong>in</strong>e) ,which has the <strong>in</strong>verted N-shape depicted <strong>in</strong> Fig. 5.22c because I 0 (V ) does. Whilemeasur<strong>in</strong>g I ∞ (V ), we hold V long enough so that all conductances reach their steadystatevalues. The steady-state value g = g ∞ (V ) can be obta<strong>in</strong>ed from the equationI − I 0 (V ) − g(V − E K ) = −I ∞ (V ) ,which says that the asymptotic steady-state current is the sum of the steady-state fastcurrent and steady-state slow current. Thereforeg = {I + I ∞ (V ) − I 0 (V )}/(V − E K )(g-nullcl<strong>in</strong>e)depicted <strong>in</strong> Fig. 5.22c. S<strong>in</strong>ce we used the I Na,p +I K -model with parameters as <strong>in</strong>Fig. 5.4b, top, we are not surprised that the V - and g-nullcl<strong>in</strong>es found here havethe same shape and relative position as those <strong>in</strong> Fig. 5.4b, top. In Ex. 5 we furtherexplore the relationship between the I-V curves and neuronal dynamics.


Conductance-Based Models 1615.2.4 Reduction to simple modelAll models discussed <strong>in</strong> this chapter can be reduced to two-dimensional systems hav<strong>in</strong>ga fast voltage variable, V , and a slower “recovery” variable, u, with N-shaped andsigmoidal nullcl<strong>in</strong>es, respectively. The decision to fire or not to fire is made at therest<strong>in</strong>g state, which is the <strong>in</strong>tersection of the nullcl<strong>in</strong>es near the left knee, as we illustrate<strong>in</strong> Fig. 5.23a. To model the subthreshold behavior of such neurons and the <strong>in</strong>itialsegment of the up-stroke of an action potential, we need to consider only a smallneighborhood of the left knee conf<strong>in</strong>ed to the shaded square <strong>in</strong> Fig. 5.23. The rest ofthe phase space is needed only to model the peak and the down-stroke of the actionpotential. If the shape of the action potential is less important than the subthresholddynamics lead<strong>in</strong>g to this action potential, then we can reta<strong>in</strong> detailed <strong>in</strong>formationabout the left knee and its neighborhood and simplify the vector field outside theneighborhood. This approach results <strong>in</strong> a simple model capable to exhibit quite realisticdynamics, as we see <strong>in</strong> Chap. 8.Derivation via nullcl<strong>in</strong>esThe fast nullcl<strong>in</strong>e <strong>in</strong> Fig. 5.23b can be approximated by the quadratic parabolau = u m<strong>in</strong> + p(V − V m<strong>in</strong> ) 2 ,where (V m<strong>in</strong> , u m<strong>in</strong> ) is the location of the m<strong>in</strong>imum on the left knee, and p ≥ 0 is ascal<strong>in</strong>g coefficient. Similarly, the slow nullcl<strong>in</strong>e can be approximated by the straightl<strong>in</strong>eu = s(V − V 0 ) ,where s is the slope and V 0 is the V -<strong>in</strong>tercept. All these parameters can easily bedeterm<strong>in</strong>ed geometrically or analytically.Us<strong>in</strong>g these nullcl<strong>in</strong>es, we approximate the dynamics <strong>in</strong> the shaded region <strong>in</strong> Fig. 5.23by the system˙V = τ f{p(V − Vm<strong>in</strong> ) 2 − (u − u m<strong>in</strong> ) } ,˙u = τ s {s(V − V 0 ) − u} ,where the parameters τ f and τ s describe the fast and slow time scales. Because ofthe term (V − V m<strong>in</strong> ) 2 , the variable V can escape to <strong>in</strong>f<strong>in</strong>ity <strong>in</strong> a f<strong>in</strong>ite time. Thiscorresponds to the fir<strong>in</strong>g of an action potential, more precisely, to its upstroke. Tomodel the downstroke, we assume that V max is the peak value of the action potential,and we reset the state of the system(V, u) ← (V reset , u + u reset ) , when V = V max ,as if the spik<strong>in</strong>g trajectory disappears at the right edge and appears at the left edge<strong>in</strong> Fig. 5.23b. Here V reset and u reset are parameters. Appropriate re-scal<strong>in</strong>g of variables


162 Conductance-Based Modelscurrent, I0-k(v-v r )(v-v t )+b(v-v r )I (V)-k(v-v r )(v-v t )I 0 (V)v rv tmembrane potential, vFigure 5.24: The relationship betweenthe parameters of the simple model(5.7, 5.8) and the <strong>in</strong>stantaneous andsteady-state I-V relations, I 0 (V ) andI ∞ (V ), respectively.transforms the simple model <strong>in</strong>to the equivalent form˙v = I + v 2 − u if v ≥ 1, then (5.5)˙u = a(bv − u) v ← c, u ← u + d (5.6)hav<strong>in</strong>g only four dimensionless parameters.Derivation via I-V relationsThe parameters of the simple model can be derived us<strong>in</strong>g <strong>in</strong>stantaneous (peak) andsteady-state I-V relations. Let us represent the model <strong>in</strong> the follow<strong>in</strong>g equivalent formC ˙v = k(v − v r )(v − v t ) − u + I if v ≥ v peak , then (5.7)u = a{b(v − v r ) − u} v ← c, u ← u + d (5.8)where v is the membrane potential, u is the recovery current, and C is the membranecapacitance. The quadratic polynomial −k(v − v r )(v − v t ) approximates thesubthreshold part of the <strong>in</strong>stantaneous I-V relation I 0 (V ). Here, v r is the rest<strong>in</strong>g membranepotential, and v t is the <strong>in</strong>stantaneous threshold potential, as <strong>in</strong> Fig. 5.24. Thatis, <strong>in</strong>stantaneous depolarizations above v t result <strong>in</strong> spike response. The polynomial−k(v − v r )(v − v t ) + b(v − v r ) approximates the subthreshold part of the steady-stateI-V relation I ∞ (V ). When b < 0, its maximum approximates the rheobase current ofthe neuron, i.e., the m<strong>in</strong>imal amplitude of a dc-current needed to fire a cell. Its derivativewith respect to v at v = v r , i.e., b − k(v r − v t ), corresponds to the rest<strong>in</strong>g <strong>in</strong>putconductance, which is the <strong>in</strong>verse of the <strong>in</strong>put resistance. Know<strong>in</strong>g both the rheobaseand the <strong>in</strong>put resistance of a neuron, one could determ<strong>in</strong>e the parameters k and b, aswe do <strong>in</strong> Chap. 8. This method does not work when b > 0.The sum of all slow currents that modulate the spike-generation mechanism arecomb<strong>in</strong>ed <strong>in</strong> the phenomenological variable u with outward currents taken with theplus sign. The form of (5.8) ensures that u = 0 at rest, i.e., when I = 0 and v = v r .The sign of b determ<strong>in</strong>es whether u is an amplify<strong>in</strong>g (b < 0) or a resonant (b > 0)variable. In the latter case, the neuron sags <strong>in</strong> response to hyperpolarized pulses ofcurrent, peaks <strong>in</strong> response to depolarized subthreshold pulses, and produces rebound(post-<strong>in</strong>hibitory) responses. The recovery time constant is a. The spike cut-off value is


Conductance-Based Models 163v peak , and the voltage reset value is c. The parameter d describes the total amount ofoutward m<strong>in</strong>us <strong>in</strong>ward currents activated dur<strong>in</strong>g the spike and affect<strong>in</strong>g the after-spikebehavior. All these parameters can be easily fit to any particular neuron type, as weshow <strong>in</strong> Chap. 8.Review of Important Concepts• Amplify<strong>in</strong>g gat<strong>in</strong>g variables describe activation of an <strong>in</strong>ward currentor <strong>in</strong>activation of an outward current. They amplify voltage changes.• Resonant gat<strong>in</strong>g variables describe <strong>in</strong>activation of an <strong>in</strong>ward currentor activation of an outward current. They resist voltage changes.• To exhibit excitability, it is enough to have one amplify<strong>in</strong>g and oneresonant gat<strong>in</strong>g variable <strong>in</strong> a neuronal model.• Many models can be reduced to two-dimensional systems with oneequation for voltage and <strong>in</strong>stantaneous amplify<strong>in</strong>g currents, and oneequation for resonant gat<strong>in</strong>g variable.• The behavior of a two-dimensional model depends on the positionof its nullcl<strong>in</strong>es. Many models have an N-shaped V -nullcl<strong>in</strong>e and asigmoid shaped nullcl<strong>in</strong>e for the gat<strong>in</strong>g variable.• There is a relationship between nullcl<strong>in</strong>es and I-V curves.• Quite different electrophysiological models can have similar nullcl<strong>in</strong>es,and hence essentially the same dynamics.• The spike-generation mechanism of detailed electrophysiologicalmodels depends on the dynamics near the left knee of the fast V -nullcl<strong>in</strong>e, and it can be captured by a simple model (5.5, 5.6).Bibliographical NotesRichard FitzHugh pioneered the usage of phase planes and nullcl<strong>in</strong>es to study theHodgk<strong>in</strong>-Huxley model (FitzHugh 1955). Later, he suggested a simple model withN-shaped cubic V -nullcl<strong>in</strong>e and a straight-l<strong>in</strong>e slow nullcl<strong>in</strong>e, known as the FitzHugh-Nagumo model, to illustrate the mechanism of excitability of the Hodgk<strong>in</strong>-Huxleysystem. However, it was Kr<strong>in</strong>skii and Kokoz (1973) who first discovered the relationshipn(t) + h(t) ≈ const and were thus able to reduce the four-dimensional Hodgk<strong>in</strong>-Huxleymodel to a two-dimensional system. S<strong>in</strong>ce then, the phase plane analysis of neuronalmodels became standard, at least <strong>in</strong> Russian language literature.Current awareness of the geometrical methods of phase plane analysis of neuronal


164 Conductance-Based Modelsmodels is mostly due to the sem<strong>in</strong>al paper by John R<strong>in</strong>zel and Bard Ermentrout “Analysisof Neural Excitability and Oscillations”, published as a chapter <strong>in</strong> Koch and Segev’sbook Methods <strong>in</strong> Neuronal Model<strong>in</strong>g (1989, second edition <strong>in</strong> 1999). Not only did they<strong>in</strong>troduced the geometrical methods to a wide computational neuroscience audience,but also were able to expla<strong>in</strong> a number of outstand<strong>in</strong>g problems, such as the orig<strong>in</strong> ofClass 1 and 2 excitability observed by Hodgk<strong>in</strong> <strong>in</strong> 1949.R<strong>in</strong>zel and Ermentrout illustrated most of the concepts us<strong>in</strong>g the Morris-Lecar(1981) model, which is a I Ca + I K -m<strong>in</strong>imal voltage-gated model equivalent to the I Na,p+ I K -model considered above. Due to its simplicity, the Morris-Lecar model is widelyused <strong>in</strong> computational neuroscience research. This is the reason we use its analogue,the I Na,p + I K -model, throughout the book.Hutcheon and Yarom (2000) suggested to classify all currents <strong>in</strong>to amplify<strong>in</strong>g andresonant. There have been no attempts to classify various electrophysiological mechanismsof excitability <strong>in</strong> neurons, though m<strong>in</strong>imal models, such as the I Na,t -model orthe I Ca +I K(C) -model, would not surprise most researchers. The other models wouldprobably look bizarre for classical electrophysiologists, though they provide a goodopportunity to practice geometrical phase plane analysis and support FitzHugh’s observationthat N-shaped V -nullcl<strong>in</strong>e is the key characteristic of neuronal dynamics.Izhikevich (2003) took advantage of this observation and suggested the simple model(5.5, 5.6) that captures the spike-generation mechanism of many known neuronal types,see Chap. 8.Exercises1. Show that the I A -model cannot have a limit cycle attractor when I A has <strong>in</strong>stantaneousactivation k<strong>in</strong>etics. (H<strong>in</strong>t: Use Bendixson criterion.)2. When the <strong>in</strong>jected dc-current I or the Na + maximal conductance g Na <strong>in</strong> theI Na,p +I K -model have large values, the excited state (V ≈ −20 mV) becomesstable. Sketch possible <strong>in</strong>tersections of nullcl<strong>in</strong>es of the model.3. Us<strong>in</strong>g I as a bifurcation parameter, determ<strong>in</strong>e the saddle-node bifurcation diagramof• the I Na,t -model with parameters as <strong>in</strong> Fig. 5.6a,• the I A -model with parameters as <strong>in</strong> Fig. 5.14a.4. Why is g <strong>in</strong> Fig. 5.22c negative when V is hyperpolarized?5. In Fig. 5.25 we plot the currents that constitute the right-hand side of the voltageequation (5.3),I − I fast (V ) and I slow (V ) = g(V − E K ) ,on the (V, I)-plane. The curves def<strong>in</strong>e fast and slow movements of the state ofthe system. Interpret the figure. (H<strong>in</strong>t: treat the curves as “sort-of-nullcl<strong>in</strong>es”).


Conductance-Based Models 165500I slow (V) = g(V-E K )500I slow (V) = g(V-E K )400400membrane current, Imembrane current, I3002001000-100 E K -60 -20 0 20membrane voltage, V (mV)5004003002001000 15I-I fast (V)0 15I-I fast (V)I slow (V) = g(V-E K )membrane current, I membrane current, I3002001000E K-100 -60 -20 0 20membrane voltage, V (mV)5004003002001000 15I-I fast (V)0 15I-I fast (V)I slow (V) = g(V-E K )00-100 E K -60 -20 0 20membrane voltage, V (mV)-100 E K -60 -20 0 20membrane voltage, V (mV)Figure 5.25: Ex. 5: The (V, I)-phase plane of the I Na,p +I K -model (compare withFig. 5.4).6. Show that I Cl +I K -model can have oscillations. (H<strong>in</strong>t: <strong>in</strong>ject negative dc-currentso that the voltage-gated Cl − current becomes <strong>in</strong>ward/amplify<strong>in</strong>g).7. (NMDA+I K -model) Show that a neuronal model consist<strong>in</strong>g of an NMDA currentand a resonant current, say I K , can exhibit excitability and periodic spik<strong>in</strong>g.8. The Nernst potential of an ion is a function of its concentration <strong>in</strong>side/outside thecell membrane, which may change. Consider the I Na,p +E Na ([Na + ] <strong>in</strong>/out )-modeland show that it can exhibit excitability and oscillations on a slow time scale.9. Determ<strong>in</strong>e when the I A -model has a limit cycle attractor without assum<strong>in</strong>gτ h (V ) ≪ τ m (V ).10. [Ph.D.] There are Na + -gated and Cl − -gated currents besides the Ca 2+ -gatedcurrents considered <strong>in</strong> this book. In addition, the Nernst potentials may changeas concentrations of ions <strong>in</strong>side/outside the cell membrane change. This maylead to new m<strong>in</strong>imal models. Classify and study all these models.


166 Conductance-Based Models


Chapter 6BifurcationsNeuronal models can be excitable for some values of parameters, and fire spikes periodicallyfor other values. These two types of dynamics correspond to a stable equilibriumand a limit cycle attractor, respectively. When the parameters change, e.g., the <strong>in</strong>jecteddc-current <strong>in</strong> Fig. 6.1 ramps up, the models can exhibit a bifurcation — a transitionfrom one qualitative type of dynamics to another. We consider transitions away fromequilibrium po<strong>in</strong>t <strong>in</strong> Sect. 6.1 and transitions away from a limit cycle <strong>in</strong> Sect. 6.2.All these transitions can be reliably observed when only one parameter, <strong>in</strong> our caseI, changes. Mathematicians refer to such as be<strong>in</strong>g bifurcations of codimension-1. Inthis chapter we provide def<strong>in</strong>itions and examples of all codimension-1 bifurcations of anequilibrium and a limit cycle that can occur <strong>in</strong> two-dimensional systems. In Sect. 6.3 wemention some codimension-1 bifurcations <strong>in</strong> high-dimensional systems, as well as somecodimension-2 bifurcations. In the next chapter we discuss how the type of bifurcationdeterm<strong>in</strong>es a cell’s neuro-computational properties.6.1 Equilibrium (Rest State)A neuron is excitable because its rest<strong>in</strong>g state is near a bifurcation, i.e., near a transitionfrom quiescence to periodic spik<strong>in</strong>g. Typically, such a bifurcation can be revealed by<strong>in</strong>ject<strong>in</strong>g a ramp current, as we do <strong>in</strong> Fig. 6.1. The four bifurcations <strong>in</strong> the figurehave qualitatively different properties, summarized <strong>in</strong> Fig. 6.2. In this section weuse analytical and geometrical tools to understand what the differences among thebifurcations are.Recall (see Chap. 4) that an equilibrium of a dynamical system is stable if all theeigenvalues of the Jacobian matrix at the equilibrium have negative real parts. Whena parameter, say I, changes, two events can happen:1. A negative eigenvalue <strong>in</strong>creases and becomes 0. This happens at the saddle-nodebifurcation: the equilibrium disappears.2. Two complex-conjugate eigenvalues with negative real part approach the imag<strong>in</strong>aryaxis and become purely imag<strong>in</strong>ary. This happens at the Andronov-Hopf167


168 Bifurcationssaddle-node bifurcation1 mssaddle-node on <strong>in</strong>variant circle bifurcation20 mV100 ms-60 mV 60 pAsupercritical Andronov-Hopf bifurcation5 ms20 mV100 mssubcritical Andronov-Hopf bifurcation-60 mV2 nAFigure 6.1: Transitions from rest<strong>in</strong>g to tonic (periodic) spik<strong>in</strong>g occur via bifurcations ofequilibrium (marked by arrows). Saddle-node on <strong>in</strong>variant circle bifurcation: <strong>in</strong> vitrorecord<strong>in</strong>g of pyramidal neuron of rat’s primary visual cortex. Subcritical Andronov-Hopf bifurcation: <strong>in</strong> vitro record<strong>in</strong>g of bra<strong>in</strong>stem mesencephalic V neuron. The othertwo traces are simulations of the I Na,p +I K -model.


Bifurcations 169fastBifurcation of an equilibrium subthreshold amplitude frequencyoscillations of spikes of spikessaddle-node no non-zero non-zerosaddle-node on <strong>in</strong>variant circle no non-zero A √ I−I b → 0supercritical Andronov-Hopf yes A √ I−I b → 0 non-zerosubcritical Andronov-Hopf yes non-zero non-zeroFigure 6.2: Summary of codimension-1 bifurcations of an equilibrium. Here, I denotesthe amplitude of the <strong>in</strong>jected current, I b is the bifurcation value, A is a parameter thatdepends on the biophysical details.bifurcation: the equilibrium loses stability, but does not disappear.Thus, there are only two qualitative events that can happen with a stable equilibrium<strong>in</strong> a dynamical system of arbitrary dimension: It can either disappear or lose stability.Of course, there could be a third event: All eigenvalues cont<strong>in</strong>ue to have negative realparts, <strong>in</strong> which case the equilibrium rema<strong>in</strong>s stable.S<strong>in</strong>ce any equilibrium of a neuronal model is the zero of the steady-state I-V curveI ∞ (V ) (the net current at the equilibrium must be zero), analysis of the shape of theI-V curve can provide an <strong>in</strong>valuable <strong>in</strong>formation about possible bifurcations of the reststate.Two typical steady-state I-V curves are depicted <strong>in</strong> Fig. 6.3. The I-V curve <strong>in</strong>Fig. 6.3a has a region with a negative slope so that it may have 3 equilibria: the leftequilibrium is probably 1 stable, the middle is unstable, and the right equilibrium couldbe stable or unstable depend<strong>in</strong>g on the k<strong>in</strong>etics of the gat<strong>in</strong>g variables (it is stable <strong>in</strong>the one-dimensional case, i.e., when gat<strong>in</strong>g variables have <strong>in</strong>stantaneous k<strong>in</strong>etics). TheI-V curve <strong>in</strong> Fig. 6.3b is monotone. A positive (<strong>in</strong>ward) <strong>in</strong>jected dc-current I shifts theI-V curves down. This leads to the disappearance of the equilibrium <strong>in</strong> Fig. 6.3a, butnot <strong>in</strong> Fig. 6.3b. Therefore, Fig. 6.3a corresponds to the saddle-node bifurcation andFig. 6.3b to Andronov-Hopf bifurcation. When exactly the equilibrium loses stability<strong>in</strong> Fig. 6.3b cannot be <strong>in</strong>ferred from the I-V relations (for this, we need to considerthe full neuronal model). But what we can <strong>in</strong>fer is that the bifurcation cannot be ofthe saddle-node type. Surpris<strong>in</strong>gly, non-monotonic I-V curves result <strong>in</strong> saddle-nodebifurcations but do not exclude Andronov-Hopf bifurcations, as the reader is asked todemonstrate <strong>in</strong> Ex. 8. This phenomenon is relevant to the cortical pyramidal neurons1 It may be unstable; see Ex. 8


170 Bifurcationssaddle-node bifurcationAndnronov-Hopf bifurcationrest stateI(V)rest stateI(V)00noequilibrianew equilibriumI=0I>0I=0I>0-100 -50 0membrane voltage, V (mV)(a)-100 -50 0membrane voltage, V (mV)(b)Figure 6.3: Steady-state I-V curves of the I Na,p +I K -model with high-threshold (left)or low-threshold (right) K + current (parameters as <strong>in</strong> Fig. 4.1).considered <strong>in</strong> the last chapter.6.1.1 Saddle-node (fold)We provided the def<strong>in</strong>ition of a saddle-node bifurcation <strong>in</strong> one-dimensional systems <strong>in</strong>Sect. 3.3.4, and the reader is encouraged to look at that section and Fig. 4.31 beforeproceed<strong>in</strong>g further.A k-dimensional dynamical systemẋ = f(x, b) ,x ∈ R khav<strong>in</strong>g an equilibrium po<strong>in</strong>t x sn for some value of the bifurcation parameter b sn (i.e.,f(x sn , b sn ) = 0) exhibits saddle-node (also known as fold) bifurcation, if the equilibriumis non-hyperbolic with a simple zero eigenvalue, the function f is non-degenerate, andit is transversal with respect to b. The first condition is easy to check:• (Non-hyperbolicity) The Jacobian k × k matrix of partial derivatives at the equilibrium(see Sect. 4.2.2) has exactly one zero eigenvalue, and the other eigenvalueshave non-zero real parts.In general, the rema<strong>in</strong><strong>in</strong>g two conditions have complicated forms, s<strong>in</strong>ce they <strong>in</strong>volveprojections of the vector field on the center manifold, which is tangent to the eigenvectorcorrespond<strong>in</strong>g to the zero eigenvalue of the Jacobian matrix. However, there is ashortcut for conductance-based neuronal models.Let I(V, b) denote the steady-state I-V relation, which can be measured experimentally,divided by the membrane capacitance C. For example, I(V, I) = {I − I ∞ (V )}/Cwhen the <strong>in</strong>jected dc-current I is used as a bifurcation parameter. We substitute themulti-dimensional neuronal model by the one-dimensional system ˙V = I(V, b). From


Bifurcations 171I(V, b) = 0 (equilibrium condition) we f<strong>in</strong>d b = I ∞ (V ). Non-hyperbolicity conditionimplies I V (V, b) = 0, so that the bifurcation occur at the local maxima and m<strong>in</strong>ima ofI ∞ (V ). We considered all these properties <strong>in</strong> Chap. 3.• (Non-degeneracy) The second-order derivative of I(V, b sn ) with respect to V isnon-zero, that is,a = 1 ∂ 2 I(V, b sn )≠ 0 (at V = V2 ∂V 2 sn ) . (6.1)That is, the piece of the I-V curve, I ∞ (V ), at the bifurcation po<strong>in</strong>t, V sn , lookslike the square parabola.• (Transversality) Function I(V, b) is non-degenerate with respect to the bifurcationparameter b; that is,c = ∂I(V sn, b)∂b≠ 0 (at b = b sn ) .This condition is always satisfied when the <strong>in</strong>jected dc-current I is the bifurcationparameter, because ∂I/∂b = ∂I/∂I = 1/C.The saddle-node bifurcation has codimension 1 because only one condition (non-hyperbolicity)<strong>in</strong>volves strict equality (“=”), and the other two <strong>in</strong>volve <strong>in</strong>equalities (“≠”). The dynamicsof multi-dimensional neuronal systems near a saddle-node bifurcation can bereduced to that of the topological normal form˙V = c(b − b sn ) + a(V − V sn ) 2 , (6.2)where V is the membrane voltage, and a and c are def<strong>in</strong>ed above. In the context ofneuronal models, this equation with an after-spike resett<strong>in</strong>g is called the quadratic<strong>in</strong>tegrate-and-fire neuron, which we discuss <strong>in</strong> Chapters 3 and 8.Example: The I Na,p +I K -modelLet us use the I Na,p +I K -model (4.1, 4.2) with high-threshold K + current to illustratethese conditions. The saddle-node bifurcation occurs when the V -nullcl<strong>in</strong>e touchesthe n-nullcl<strong>in</strong>e, as <strong>in</strong> Fig. 6.4. Solv<strong>in</strong>g the equations numerically, we f<strong>in</strong>d that thisoccurs when I sn = 4.51 and (V sn , n sn ) = (−61, 0.0007). The Jacobian matrix at theequilibrium,( )0.0435 −290L =,0.00015 −1has two eigenvalues λ 1 = 0 and λ 2 = −0.9565, with correspond<strong>in</strong>g eigenvectors( )( )11v 1 =and v0.000152 =,0.0034


172 Bifurcations0.5v 2V-nullcl<strong>in</strong>eK + activation, n0.40.30.20.001v 10-63 -62 -61 -60 -59 -580.10n-nullcl<strong>in</strong>e-80 -70 -60 -50 -40 -30 -20 -10 0membrane voltage, V (mV)Figure 6.4: Saddle-node bifurcation <strong>in</strong> the I Na,p +I K -model (4.1, 4.2) with highthresholdK + current (parameters as <strong>in</strong> Fig. 4.1a) and I = 4.51.-50membrane voltage, V (mV)-55-60-65-70normal formI Na +I K -model-75-15 -10 -5 0 5 10<strong>in</strong>jected dc-current, IFigure 6.5: Bifurcation diagrams of the topological normal form (6.3) and the I Na,p +I K -model (4.1, 4.2).


Bifurcations 173(a) saddle-node bifurcationlimitcyclenodesaddlesaddle-node(b) saddle-node on <strong>in</strong>variant circle (SNIC) bifurcation<strong>in</strong>variant circlenode saddle saddle-nodeFigure 6.6: Two types of saddle-node bifurcation.depicted <strong>in</strong> the <strong>in</strong>set <strong>in</strong> Fig. 6.4. (It is easy to check that Lv 1 = 0 and Lv 2 = −0.9565v 2 .)The non-degeneracy and transversality conditions yields a = 0.1887 and c = 1, so thatthe topological normal form for the I Na,p +I K -model is˙V = (I − 4.51) + 0.1887(V + 61) 2 , (6.3)which can be solved analytically. The correspond<strong>in</strong>g bifurcation diagrams are depicted<strong>in</strong> Fig. 6.5. There is no surprise that there is a fairly good match when I is near thebifurcation value.6.1.2 Saddle-node on <strong>in</strong>variant circleAs its name stands, saddle-node on <strong>in</strong>variant circle bifurcation (also known as SNICor SNLC bifurcation) is a standard saddle-node bifurcation described above with anadditional caveat: it occurs on an <strong>in</strong>variant circle, compare Fig. 6.6a and b. Here,the <strong>in</strong>variant circle consists of two trajectories connect<strong>in</strong>g the node and the saddle,called heterocl<strong>in</strong>ic trajectories. It is called <strong>in</strong>variant because any solution start<strong>in</strong>g onthe circle rema<strong>in</strong>s on the circle. As the saddle and node coalesce, the small trajectoryshr<strong>in</strong>ks and the large heterocl<strong>in</strong>ic trajectory becomes a homocl<strong>in</strong>ic <strong>in</strong>variant circle, i.e.,orig<strong>in</strong>at<strong>in</strong>g and term<strong>in</strong>at<strong>in</strong>g at the same po<strong>in</strong>t. When the po<strong>in</strong>t disappears, the circlebecomes a limit cycle.


174 BifurcationsBoth types of the bifurcation can occur <strong>in</strong> the I Na,p +I K -model as we show <strong>in</strong> Fig. 6.7.The difference between top and bottom of the figure is the time constant τ(V ) of theK + current. S<strong>in</strong>ce the K + current has high threshold, the time constant does not affectdynamics at rest, but it makes a huge difference when action potential is generated.If the current is fast (top), it activates dur<strong>in</strong>g the upstroke thereby decreas<strong>in</strong>g theamplitude of action potential, and deactivates dur<strong>in</strong>g the downstroke thereby result<strong>in</strong>g<strong>in</strong> overshoot and another action potential. In contrast, slower K + current (bottom)does not have time to deactivate dur<strong>in</strong>g the downstroke, thereby result<strong>in</strong>g <strong>in</strong> undershoot(short after-hyperpolarization), with V go<strong>in</strong>g below the rest<strong>in</strong>g state.From the geometrical po<strong>in</strong>t of view, the phase portraits <strong>in</strong> Fig. 6.6b and <strong>in</strong> Fig. 6.7,bottom, have the same topological structure: there is a homocl<strong>in</strong>ic trajectory (an<strong>in</strong>variant circle) that orig<strong>in</strong>ates at the saddle-node po<strong>in</strong>t, leaves its small neighborhood(to fire an action potential), then reenters the neighborhood aga<strong>in</strong>, and term<strong>in</strong>atesat the saddle-node po<strong>in</strong>t. This homocl<strong>in</strong>ic trajectory is a limit cycle attractor with<strong>in</strong>f<strong>in</strong>ite period, which corresponds to fir<strong>in</strong>g with zero frequency. This and other neurocomputationalfeatures of saddle-node bifurcations are discussed <strong>in</strong> the next chapter.Below we only explore how the frequency of oscillation depends on the bifurcationparameter, e.g., on the <strong>in</strong>jected dc-current I.A remarkable fact is that we can estimate the frequency of the large-amplitude limitcycle attractor by consider<strong>in</strong>g a small neighborhood of the saddle-node po<strong>in</strong>t. Indeed,a trajectory on the limit cycle generates a fast spike from po<strong>in</strong>t B to A <strong>in</strong> Fig. 6.8 andthen slowly moves from A to B (shaded region <strong>in</strong> the figure) because the vector field(the velocity) <strong>in</strong> the neighborhood between A and B is very small. The duration of thestereotypical action potentials, denoted here as T 1 , is relatively constant and does notdepend much on the <strong>in</strong>jected current I. In contrast, the time spent <strong>in</strong> the neighborhood(A, B) depends significantly on I. S<strong>in</strong>ce the behavior <strong>in</strong> the neighborhood is describedby the topological normal form (6.2), we can estimate the time the trajectory spendsthere <strong>in</strong> terms of the parameters a, b and c (see Ex. 3). This yieldsT 2 =π√ac(b − bsn ) ,where the parameters a, b, and c are those def<strong>in</strong>ed <strong>in</strong> the previous section. So theperiod of one oscillation is T = T 1 + T 2 .In Fig. 6.8, top, we illustrate the accuracy of this estimation us<strong>in</strong>g the I Na,p +I K -model, whose topological normal form (6.3) was derived earlier. The duration of theaction potential is T 1 = 4.7 ms, and the duration of time the voltage variable spends<strong>in</strong> the shaded neighborhood (A,B) (here −61 ± 11 mV) is approximated byT 2 =π√0.1887(I − 4.51)(ms)The analytical curveω = 1000T 1 + T 2(Hz)


Bifurcations 175fast K+ current0.60.5n-nullcl<strong>in</strong>eK + activation variable, n0.40.30.2V-nullcl<strong>in</strong>e0.10saddle-node0.60.5slow K+ currentn-nullcl<strong>in</strong>e0.4K + activation variable, n0.30.20.1V-nullcl<strong>in</strong>e<strong>in</strong>variant circle0saddle-nodeon <strong>in</strong>variant circle-80 -70 -60 -50 -40 -30 -20 -10 0 10 20membrane voltage, V (mV)Figure 6.7: Saddle-node bifurcation <strong>in</strong> the I Na,p +I K -model with high-threshold K +current can be off the limit cycle (top) or on the <strong>in</strong>variant circle (bottom). Parametersas <strong>in</strong> Fig. 4.1a with τ(V ) = 0.152 (top) or τ(V ) = 1 (bottom).


176 BifurcationsV-nullcl<strong>in</strong>e0.5K + gat<strong>in</strong>g variable, n0.40.30.210080604020frequency, Hz1000/T 2theoreticalnumerical0.1A04.5 5 5.5<strong>in</strong>jected dc-current, I0n-nullcl<strong>in</strong>eB-80 -70 -60 -50 -40 -30 -20 -10 0 10 20membrane voltage, V (mV)-30cut spikecut spikemembrane voltage, V (mV)-40-50-60-70BA-80T 1tan tT 20 10 20 30 40 50 60 70 80 90 100time, t (ms)Figure 6.8: The I Na,p +I K -model can a fire periodic tra<strong>in</strong> of action potentials witharbitrary small frequency when it is near a saddle-node on <strong>in</strong>variant circle bifurcation.The trajectory moves fast from po<strong>in</strong>t B to A (a spike) and slowly <strong>in</strong> the shaded regionfrom po<strong>in</strong>t A to B.


Bifurcations 177matches numerically the found frequency of oscillation (Fig. 6.8, top) <strong>in</strong> a fairly broadfrequency range. For comparison, we plot the curve 1000/T 2 to show that neglect<strong>in</strong>gthe duration of the spike, T 1 , can be justified only when I is very close to the bifurcationpo<strong>in</strong>t.6.1.3 Supercritical Andronov-HopfIf a neuronal model has a monotonic steady-state I-V relation, a saddle-node bifurcationcannot occur. The rest<strong>in</strong>g state <strong>in</strong> such a model does not disappear, but it losesstability, typically via an Andronov-Hopf (sometimes called Hopf) bifurcation. Thelost of stability is accompanied either by an appearance of a stable limit cycle (supercriticalAndronov-Hopf) or by a disappearance of an unstable limit cycle (subcriticalAndronov-Hopf).Let us consider a two-dimensional system˙v = F (v, u, b)˙u = G(v, u, b)(6.4)and suppose that (v, u) = (0, 0) is an equilibrium when the bifurcation parameterb = 0, that is, F (0, 0, 0) = G(0, 0, 0) = 0. This system undergoes an Andronov-Hopfbifurcation at the equilibrium if the follow<strong>in</strong>g three conditions are satisfied:• (Non-hyperbolicity) The Jacobian 2 × 2 matrix of partial derivatives at the equilibrium(see Sect. 4.2.2), ( )Fv FL =u,G v G uhas a pair of purely imag<strong>in</strong>ary eigenvalues, ±iω ∈ C with ω ≠ 0. That is,tr L = F v + G u = 0 and ω 2 = det L = F v G u − F u G v > 0 at v = u = b = 0.The l<strong>in</strong>ear change of variablesconverts (6.4) <strong>in</strong>to the formwhere functionsv = x and F u u = −F v x − ωy (6.5)ẋ = −ωy + f(x, y)ẏ = ωx + g(x, y) ,f(x, y) = F (v, u) + ωy and g(x, y) = −(F v F (v, u) + F u G(v, u))/ω − ωxhave no l<strong>in</strong>ear terms <strong>in</strong> x and y. Now we are ready to state the other two conditions:• (Non-degeneracy) The parameter(6.6)a = 1 16 {f xxx + f xyy + g xxy + g yyy }+ 116ω {f xy(f xx + f yy ) − g xy (g xx + g yy ) − f xx g xx + f yy g yy }is non-zero.(6.7)


178 Bifurcationsrstable limit cycleunstable limitcyclerstableequilibriumunstable equilibriumstable equilibriumunstableequilibriumstable limit cyclecunstable limitcyclecSupercritical (a0)Figure 6.9: Andronov-Hopf bifurcation: A stable equilibrium becomes unstable <strong>in</strong>system (6.8, 6.9).• (Transversality) Let c(b) ± iω(b) denote the complex-conjugate eigenvalues of theJacobian matrix of (6.4) for b near 0, with c(0) = 0 and ω(0) = ω The real part,c(b), must be non-degenerate with respect to b, that is, c ′ (0) ≠ 0.The Andronov-Hopf bifurcation has codimension one, s<strong>in</strong>ce only one condition <strong>in</strong>volvesstrict equality (tr L = 0), and the other two <strong>in</strong>volve <strong>in</strong>equalities (“≠”).The sign of a determ<strong>in</strong>es the type of the Andronov-Hopf bifurcation, depicted <strong>in</strong>Fig. 6.9:• Supercritical Andronov-Hopf bifurcation occurs when a < 0. It corresponds to astable limit cycle appear<strong>in</strong>g from a stable equilibrium.• Subcritical Andronov-Hopf bifurcation occurs when a > 0. It corresponds to anunstable limit cycle shr<strong>in</strong>k<strong>in</strong>g to a stable equilibrium.F<strong>in</strong>d<strong>in</strong>g a <strong>in</strong> applications could be challeng<strong>in</strong>g. A few useful examples are considered<strong>in</strong> Exercises 14-18.Any system undergo<strong>in</strong>g an Andronov-Hopf bifurcation can be reduced by a changeof variables to the topological normal form (see also Ex. 4)ṙ = c(b)r + ar 3 , (6.8)˙ϕ = ω(b) + dr 2 , (6.9)where r ≥ 0 is the amplitude (radius), and ϕ is the phase (angle) of oscillation, as <strong>in</strong>Fig. 6.10, and a, b, c(b), and ω(b) as above.


Bifurcations 179r0ϕFigure 6.10: Polar coord<strong>in</strong>ates: r is the amplitude(radius) and ϕ is the phase (angle) of oscillation.The function c(b) <strong>in</strong> the normal form (6.8,6.9) determ<strong>in</strong>es the stability of the equilibriumr = 0 correspond<strong>in</strong>g to a non-oscillatory state. (Stable for c < 0 and unstablefor c > 0, regardless of the value of a). The function ω(b) determ<strong>in</strong>es the frequency ofdamped or susta<strong>in</strong>ed oscillations around this state. The parameter d describes how thefrequency of oscillation depends on its amplitude. A state-dependent change of timecan remove the term dr 2 from (6.9) (Kuznetsov 1995), so many assume that d = 0 tostart with.Example: The I Na,p +I K -modelLet us use the I Na,p +I K -model with low-threshold K + current <strong>in</strong> Fig. 6.11 to illustratethe three conditions above. As the magnitude of the <strong>in</strong>jected dc-current I <strong>in</strong>creases,the equilibrium loses stability and gives birth to a stable limit cycle with grow<strong>in</strong>gamplitude. Us<strong>in</strong>g simulations we f<strong>in</strong>d that the bifurcation occurs when I ah = 14.66and (V ah , n ah ) = (−56.5, 0.09). The Jacobian matrix at the equilibrium,(L =1 −3350.0166 −1has a pair of complex conjugate eigenvalues ±2.14i, so the non-hyperbolicity conditionis satisfied. Next, we f<strong>in</strong>d numerically (<strong>in</strong> Fig. 6.12 or analytically <strong>in</strong> Ex. 9) that theeigenvalues at the equilibrium can be approximated byc(I) + ω(I)i ≈ 0.03{I − 14.66} ± (2.14 + 0.04{I − 14.66})i<strong>in</strong> a neighborhood of the bifurcation po<strong>in</strong>t I = 14.66. S<strong>in</strong>ce the slope of c(I) is non-zero,the transversality condition is also satisfied. Us<strong>in</strong>g Ex. 17 we f<strong>in</strong>d that a = −0.0026 andd = −0.0029, so that the non-degeneracy condition is also satisfied, and the bifurcationis of the supercritical type. The correspond<strong>in</strong>g topological normal form isṙ = 0.03{I − 14.66}r − 0.0026r 3 ,˙ϕ = (2.14 + 0.04{I − 14.66}) − 0.0029r 2 .To analyze the normal form we consider the r-equation and neglect the phase variableϕ. Fromr(c(b) + ar 2 ) = 0),


180 BifurcationsnVstable limit cyclesI=40stableunstableII=30max/m<strong>in</strong> of oscillations ofmembrane potential, mV0-10-20-30-40-50-60-70-80supercriticalAndronov-Hopfbifurcationstablestable cyclesunstable equiliblia0 20 40 60 80<strong>in</strong>jected dc-current, II=20I=25I=15I=10I=0I=-100.70.60.50.4supercritical Andronov-Hopfbifurcation at I=14.66Ceigenvalues c + iωV-nullcl<strong>in</strong>e0.30.20.1n-nullcl<strong>in</strong>eimag<strong>in</strong>ary part, ω00-80 -60 -40 -20 0 20membrane voltage, V (mV)stable unstable0real part, cFigure 6.11: Supercritical Andronov-Hopf bifurcation <strong>in</strong> the I Na,p +I K -model with lowthresholdK + current: As a bifurcation parameter I <strong>in</strong>creases, an equilibrium losesstability and gives birth to a stable limit cycle with grow<strong>in</strong>g amplitude. Parameters as<strong>in</strong> Fig. 4.1b.


Bifurcations 1813210eigenvalues = c(I) + ω(I)il<strong>in</strong>earization1510514.66 14.66-1000 10 20 30 0 10 20 30 0 10 20 30<strong>in</strong>jected dc-current, I <strong>in</strong>jected dc-current, I <strong>in</strong>jected dc-current, Inumericalω(I)c(I)membrane voltage, (mV)amplitudenumericaltheoreticalfrequency, 2π/period (rads/ms)2.521.510.5frequencytheoreticalnumericalFigure 6.12: Supercritical Andronov-Hopf bifurcation <strong>in</strong> the I Na,p +I K -model with lowthresholdK + current (see Fig. 6.11). Dots — numerical simulation of the full model,cont<strong>in</strong>uous curves — analytical results us<strong>in</strong>g the topological normal form (6.8, 6.9).we conclude that r = 0 is an equilibrium for any value of c(b). S<strong>in</strong>ce(c(b)r + ar 3 ) r = c(b) at r = 0,the equilibrium is stable for c(b) < 0 and unstable for c(b) > 0, as we illustrate <strong>in</strong>Fig. 6.13. Indeed, the rest state <strong>in</strong> the I Na,p +I K -model is stable when I < 14.66 andunstable when I > 14.66.When c(b) > 0, the normal form has a family of stable periodic solutions withamplituder = √ c(b)/|a| and (frequency) = ω(b) + d c(b)/|a| .Hence, the I Na,p +I K -model has a family of periodic attractors withandr = √ 0.03{I − 14.66}/0.0026(frequency) = (2.14 + 0.04{I − 14.66}) − 0.0029 · 0.03{I − 14.66}/0.0026 ,depicted <strong>in</strong> Fig. 6.12. We see that the topological normal form describes not onlyqualitatively but also quantitatively the full I Na,p +I K -model near the Andronov-Hopfbifurcation.6.1.4 Subcritical Andronov-HopfNeuronal models with monotonic steady-state I-V relations can also exhibit subcriticalAndronov-Hopf bifurcations, as we illustrate <strong>in</strong> Fig. 6.16 us<strong>in</strong>g the I Na,p +I K -modelhav<strong>in</strong>g low-threshold K + current and a steep activation curve for Na + current. The


¡ ¡¡ ¡¡ ¡ ¡¡ ¡ ¡182 Bifurcations4r'c(b)


Bifurcations 1831K+ activation variable, n0.5stablelimit cycleunstable limitcyclen-nullcl<strong>in</strong>eV-nullcl<strong>in</strong>estable equilibrium (rest)0-80 -70 -60 -50 -40 -30 -20 -10 0 10membrane voltage, V (mV)Figure 6.14: Phase portrait of the I Na,p +I K -model: Unstable limit cycle (dashed circle)is often surrounded by a stable one (solid circle) <strong>in</strong> two-dimensional neuronal models.F<strong>in</strong>ally, notice that there is always a bistability, i.e., co-existence, of the rest<strong>in</strong>gattractor and some other attractor near a subcritical Andronov-Hopf bifurcation <strong>in</strong>2-dimensional conductance-based models, as <strong>in</strong> Fig. 6.14 (<strong>in</strong> non-neural models, thetrajectories could go to <strong>in</strong>f<strong>in</strong>ity and there need not be bistability). The bistabilitymust also be present at the saddle-node bifurcation of an equilibrium, but may or maynot be present at the saddle-node on <strong>in</strong>variant circle or at a supercritical Andronov-Hopf bifurcation.Delayed loss of stabilityIn Fig. 6.17a we <strong>in</strong>ject a ramp of current <strong>in</strong>to the I Na,p +I K -model to drive it slowlythrough the subcritical Andronov-Hopf bifurcation po<strong>in</strong>t I ≈ 48.75 (see Fig. 6.16). Wechoose the ramp so that the bifurcation occurs exactly at t = 100. Even though thefocus equilibrium is unstable for t > 100, the membrane potential rema<strong>in</strong>s near -50mV as if the equilibrium were still stable. This phenomenon, discovered by Shishkova(1973), is called delayed loss of stability. It is ubiquitous <strong>in</strong> simulations of smoothdynamical systems near subcritical or supercritical Andronov-Hopf bifurcations.The mechanism of delayed loss of stability is quite simple. The state of the systemis attracted to the stable focus while t < 100. Even though the focus loses stabilityat t = 100, the state of the system is <strong>in</strong>f<strong>in</strong>itesimally close to the equilibrium, so ittakes a long time to diverge from it. The longer the convergence to the equilibrium,the longer the divergence from it, hence the noticeable delay. The delay has an upperbound that depends on the smoothness of the dynamical system (Nejshtadt 1985).It can be shortened or even reversed (advanced loss of stability) by weak noise thatis always present <strong>in</strong> neurons. This may expla<strong>in</strong> why the delay has never been seen


¡ ¡ ¡¡ ¡ ¡¡ ¡ ¡¡ ¡ ¡184 Bifurcations4r'c(b)


Bifurcations 185unstable limit cyclesnVI=49stableunstableII=48.75max/m<strong>in</strong> of oscillations ofmembrane potential, mV0-20-40-60-80u nstablestablelimit cyclesunstable45 50 55<strong>in</strong>jected current, II=48I=47I=45subcritical Andronov-Hopfbifurcation at I=48.8I=43Ceigenvalues c + iω10.80.6V-nullcl<strong>in</strong>eimag<strong>in</strong>ary part, ω00.4stable unstable0real part, c0.20n-nullcl<strong>in</strong>e-80 -60 -40 -20 0 20membrane voltage, V (mV)Figure 6.16: Subcritical Andronov-Hopf bifurcation <strong>in</strong> the I Na,p +I K -model: As thebifurcation parameter I <strong>in</strong>creases, an unstable limit cycle (dashed circle; see alsoFig. 6.14) shr<strong>in</strong>ks to an equilibrium and makes it lose stability. Parameters as <strong>in</strong>Fig. 4.1b except g L = 1, g Na = g K = 4, and the Na + activation function has V 1/2 = −30mV and k = 7.


186 Bifurcationsmembrane potential, V (mV)0-20-40-60(a)stable limitcycleunstable limit cyclestablesubcriticalAndronov-Hopfbifurcationunstabledelayedloss of stabilitymembrane potential, V (mV)0-20-40-60(b)stable limitnoise-<strong>in</strong>ducedsusta<strong>in</strong>ed oscillationscycleunstable limit cyclesubcriticalAndronov-Hopfbifurcation<strong>in</strong>jectedcurrent, I(t)60400 20 40 60 80 100 120 140 160 180 200time, tFigure 6.17: Delayed loss of stability (a) and noise-<strong>in</strong>duced susta<strong>in</strong>ed oscillations (b)near subcritical Andronov-Hopf bifurcation. Shown are simulations of the I Na,p +I K -model with parameters as <strong>in</strong> Fig. 6.16 and the same <strong>in</strong>itial conditions. Small conductancenoise is added <strong>in</strong> (b) to unmask oscillations.wan<strong>in</strong>g rhythmic activity seen <strong>in</strong> the figure (see also Ex. 3 <strong>in</strong> Chap. 7). In Chapters 8and 9 we present many examples of noise-<strong>in</strong>duced susta<strong>in</strong>ed oscillations <strong>in</strong> biologicalneurons, and <strong>in</strong> Chap. 7 we study their neuro-computational properties.6.2 Limit Cycle (Spik<strong>in</strong>g State)In the previous section we considered all codimension-1 bifurcations of equilibria, whichtypically correspond to transitions from rest<strong>in</strong>g to spik<strong>in</strong>g states <strong>in</strong> neuronal models.Below we consider all codimension-1 bifurcations of limit cycle attractors on a phaseplane. These bifurcations typically correspond to transitions from repetitive spik<strong>in</strong>g torest<strong>in</strong>g behavior, as we illustrate <strong>in</strong> Fig. 6.18, and they will be important <strong>in</strong> Chap. 9where we consider burst<strong>in</strong>g dynamics.


Bifurcations 187saddle-node on <strong>in</strong>variant circle bifurcation10 mV50 ms-30 mV700 pAsupercritical Andronov-Hopf bifurcation20 mV100 ms-60 mV-10 mV180 pAfold limit cycle bifurcation20 mV10 ms-40 mVsaddle homocl<strong>in</strong>ic orbit bifurcation20 mV50 msFigure 6.18: Transitions from tonic (periodic) spik<strong>in</strong>g to rest<strong>in</strong>g occurs via bifurcationsof a limit cycle attractors (marked by arrows). Saddle-node on <strong>in</strong>variant circlebifurcation: record<strong>in</strong>g of layer 5 pyramidal neuron <strong>in</strong> rat’s visual cortex. SupercriticalAndronov-Hopf bifurcation: excitation block <strong>in</strong> pyramidal neuron of rat’s visual cortex.Fold limit cycle bifurcation: bra<strong>in</strong>stem mesencephalic V neuron of rat. Saddlehomocl<strong>in</strong>ic orbit bifurcation: neuron <strong>in</strong> pre-Boltz<strong>in</strong>ger complex of rat bra<strong>in</strong>stem (dataprovided by C.A. Del Negro and J.L. Feldman).


188 BifurcationsBifurcation of a limit cycle attractor amplitude frequencysaddle-node on <strong>in</strong>variant circle non-zero A √ I−I b → 0supercritical Andronov-Hopf A √ I−I b → 0 non-zerofold limit cycle non-zero non-zerosaddle homocl<strong>in</strong>ic orbit non-zero−Aln |I−I b | →0Figure 6.19: Summary of codimension-1 bifurcations of a limit cycle attractor on aplane. Here, I denotes the amplitude of the <strong>in</strong>jected current, I b is the bifurcationvalue, A is a parameter that depends on the biophysical details.In Fig. 6.19 we summarize how the bifurcations affect a periodic attractor. Saddlenodeon <strong>in</strong>variant circle and saddle homocl<strong>in</strong>ic orbit bifurcations <strong>in</strong>volve homocl<strong>in</strong>ictrajectories hav<strong>in</strong>g <strong>in</strong>f<strong>in</strong>ite period (zero frequency). They result <strong>in</strong> oscillations withdrastically <strong>in</strong>creas<strong>in</strong>g <strong>in</strong>terspike <strong>in</strong>tervals as the system approaches the bifurcation state(see Fig. 6.18).In contrast, supercritical Andronov-Hopf bifurcation results <strong>in</strong> oscillations withvanish<strong>in</strong>g amplitude, as one can clearly see <strong>in</strong> Fig. 6.18. If neither the frequency northe amplitude vanishes, then the bifurcation is of the fold limit cycle type. Indeed,the amplitude and the <strong>in</strong>terspike period are constant before the arrow <strong>in</strong> Fig. 6.18correspond<strong>in</strong>g to the fold limit cycle bifurcation. Damped small-amplitude oscillationafter the arrow occurs because of oscillatory convergence to the equilibrium.We start with a brief review of the saddle-node on <strong>in</strong>variant circle and the supercriticalAndronov-Hopf bifurcations, which we considered <strong>in</strong> detail <strong>in</strong> Sect. 6.1. Thesebifurcations can expla<strong>in</strong> not only transitions from rest to spik<strong>in</strong>g but also transitionsfrom spik<strong>in</strong>g to rest states. Then, we consider fold limit cycle and saddle homocl<strong>in</strong>icorbit bifurcations.6.2.1 Saddle-node on <strong>in</strong>variant circleA stable limit cycle can disappear via a saddle-node on <strong>in</strong>variant circle bifurcationas depicted <strong>in</strong> Fig. 6.20. The necessary condition for such a bifurcation is that thesteady-state I-V relation is not monotonic.We considered this bifurcation <strong>in</strong> Sect. 6.1.2 as a bifurcation from an equilibriumto a limit cycle; that is from left to right <strong>in</strong> Fig. 6.20. Now consider it from rightto left: As a bifurcation parameter changes, e.g., the <strong>in</strong>jected dc-current I decreases,a stable limit cycle (circle <strong>in</strong> Fig. 6.20, right) disappears because there is a saddle-


Bifurcations 189bifurcation parameter changes<strong>in</strong>variant circle<strong>in</strong>variant circlestable limit cyclenode saddle saddle-nodeFigure 6.20: Saddle-node on <strong>in</strong>variant circle (SNIC) bifurcation of a limit cycle attractor.bifurcation parameter changesstable equilibriumstable limit cycleunstable equilibriumFigure 6.21: Supercritical Andronov-Hopf bifurcation of a limit cycle attractor.node bifurcation (Fig. 6.20, center) that breaks the cycle and gives birth to a pairof equilibria– stable node and unstable saddle (Fig. 6.20, left). After the bifurcation,the limit cycle becomes an <strong>in</strong>variant circle consist<strong>in</strong>g of a union of two heterocl<strong>in</strong>ictrajectories.Depend<strong>in</strong>g on the direction of change of a bifurcation parameter, the saddle-nodeon <strong>in</strong>variant circle bifurcation can expla<strong>in</strong> either appearance or disappearance of alimit cycle attractor. In either case, the amplitude of the limit cycle rema<strong>in</strong>s relativelyconstant but its period becomes <strong>in</strong>f<strong>in</strong>ite at the bifurcation po<strong>in</strong>t because the cyclebecomes a homocl<strong>in</strong>ic trajectory to the saddle-node equilibrium (Fig. 6.20, center). Aswe showed <strong>in</strong> Sect. 6.1.2 (see Fig. 6.8), the frequency of oscillation scales as √ I − I bwhen the bifurcation parameter approaches the bifurcation value I b .6.2.2 Supercritical Andronov-HopfA stable limit cycle can shr<strong>in</strong>k to a po<strong>in</strong>t via supercritical Andronov-Hopf bifurcation<strong>in</strong> Fig. 6.21, which we considered <strong>in</strong> Sect. 6.1.3. Indeed, as the bifurcation parameterchanges, e.g., the <strong>in</strong>jected dc-current I <strong>in</strong> Fig. 6.11 decreases, the amplitude of thelimit cycle attractor vanishes, and the cycle becomes just a stable equilibrium. Aswe showed <strong>in</strong> Sect. 6.1.3 (see Fig. 6.12), the amplitude scales as √ I − I b when the


190 Bifurcationsstableunstablelimit cyclelimitcyclefoldlimitcycleFigure 6.22: Fold limit cycle bifurcation: A stable and an unstable limit cycle approachand annihilate each other.bifurcation parameter approaches the bifurcation value I b .6.2.3 Fold limit cycleA stable limit cycle can appear (or disappear) via the fold limit cycle bifurcationdepicted <strong>in</strong> Fig. 6.22. Let us consider the figure from left to right, which corresponds tothe disappearance of the limit cycle, and hence to the disappearance of periodic spik<strong>in</strong>gactivity. As the bifurcation parameter changes, the stable limit cycle is approached byan unstable one, they coalesce and annihilate each other. At the po<strong>in</strong>t of annihilation,there is a periodic orbit, but it is neither stable nor unstable. More precisely, it is stablefrom the side correspond<strong>in</strong>g to the stable cycle (outside <strong>in</strong> Fig. 6.22), and unstable fromthe other side (<strong>in</strong>side <strong>in</strong> Fig. 6.22). This periodic orbit is referred to as be<strong>in</strong>g a fold(also known as a saddle-node) limit cycle, and it is analogous to the fold (saddle-node)equilibrium studied <strong>in</strong> Sect. 6.1. Consider<strong>in</strong>g Fig. 6.22 from right to left expla<strong>in</strong>s howa stable limit cycle can appear seem<strong>in</strong>gly out of nowhere: As a bifurcation parameterchanges, a fold limit cycle appears, which then bifurcates <strong>in</strong>to a stable limit cycle andan unstable one.Fold limit cycle bifurcation can occur <strong>in</strong> the I Na,p +I K -model hav<strong>in</strong>g low-thresholdK + current, as we demonstrate <strong>in</strong> Fig. 6.23. The top phase portrait correspond<strong>in</strong>g toI = 43 is the same as the one <strong>in</strong> Fig. 6.16. In that figure we studied how the equilibriumloses stability via subcritical Andronov-Hopf bifurcation, which occurs when anunstable limit cycle shr<strong>in</strong>ks to a po<strong>in</strong>t. We never questioned where the unstable limitcycle came from. Neither were we concerned with the existence of a large-amplitudestable limit cycle correspond<strong>in</strong>g to the periodic spik<strong>in</strong>g state. In Fig. 6.23 we studythis problem. We decrease the bifurcation parameter I to see what happens with thelimit cycles. As I approaches the bifurcation value 42.18, the unstable and stable limitcycles approach and annihilate each other. When I is less than the bifurcation value,there are no periodic orbits, only one stable equilibrium correspond<strong>in</strong>g to the rest<strong>in</strong>gstate.Notice that the fold limit cycle bifurcation expla<strong>in</strong>s how (un)stable limit cycles


Bifurcations 191I=43limitcyclesstableI=42.5unstable limitcyclesfold limit cycleI=42.35foldlimitcycleI=42.181I=42V-nullcl<strong>in</strong>em<strong>in</strong>/max of oscillation ofmembrane potential, mV0-20-40-60bifurcationstablelimit cyclesunstable limit cyclesstable equilibria0.5n-nullcl<strong>in</strong>e-8040 45 50<strong>in</strong>jected current, I0-80 -60 -40 -20 0membrane voltage, mVFigure 6.23: Fold limit cycle bifurcation <strong>in</strong> the I Na,p +I K -model. As the bifurcationparameter I decreases, the stable and unstable limit cycles approach and annihilateeach other. Parameters as <strong>in</strong> Fig. 6.16.


192 Bifurcationsamplitude (max-m<strong>in</strong>), mV8070605040302010fold limit cyclebifurcationstable limit cyclesunstable limit cyclessubcriticalAndronov-Hopfbifurcation040 41 42 43 44 45 46 47 48 49 50<strong>in</strong>jected dc-current, IFigure 6.24: Bifurcation diagram of the I Na,p +I K -model. Parameters as <strong>in</strong> Fig. 6.16.appear or disappear, but it does not expla<strong>in</strong> how stable periodic spik<strong>in</strong>g behaviorappears. Indeed, let us start with I = 42 <strong>in</strong> Fig. 6.23 and slowly <strong>in</strong>crease the parameter.The state of the I Na,p +I K -model is at the stable equilibrium. When I passes thebifurcation value, a large-amplitude stable limit cycle correspond<strong>in</strong>g to periodic spik<strong>in</strong>gappears, yet the model is still quiescent, because it is still at the stable equilibrium.Thus, the limit cycle is just a geometrical object <strong>in</strong> the phase space that corresponds tospik<strong>in</strong>g behavior. However, to actually exhibit spik<strong>in</strong>g, the state of the system must besomehow pushed <strong>in</strong>to the attraction doma<strong>in</strong> of the cycle, say by external stimulation.This issue is related to the computational properties of neurons, and it is discussed <strong>in</strong>detail <strong>in</strong> the next chapter.In Fig. 6.24 we depict the bifurcation diagram of the I Na,p +I K -model. For eachvalue of I, we simulate the model forward (t → ∞) to f<strong>in</strong>d the stable limit cycle andbackward (t → −∞) to f<strong>in</strong>d the unstable limit cycle. Then we plot their amplitudes(maximal voltage m<strong>in</strong>us m<strong>in</strong>imal voltage along the limit cycle) on the (I, V )-plane.One can clearly see that there is a fold limit cycle bifurcation (left) and a subcriticalAndronov-Hopf bifurcation (right). The left part of the bifurcation diagram looksexactly like the one for saddle-node bifurcation, which expla<strong>in</strong>s why the fold limitcycle bifurcation is often referred to as fold or saddle-node of periodics.The similarity of the fold limit cycle bifurcation and the saddle-node bifurcationis not a co<strong>in</strong>cidence. Stability of limit cycles can be studied us<strong>in</strong>g Floquet theory,Po<strong>in</strong>caré cross-section maps (Kuznetsov 1995), or by brute force, e.g., by reduc<strong>in</strong>gthe model to an appropriate polar coord<strong>in</strong>ate system. When a limit cycle attractorundergoes fold limit cycle bifurcation, its radius undergoes saddle-node bifurcation(this is a h<strong>in</strong>t to Ex. 10).


Bifurcations 193homocl<strong>in</strong>ic orbitstablelimit cyclesaddle saddle saddlea. supercritical saddle homocl<strong>in</strong>ic orbit bifurcationhomocl<strong>in</strong>ic orbitunstablelimit cyclesaddle saddle saddleb. subcritical saddle homocl<strong>in</strong>ic orbit bifurcationFigure 6.25: Saddle homocl<strong>in</strong>ic orbit bifurcation.


194 Bifurcationshomocl<strong>in</strong>ic orbithomocl<strong>in</strong>ic orbitsaddle homocl<strong>in</strong>ic orbitbifurcationsaddle-node on <strong>in</strong>variant circlebifurcationFigure 6.26: Two bifurcations <strong>in</strong>volv<strong>in</strong>g homocl<strong>in</strong>ic trajectories to an equilibrium.6.2.4 Homocl<strong>in</strong>icA limit cycle can appear or disappear via a saddle homocl<strong>in</strong>ic orbit bifurcation, asdepicted <strong>in</strong> Fig. 6.25. As the bifurcation parameter changes, the cycle becomes ahomocl<strong>in</strong>ic orbit to the saddle equilibrium, and its period becomes <strong>in</strong>f<strong>in</strong>ite. After thebifurcation, the cycle no longer exists. A necessary condition for such a bifurcation isthat the steady-state I-V relation is not monotonic.One should be careful to dist<strong>in</strong>guish the saddle homocl<strong>in</strong>ic orbit bifurcation fromthe saddle-node on <strong>in</strong>variant circle bifurcation depicted <strong>in</strong> Fig. 6.26. Indeed, it mightbe easy to confuse these bifurcations, s<strong>in</strong>ce both <strong>in</strong>volve an equilibrium and a largeamplitudehomocl<strong>in</strong>ic trajectory that becomes a limit cycle. The key difference is thatthe equilibrium is a saddle <strong>in</strong> the former and a saddle-node <strong>in</strong> the latter. The saddleequilibrium persists as the bifurcation parameter changes, whereas the saddle-nodeequilibrium disappears or bifurcates <strong>in</strong>to two po<strong>in</strong>ts, depend<strong>in</strong>g on the direction ofchange of the bifurcation parameter.Recall that a saddle on a plane has two real eigenvalues of opposite signs. Theirsum, λ 1 + λ 2 , is called the saddle quantity.• If λ 1 + λ 2 < 0, then the saddle homocl<strong>in</strong>ic orbit bifurcation is supercritical, whichcorresponds to the (dis)appearance of a stable limit cycle.• If λ 1 + λ 2 > 0, then the saddle homocl<strong>in</strong>ic orbit bifurcation is subcritical, whichcorresponds to the (dis)appearance of an unstable limit cycle.Thus, the saddle quantity plays the same role as the parameter a <strong>in</strong> the Andronov-Hopfbifurcation. The supercritical saddle homocl<strong>in</strong>ic orbit bifurcation is more common <strong>in</strong>neuronal models than the subcritical one due to the reason expla<strong>in</strong>ed <strong>in</strong> Sect. 6.3.6.Hence, we consider only the supercritical case below, and we drop the word “supercritical”for the sake of brevity.


Bifurcations 195stable manifoldout<strong>in</strong>?saddleunstable manifoldFigure 6.27: Saddle homocl<strong>in</strong>ic orbit bifurcationoccurs when the stable and unstable submanifolds ofthe saddle make a loop.A useful way to look at the bifurcation is to note that the saddle has one stableand one unstable direction on a phase plane. There are two orbits associated withthese directions, called the stable and unstable submanifolds, depicted <strong>in</strong> Fig. 6.27.Typically, the submanifolds miss each other, that is, the unstable submanifold goeseither <strong>in</strong>side or outside the stable one. This could happen for two different values ofthe bifurcation parameter. One can image that as the bifurcation parameter changescont<strong>in</strong>uously from one value to the other, the submanifolds jo<strong>in</strong> at some po<strong>in</strong>t and forma s<strong>in</strong>gle homocl<strong>in</strong>ic trajectory that starts and ends at the saddle.The saddle homocl<strong>in</strong>ic orbit bifurcation is ubiquitous <strong>in</strong> neuronal models, and it caneasily be observed <strong>in</strong> the I Na,p +I K -model with fast K + conductance, as we illustrate<strong>in</strong> Fig. 6.28. Let us start with I = 7 (top of Fig. 6.28) and decrease the bifurcationparameter I. First, there is only a stable limit cycle correspond<strong>in</strong>g to periodic spik<strong>in</strong>gactivity. When I decreases, a stable and an unstable equilibrium appear via saddlenodebifurcation (not shown <strong>in</strong> the figure), but the state of the model is still on thelimit cycle attractor. Further decrease of I moves the saddle equilibrium closer tothe limit cycle (case I = 4 <strong>in</strong> the figure), until the cycle becomes an <strong>in</strong>f<strong>in</strong>ite periodhomocl<strong>in</strong>ic orbit to the saddle (case I ≈ 3.08), and then disappears (case I = 1). Atthis moment, the state of the system approaches the stable equilibrium, and the tonicspik<strong>in</strong>g stops.Similarly to the fold limit cycle bifurcation, the saddle homocl<strong>in</strong>ic orbit bifurcationexpla<strong>in</strong>s how the limit cycle attractor correspond<strong>in</strong>g to periodic spik<strong>in</strong>g behavior appearsand disappears. However, it does not expla<strong>in</strong> the transition to periodic spik<strong>in</strong>gbehavior. Indeed, when I = 4 <strong>in</strong> Fig. 6.28, the limit cycle attractor exists, yet theneuron may still be quiescent because its state may be at the stable node. The periodicspik<strong>in</strong>g behavior appears only after external perturbations push the state of thesystem <strong>in</strong>to the attraction doma<strong>in</strong> of the limit cycle attractor, or I <strong>in</strong>creases furtherand the stable node disappears via a saddle-node bifurcation.We can use l<strong>in</strong>ear theory to estimate the frequency of the limit cycle attractornear saddle homocl<strong>in</strong>ic orbit bifurcation. Because the vector field is small near theequilibrium, the periodic trajectory slowly passes through a small neighborhood of theequilibrium, then quickly makes a rotation and returns to the neighborhood, as weillustrate <strong>in</strong> Fig. 6.29. Let T 1 denote the time required to make one rotation (dashed


196 Bifurcations0.80.60.4stable limit cyclen-nullcl<strong>in</strong>eI=7limit cyclesnVIV-nullcl<strong>in</strong>e0.20-80 -60 -40 -20 0membrane voltage, mVI=4nodesaddlestable limit cyclestable limit cyclenodesaddleI=3.08homocl<strong>in</strong>ic orbithomocl<strong>in</strong>ic orbitnodesaddleI=1m<strong>in</strong>/max of oscillations ofmembrane potential, mV0-20-40-60homocl<strong>in</strong>icbifurcationsaddlesnodeslimit cycles-800 2 4 6<strong>in</strong>jected dc-current, InodesaddleFigure 6.28: Saddle homocl<strong>in</strong>ic orbit bifurcation <strong>in</strong> the I Na,p +I K -model with parametersas <strong>in</strong> Fig. 4.1a and fast K + current (τ(V ) = 0.16). As the bifurcation parameter Idecreases, the stable limit cycle becomes a homocl<strong>in</strong>ic orbit to a saddle.


Bifurcations 1970.8-100.70.040.60.02-20K + activation gate, n0.50.40.30.2V-nullcl<strong>in</strong>e0saddleT 1-60 -55 -50n-nullcl<strong>in</strong>elimit cyclemembrane voltage, V (mV)-30-40T 2T 20.1T 10-80 -70 -60 -50 -40 -30 -20 -10 0 10membrane voltage, V (mV)-50T 1 T 2-600 1 2 3 4 5 6time, msFigure 6.29: The period of the limit cycle is T = T 1 + T 2 with T 2 → ∞ as the cycleapproaches the saddle equilibrium. Shown is the I Na,p +I K -model with I = 3.5.part of the limit cycle <strong>in</strong> the figure) and T 2 denote the time spent <strong>in</strong> the small neighborhoodof the saddle equilibrium (cont<strong>in</strong>uous part of the limit cycle <strong>in</strong> the shadowedregion), so that the period of limit cycle is T = T 1 + T 2 . While T 1 is relatively constant,T 2 → ∞ as I approaches the bifurcation value I b = 3.08, and the limit cycleapproaches the saddle. In Ex. 11 we show thatT 2 = − 1 λ 1ln{τ(I − I b )} ,where λ 1 is the positive (unstable) eigenvalue of the saddle, and τ is a parameter thatdepends on the size of the neighborhood, global features of the vector field, etc. Wecan represent the period, T , <strong>in</strong> the formT (I) = − 1 λ 1ln{τ 1 (I − I b )} ,where a s<strong>in</strong>gle parameter τ 1 = τe −λ 1T 1accounts for all global features of the model,<strong>in</strong>clud<strong>in</strong>g the width of the action potential, the shape of the limit cycle, etc. One caneasily determ<strong>in</strong>e τ 1 if the eigenvalue λ 1 and the period of the limit cycle is known for atleast one value of I. The I Na,p +I K -model has τ 1 = 0.2, as we show <strong>in</strong> Fig. 6.30. Noticethat the theoretical frequency 1000/T (I) matches the numerically found frequency <strong>in</strong>a broad range. Also, notice how imprecise the numerical results are (see <strong>in</strong>set <strong>in</strong> thefigure).


198 Bifurcationsfrequency (Hz)400300200200numericaltheoretical1001/ln10003.08 3.0903 3.1 3.2 3.3 3.4 3.5 3.6<strong>in</strong>jected dc-current, IFigure 6.30: Frequency of spik<strong>in</strong>g <strong>in</strong> the I Na,p +I K -model with parameters as <strong>in</strong>Fig. 6.28 near a saddle homocl<strong>in</strong>ic orbit bifurcation. Dots are numerical results; thecont<strong>in</strong>uous curve is ω(I) = 1000λ(I)/{− ln(0.2(I − 3.0814))}, where the eigenvalueλ(I) = 0.87 √ 4.51 − I was obta<strong>in</strong>ed from the normal form (6.3). The <strong>in</strong>set shows amagnified region near the bifurcation value I = 3.0814.big homocl<strong>in</strong>ic orbitlimit cycleFigure 6.31: Big saddle homocl<strong>in</strong>ic orbit bifurcation.Both, the saddle-node on <strong>in</strong>variant circle bifurcation and the saddle homocl<strong>in</strong>icorbit bifurcation result <strong>in</strong> spik<strong>in</strong>g with decreas<strong>in</strong>g frequency so that their frequencycurrent(F-I) curves go cont<strong>in</strong>uously to zero. The key difference is that the formerasymptotes as √ I − I b , whereas the latter as 1/ ln(I − I b ). The strik<strong>in</strong>g feature of thelogarithmic decay <strong>in</strong> Fig. 6.30 is that the frequency is greater than 100 Hz and thetheoretical curve does not seem to go to zero for all I except those <strong>in</strong> an <strong>in</strong>f<strong>in</strong>itesimalneighborhood of the bifurcation value I b . Such a neighborhood is almost impossible tocatch numerically, let alone experimentally <strong>in</strong> real neurons.Many neuronal models, and even some cortical pyramidal neurons (see Fig. 7.42)exhibit a saddle homocl<strong>in</strong>ic orbit bifurcation depicted <strong>in</strong> Fig. 6.31. Here, the unstablemanifold of a saddle returns to the saddle along the opposite side, thereby mak<strong>in</strong>g abig loop, hence the name big saddle homocl<strong>in</strong>ic orbit bifurcation. This k<strong>in</strong>d of bifurcationoften occurs when an excitable system is near a codimension-2 Bogdanov-Takensbifurcation considered <strong>in</strong> Sect. 6.3.3, and it has the same properties as the “small”homocl<strong>in</strong>ic orbit bifurcation considered above: it could be subcritical or supercritical,


Bifurcations 199heterocl<strong>in</strong>ic orbitFigure 6.32: Heterocl<strong>in</strong>ic orbit bifurcation does not change the existence of stability ofany equilibrium or periodic orbit.depend<strong>in</strong>g on the saddle quantity, it results <strong>in</strong> a logarithmic F-I curve, and it impliesthe co-existence of attractors. All methods of analysis of excitable systems near “small”saddle homocl<strong>in</strong>ic orbit bifurcations can also be applied to the case <strong>in</strong> Fig. 6.31.6.3 Other Interest<strong>in</strong>g CasesSaddle-node and Andronov-Hopf bifurcations of equilibria comb<strong>in</strong>ed with fold limitcycle, homocl<strong>in</strong>ic orbit bifurcation, and heterocl<strong>in</strong>ic orbit bifurcation (see Fig. 6.32)exhaust all possible bifurcations of codimension-1 on a plane. These bifurcations canalso occur <strong>in</strong> higher-dimensional systems. Below we discuss additional codimension-1bifurcations <strong>in</strong> three-dimensional phase space, and then we consider some codimension-2 bifurcations that play an important role <strong>in</strong> neuronal dynamics. The first-time readermay read only Sect. 6.3.6 and skip the rest.6.3.1 Three-dimensional phase spaceSo far we considered four bifurcations of equilibria and four bifurcations of limit cycleson a phase plane. The same eight bifurcations can appear <strong>in</strong> multi-dimensional systems.Below we briefly discuss the new k<strong>in</strong>ds of bifurcations that are possible <strong>in</strong> a threedimensionalphase space but cannot occur on a plane.First, there are no new bifurcations of equilibria <strong>in</strong> multi-dimensional phase space.Indeed, what could possibly happen with the Jacobian matrix of an equilibrium ofa multi-dimensional dynamical system? A simple zero eigenvalue would result <strong>in</strong>a saddle-node bifurcation, and a simple pair of purely imag<strong>in</strong>ary complex-conjugateeigenvalues would result <strong>in</strong> an Andronov-Hopf bifurcation. Both are exactly the sameas <strong>in</strong> the lower-dimensional systems considered before. Thus, add<strong>in</strong>g dimensions to adynamical system does not create new possibilities for bifurcations of equilibria.In contrast, add<strong>in</strong>g the third dimension to a planar dynamical system creates newpossibilities for bifurcations of limit cycles, some of which are depicted <strong>in</strong> Fig. 6.33.Below we briefly describe these bifurcations.The saddle-focus homocl<strong>in</strong>ic orbit bifurcation <strong>in</strong> Fig. 6.33 is similar to the saddlehomocl<strong>in</strong>ic orbit bifurcation considered <strong>in</strong> Sect. 6.2.4 except that the equilibrium hasa pair of complex-conjugate eigenvalues and a non-zero real eigenvalue. The homocl<strong>in</strong>icorbit orig<strong>in</strong>ates <strong>in</strong> the subspace spanned by the eigenvector correspond<strong>in</strong>g to


200 BifurcationsSaddle-Focus Homocl<strong>in</strong>ic Orbit BifurcationSubcritical Flip BifurcationSubcritical Neimark-Sacker BifurcationBlue-Sky CatastropheFold Limit Cycle on Homocl<strong>in</strong>ic Torus BifurcationFigure 6.33: Some codimension-1 bifurcations of limit cycles <strong>in</strong> three-dimensional phasespace (modified from Izhikevich 2000).


Bifurcations 201the real eigenvalue (as <strong>in</strong> Fig. 6.33) and term<strong>in</strong>ates along the subspace spanned by theeigenvectors correspond<strong>in</strong>g to the complex-conjugate pair. The reverse direction is alsopossible. Depend<strong>in</strong>g on the direction and the relative magnitude of the eigenvalues,this bifurcation can result <strong>in</strong> the (dis)appearance or a stable (supercritical) or unstable(subcritical) twisted large-period orbit.The subcritical flip bifurcation <strong>in</strong> Fig. 6.33 occurs when a stable periodic orbitis surrounded by an unstable orbit of twice the period. The unstable periodic orbitshr<strong>in</strong>ks to the stable one and makes it lose stability. This bifurcation is similar to thepitchfork bifurcation studied below, except it has codimension 1 (pitchfork bifurcationhas <strong>in</strong>f<strong>in</strong>ite codimension unless one considers dynamical systems with symmetry). Asupercritical flip bifurcation is similar, except that an unstable cycle is surrounded bya stable double-period cycle.The subcritical Neimark-Sacker bifurcation <strong>in</strong> Fig. 6.33 occurs when a stable periodicorbit is surrounded by an unstable <strong>in</strong>variant torus. The latter shr<strong>in</strong>ks and makesthe periodic orbit lose its stability. In some sense, which we will not elaborate on here,this bifurcation is similar to the supercritical Andronov-Hopf bifurcation of an equilibrium.The supercritical Neimark-Sacker bifurcation occurs when an unstable orbit issurrounded by a stable <strong>in</strong>variant torus.The blue-sky catastrophe <strong>in</strong> Fig. 6.33 occurs when a small amplitude stable limitcycle disappears and a large-amplitude large-period stable orbit appears out of nowhere(from the blue sky): The orbit has an <strong>in</strong>f<strong>in</strong>ite period at the bifurcation, yet it isnot homocl<strong>in</strong>ic to any equilibrium. A careful analysis shows that the large orbit ishomocl<strong>in</strong>ic to the small limit cycle at the moment the cycle disappears. In somesense, which we elaborate on later, this bifurcation is similar to the saddle-node on<strong>in</strong>variant circle bifurcation (see Ex. 20). In particular, both bifurcations share thesame asymptotics.The fold limit cycle on homocl<strong>in</strong>ic torus bifurcation <strong>in</strong> Fig. 6.33 is similar to theblue-sky catastrophe except that the disappearance of the small periodic orbit results<strong>in</strong> a large-amplitude torus (quasi-periodic) attractor.6.3.2 Cusp and pitchforkRecall that an equilibrium x b of a one-dimensional system ẋ = f(x, b) is at a saddlenodebifurcation when f x = 0 (first derivative of f) but f xx ≠ 0 (second derivativeof f) at the equilibrium. The latter is called the non-degeneracy condition, and itguarantees that the system dynamics is equivalent to that of ẋ = c(b) + x 2 .If f x = 0 and f xx = 0, but f xxx ≠ 0, then the equilibrium is at the codimension-2cusp bifurcation, and the behavior of the system near the equilibrium can be describedby the topological normal formẋ = c 1 (b) + c 2 (b)x + ax 3 ,wherec 1 (b) = f(x b , b) , c 2 (b) = f x (x b , b) , a = f xxx /6 ≠ 0 ,


202 Bifurcationsc 1c 2xFigure 6.34: Cusp surface.<strong>in</strong> particular, c 1 = c 2 = 0 at the cusp po<strong>in</strong>t. The cusp bifurcation is supercritical whena < 0 and subcritical otherwise. It is expla<strong>in</strong>ed by the shape of the surfacec 1 + c 2 x + ax 3 = 0depicted <strong>in</strong> Fig. 6.34.Let us treat c 1 and c 2 as two <strong>in</strong>dependent parameters, and check that there aresaddle-node bifurcations <strong>in</strong> any neighborhood of the cusp po<strong>in</strong>t. The bifurcation setsof the topological normal form can easily be found. Differentiat<strong>in</strong>g c 1 + c 2 x + ax 3 withrespect to x gives c 2 +3ax 2 . Equat<strong>in</strong>g both of these expressions to zero and elim<strong>in</strong>at<strong>in</strong>gx gives the saddle-node bifurcation curvesc 1 = ± √ 2 ( c2) 3/2,a 3depicted at the bottom of Fig. 6.34.S<strong>in</strong>ce c 1 = c 1 (b) and c 2 = c 2 (b), vary<strong>in</strong>g the bifurcation parameter b results <strong>in</strong> apath on the (c 1 , c 2 )-plane. Depend<strong>in</strong>g on the shape and location of this path, one canget many 1-dimensional bifurcation diagrams. A summary of some special cases isdepicted <strong>in</strong> Fig. 6.35 show<strong>in</strong>g that there can be many <strong>in</strong>terest<strong>in</strong>g dynamical regimes <strong>in</strong>the vic<strong>in</strong>ity of a cusp bifurcation po<strong>in</strong>t.An important special case is when c 1 = 0 and c 2 (b) = b, so that the topologicalnormal form isẋ = bx + ax 3 .This form corresponds to a pitchfork bifurcation, whose diagram is depicted <strong>in</strong> Fig. 6.36(see also the bottom bifurcation diagram <strong>in</strong> Fig. 6.35). This bifurcation has an <strong>in</strong>f<strong>in</strong>itecodimension unless one considers dynamical systems with symmetry, e.g., ẋ = f(x, b)with f(−x, b) = −f(x, b) for all x and b.6.3.3 Bogdanov-TakensCan an equilibrium undergo Andronov-Hopf and saddle-node bifurcations simultaneously?There are two possibilities, illustrated <strong>in</strong> Fig. 6.37:


Bifurcations 203xc 2 (b)xbbxxbxxbbbc 1 (b)xxxxb b bbFigure 6.35: Summary of special cases for the supercritical cusp bifurcation. Dottedsegments are paths c 1 = c 1 (b), c 2 = c 2 (b), where b is a one-dimensional bifurcationparameter. The correspond<strong>in</strong>g bifurcation diagrams are depicted <strong>in</strong> boxes. Cont<strong>in</strong>uouscurves represent stable solutions, dashed curves represent unstable solutions (modifiedfrom Hoppensteadt and Izhikevich 1997).+b/aunstablexxstable+b/|a|stableunstablebstableunstableb-b/aunstablestable-b/|a|¡ ¡ ¡¡ ¡ ¡¡ ¡ ¡¡ ¡ ¡¡ ¡¡ ¡¡ ¡ ¡¡ ¡ ¡¡ ¡ ¡¡ ¡ ¡subcritical pitchfork, a > 0supercritical pitchfork, a < 0Figure 6.36: Pitchfork bifurcation diagrams.


204 BifurcationseigenvaluesHopffoldeigenvaluesfold-Hopf bifurcationBogdanov-Takens bifurcationFigure 6.37: Two ways an equilibrium can undergo a saddle-node (fold) and anAndronov-Hopf bifurcations simultaneously.• (Fold-Hopf) The Jacobian matrix at the equilibrium has a pair of pure imag<strong>in</strong>arycomplex-conjugate eigenvalues (Andronov-Hopf bifurcation) and one zeroeigenvalue (saddle-node bifurcation). In this case the two bifurcations occur <strong>in</strong>different subspaces.• (Bogdanov-Takens) The Jacobian matrix has two zero eigenvalues.case the two bifurcations occur <strong>in</strong> the same subspace.In thisThe fold-Hopf bifurcations occurs <strong>in</strong> systems hav<strong>in</strong>g dimension 3 and up, while theBogdanov-Takens bifurcation can occur <strong>in</strong> two-dimensional systems. Both bifurcationshave codimension-2; that is, they require 2 bifurcation parameters. Notice thatfold-Hopf bifurcation has 3 eigenvalues with zero real part, whereas Bogdanov-Takensbifurcation has only 2 zero eigenvalues. This bifurcation can on the one hand be viewedas a saddle-node bifurcation <strong>in</strong> which another (negative) eigenvalue gets arbitrary closeto zero, and on the other hand as an Andronov-Hopf bifurcation <strong>in</strong> which imag<strong>in</strong>arypart of the complex-conjugate eigenvalues goes to zero.The Jacobian matrix of an equilibrium at the Bogdanov-Takens bifurcation satisfiestwo conditions: det L = 0 (saddle-node bifurcation) and tr L = 0 (Andronov-Hopfbifurcation). For example, it can have the formL =( 0 10 0). (6.10)Because of these two conditions, the codimension of this bifurcation is 2. There arealso certa<strong>in</strong> non-degeneracy and transversality conditions (see Kuznetsov 1995). Thecorrespond<strong>in</strong>g topological normal form,˙u = v ,˙v = a + bu + u 2 + σuv ,(6.11)


Bifurcations 2051 2b1 2SN23 AH 41BT4a3 AH 4SHOAH2 3SN1SN 1 SN 2SN 1SN 2SHOsupercritical,BTSHOBTs0Figure 6.38: Bogdanov-Takens (BT) bifurcation diagram of the topological normal form(6.11). Abbreviations: AH - Andronov-Hopf bifurcation, SN - saddle-node bifurcation,SHO - saddle homocl<strong>in</strong>ic orbit bifurcation.has two bifurcation parameters, a and b, and the parameter σ = ±1 determ<strong>in</strong>es whetherthe bifurcation is subcritical or supercritical. This parameter depends on the comb<strong>in</strong>ationof the second-order partial derivatives with respect to the first variable, and it isnon-zero because of the non-degeneracy conditions (Kuznetsov 1995). Bifurcation diagramand representative phase portraits for various a, b and σ are depicted <strong>in</strong> Fig. 6.38(the case σ > 0 can be reduced to σ < 0 by the substitution t → −t and v → −v).A remarkable fact is that the saddle-node and the Andronov-Hopf bifurcations do notoccur alone. There is also a saddle homocl<strong>in</strong>ic orbit bifurcation appear<strong>in</strong>g near theBogdanov-Takens po<strong>in</strong>t.Bogdanov-Takens bifurcation occurs often <strong>in</strong> neuronal models with nullcl<strong>in</strong>es <strong>in</strong>tersect<strong>in</strong>gas <strong>in</strong> Fig. 6.39a. We show <strong>in</strong> the next chapter that this bifurcation separates<strong>in</strong>tegrators from resonators, and it could occur <strong>in</strong> some layer 5 pyramidal neurons ofrat visual cortex, as we discuss <strong>in</strong> Sect. 7.2.11 and Sect. 8.2.1. The two equilibria <strong>in</strong>the lower (left) knee of the fast nullcl<strong>in</strong>e <strong>in</strong> Fig. 6.39b are not necessarily a saddle anda stable node, but could be a saddle and an (un)stable focus, as <strong>in</strong> the phase portraits<strong>in</strong> Fig. 6.38.


206 Bifurcationsfast nullcl<strong>in</strong>eslow nullcl<strong>in</strong>eBogdanov-Takenspo<strong>in</strong>t(a)(b)Figure 6.39: Intersection of nullcl<strong>in</strong>es of a two-dimensional system result<strong>in</strong>g <strong>in</strong>Bogdanov-Takens bifurcation.I 0 =0 I 1 I 2bighomocl<strong>in</strong>ic orbitI 3 I 4 I 5smallhomocl<strong>in</strong>ic orbitsubcriticalAndronov-HopfbifurcationFigure 6.40: Transformations of phase portraits of a neuronal model near subcriticalBogdanov-Takens bifurcation po<strong>in</strong>t as the magnitude of the <strong>in</strong>jected current I <strong>in</strong>creases(here I k+1 > I k ). Shaded regions are the attraction doma<strong>in</strong>s of the equilibrium correspond<strong>in</strong>gto the rest<strong>in</strong>g state.


Bifurcations 207(a) (b) (c) (d)g=0f=0(e) (f) (g) (h)Figure 6.41: Canard (French duck) limit cycles <strong>in</strong> a relaxation oscillator (hand draw<strong>in</strong>g).Interest<strong>in</strong>gly, the global vector field structure of neuronal models with nullcl<strong>in</strong>esas <strong>in</strong> Fig. 6.39a results <strong>in</strong> the birth of a spik<strong>in</strong>g limit cycle attractor via a big saddlehomocl<strong>in</strong>ic orbit bifurcation, so the neuronal model undergoes a cascade of bifurcationsdepicted <strong>in</strong> Fig. 6.40 as the amplitude of the <strong>in</strong>jected current I <strong>in</strong>creases. The localphase portraits correspond<strong>in</strong>g to I 0 , I 1 , and I 2 are topologically equivalent to the phaseportrait “1” <strong>in</strong> Fig. 6.38, right. (The equivalence is only local near the left knee; thereis no global equivalence because of the extra equilibrium <strong>in</strong> Fig. 6.40 and because ofthe big homocl<strong>in</strong>ic or periodic orbit.) As I <strong>in</strong>creases, a stable large-amplitude spik<strong>in</strong>glimit cycle appears via a big supercritical homocl<strong>in</strong>ic orbit bifurcation at some I 1 . Itcoexists with the stable rest<strong>in</strong>g state for all I 1 < I < I 5 . At some po<strong>in</strong>t I 2 , thesaddle quantity, i.e., the sum of its eigenvalues, changes from negative to positive(it is zero at the Bogdanov-Takens bifurcation), so another saddle homocl<strong>in</strong>ic orbitbifurcation (at some I 3 ) occurs, which is subcritical, giv<strong>in</strong>g birth to an unstable limitcycle. The phase portrait at I 3 is locally topologically equivalent to the one markedSHO <strong>in</strong> Fig. 6.38. Similarly, the phase portrait at I 4 is locally equivalent to the onemarked “2” <strong>in</strong> Fig. 6.38. The unstable cycle shr<strong>in</strong>ks to the equilibrium and makes it losestability via a subcritical Andronov-Hopf bifurcation at some I 5 , which corresponds tocase AH <strong>in</strong> Fig. 6.38. Further <strong>in</strong>crease of I converts the unstable focus <strong>in</strong>to an unstablenode, which approaches the saddle and disappears via the saddle-node bifurcation SN 1<strong>in</strong> Fig. 6.38 (not shown <strong>in</strong> Fig. 6.40).6.3.4 Relaxation oscillators and CanardsLet us consider a relaxation oscillatorẋ = f(x, y, b) (fast variable)ẏ = µg(x, y, b) (slow variable)


208 Bifurcationswith fast and slow nullcl<strong>in</strong>es as <strong>in</strong> Fig. 6.41a and µ ≪ 1. Suppose that there is a stableequilibrium, as <strong>in</strong> Fig. 6.41a, for some values of the bifurcation parameter b < 0, anda stable limit cycle, as <strong>in</strong> Fig. 6.41h, for some other values b > 0. What k<strong>in</strong>d of abifurcation of the equilibrium occurs when b <strong>in</strong>creases from negative to positive, andthe slow nullcl<strong>in</strong>e passes the left knee of the fast N-shaped nullcl<strong>in</strong>e?The Jacobian matrix at the equilibrium has the form( )fx fL =yµg x µg yS<strong>in</strong>ce f x = 0 at the knee (prove this), but f y is typically not, the Jacobian matrixresembles the one for the Bogdanov-Takens bifurcation (6.10) <strong>in</strong> the limit µ = 0.However, the resemblance is only superficial, s<strong>in</strong>ce the relaxation oscillator does notsatisfy the non-degeneracy conditions. In particular, second-order partial derivativesof µg(x, y, b) vanish <strong>in</strong> the limit µ → 0, result<strong>in</strong>g <strong>in</strong> σ = 0 and <strong>in</strong> the disappearance ofthe term u 2 <strong>in</strong> the topological normal form (6.11).A purely geometrical consideration confirms that the transition from Fig. 6.41a toFig. 6.41h cannot be of the Bogdanov-Takens type, s<strong>in</strong>ce there is a unique equilibriumand no possibility for a saddle-node bifurcation, which always accompanies theBogdanov-Takens bifurcation. Actually, the equilibrium loses stability via Andronov-Hopf bifurcation that occurs whentr L = f x + µg y = 0 and det L = µ(f x g y − f y g x ) > 0 .The loss of stability typically happens not at the left knee, where f x = 0, but a littlebit to the right of the knee, where f x = −µg y > 0 (because g y < 0 <strong>in</strong> neuronal models).We already saw this phenomenon <strong>in</strong> Sect. 4.2.6 when we considered FitzHugh-Nagumomodel.An <strong>in</strong>terest<strong>in</strong>g observation is that the period of damped or susta<strong>in</strong>ed oscillationsnear the Andronov-Hopf bifurcation po<strong>in</strong>t <strong>in</strong> Fig. 6.41b is of the order 1/ √ µ, becausethe frequency ω = √ det L ≈ √ µ, whereas the period of large-amplitude relaxationoscillation is of the order 1/µ, because it takes 1/µ units of time to slide up anddown along the branches of the fast nullcl<strong>in</strong>e <strong>in</strong> Fig. 6.41h. Thus, the period of smallsubthreshold oscillations of a neural model may have no relation to the period ofspik<strong>in</strong>g, if the model is a relaxation oscillator.The Andronov-Hopf bifurcation could be supercritical or subcritical, depend<strong>in</strong>g onthe functions f and g; see Ex. 14 and Ex. 18. Figure 6.41 depicts the supercriticalcase. In the subcritical case, stable and unstable limit cycles are typically born viafold limit cycle bifurcation, then the unstable limit cycle goes through the shapes as<strong>in</strong> Fig. 6.41g,f,e,d,c, and b, and then shr<strong>in</strong>ks to a po<strong>in</strong>t.CanardsThe dist<strong>in</strong>ctive feature of limit cycles <strong>in</strong> Fig. 6.41c-g is that they follow the unstablebranch (dashed curve) of the fast nullcl<strong>in</strong>e before jump<strong>in</strong>g to the left or to the right


Bifurcations 209(stable) branches. Due to the relaxation nature of the system, the vector field ishorizontal outside the N-shape fast nullcl<strong>in</strong>e, so any transition from Fig. 6.41b toFig. 6.41h must gradually go through the stages <strong>in</strong> Fig. 6.41c-g. Because the cycle <strong>in</strong>Fig. 6.41f resembles a French duck, at least <strong>in</strong> the eyes of the French mathematiciansE. Benoit, J.-L. Callot, F. Diener, and M. Diener, who discovered this phenomenon <strong>in</strong>1977, it is often called a canard cycle.In general, any trajectory that follows the unstable branch is called a canard trajectory.Canard trajectories play an important role <strong>in</strong> def<strong>in</strong><strong>in</strong>g thresholds for resonatorneurons, as we discuss <strong>in</strong> Sect. 7.2.5. It takes of the order of 1/µ units of time toslide along the unstable branch of the fast nullcl<strong>in</strong>e. A small perturbation to the leftor to the right could result <strong>in</strong> an immediate jump to the correspond<strong>in</strong>g stable branchof the nullcl<strong>in</strong>e. Hence, the <strong>in</strong>itial conditions should be specified with an unrealisticprecision of the order of e −1/µ to follow the branch, which expla<strong>in</strong>s why the canardtrajectories are difficult to catch numerically, let alone experimentally. Consequently,the canard cycles, though stable, also exits <strong>in</strong> an exponentially small region of valuesof the parameter b. A typical simulation shows an explosion of a stable limit cyclefrom small (Fig. 6.41b) to large (Fig. 6.41h) as the parameter b is slowly varied. Insummary, canard cycles <strong>in</strong> two-dimensional relaxation oscillators play an importantrole of thresholds, but they are fragile and rather exceptional.In contrast, canard trajectories <strong>in</strong> three-dimensional relaxation oscillators (one fastand two slow variables) are generic <strong>in</strong> the sense that they exits <strong>in</strong> a wide range ofparameter values. A simple way to see this is to treat b as the second slow variable.Then, there is a set of <strong>in</strong>itial conditions correspond<strong>in</strong>g to the canard trajectories.Study<strong>in</strong>g canards <strong>in</strong> R 3 goes beyond the scope of this book (see bibliography at theend of this chapter).6.3.5 Baut<strong>in</strong>What happens when a subcritical Andronov-Hopf bifurcation becomes supercritical,that is, when the parameter a <strong>in</strong> the topological normal form for Andronov-Hopf bifurcation(6.8,6.9) changes sign? The bifurcation becomes degenerate when a = 0,and the behavior of the system is described by the topological normal form for Baut<strong>in</strong>bifurcation, which we write here <strong>in</strong> the complex formż = (c + iω)z + az|z| 2 + a 2 z|z| 4 , (6.12)where z ∈ C is a complex variable, and c, a and a 2 are real parameters. The parametersa and a 2 are called the first and second Liapunov coefficients (often spelled Lyapunovcoefficients). The Baut<strong>in</strong> bifurcation occurs when a = c = 0 and a 2 ≠ 0, and henceit has codimension 2. It is subcritical when a 2 > 0 and supercritical otherwise. Ifa 2 = 0, then one needs to consider the next term a 3 z|z| 6 <strong>in</strong> the normal form to get abifurcation of codimension-3, etc.We can easily determ<strong>in</strong>e bifurcations of the topological normal form. First of all,(6.12) undergoes Andronov-Hopf bifurcation when c = 0, which is supercritical for


210 BifurcationsSupercritical Andronov-HopfBifurcationc = 0a < 0Baut<strong>in</strong>Bifurcation0c2a - 4c a = 02c = 0a > 0FoldLimit CycleBifurcationSubcritical Andronov-HopfBifurcationaFigure 6.42: Supercritical Baut<strong>in</strong> bifurcation <strong>in</strong> (6.12); see also Fig. 9.42, left.a < 0 and subcritical otherwise. Moreover, if a and a 2 have different signs, then (6.12)undergoes fold limit cycle bifurcation whena 2 − 4ca 2 = 0 ,as we illustrate <strong>in</strong> Fig. 6.42. Thus, both Andronov-Hopf and fold limit cycle bifurcationsoccur simultaneously at the Baut<strong>in</strong> po<strong>in</strong>t a = c = 0. Many two-dimensional neuronalmodels, such as the I Na,p +I K -model with low-threshold K + current, are relatively nearthis bifurcation, which expla<strong>in</strong>s why the unstable limit cycle <strong>in</strong>volved <strong>in</strong> the subcriticalAndronov-Hopf bifurcation is usually born via fold limit cycle bifurcation. There issome evidence that rodent trigem<strong>in</strong>al <strong>in</strong>terneurons, dorsal root ganglion neurons, andmes V neuron <strong>in</strong> bra<strong>in</strong>stem are also near this bifurcation; see Sect. 9.3.3.6.3.6 Saddle-node homocl<strong>in</strong>ic orbitLet us compare the saddle-node on <strong>in</strong>variant circle bifurcation and the saddle homocl<strong>in</strong>icorbit bifurcation depicted <strong>in</strong> Fig. 6.43, top. In both cases there is a homocl<strong>in</strong>icorbit, i.e., a trajectory that orig<strong>in</strong>ates and term<strong>in</strong>ates at the same equilibrium. However,the equilibria are of different types, and the orbit returns to them along differentdirections. Now suppose a system undergoes both bifurcations simultaneously, as weillustrate <strong>in</strong> Fig. 6.43, bottom. Such a bifurcation, called saddle-node homocl<strong>in</strong>ic orbit


Bifurcations 211homocl<strong>in</strong>ic orbithomocl<strong>in</strong>ic orbithomocl<strong>in</strong>ic orbitsaddle-node on <strong>in</strong>variant circlebifurcationsaddle homocl<strong>in</strong>ic orbitbifurcationsaddle-node homocl<strong>in</strong>ic orbitbifurcationFigure 6.43: Saddle-node homocl<strong>in</strong>ic orbit bifurcation occurs when a systems undergoesa saddle-node on <strong>in</strong>variant circle and saddle homocl<strong>in</strong>ic orbit bifurcations simultaneously.bifurcation, has codimension 2, s<strong>in</strong>ce two strict conditions must be satisfied: First,the equilibrium must be at the saddle-node bifurcation po<strong>in</strong>t, i.e., must have eigenvalueλ 1 = 0. Second, the homocl<strong>in</strong>ic trajectory must return to the equilibrium alongthe non-central direction, i.e., along the stable direction correspond<strong>in</strong>g to the negativeeigenvalue λ 2 . S<strong>in</strong>ce the saddle-node quantity, λ 1 + λ 2 , is always negative, thisbifurcation always results <strong>in</strong> the (dis)appearance of a stable limit cycle.In Fig. 6.44 we illustrate the saddle-node homocl<strong>in</strong>ic orbit bifurcation us<strong>in</strong>g theI Na,p +I K -model with two bifurcation parameters: the <strong>in</strong>jected dc-current I and the K +time constant τ. The bifurcation occurs at the po<strong>in</strong>t (I, τ) = (4.51, 0.17). Notice thatthere are three other codimension-1 bifurcation curves converg<strong>in</strong>g to this codimension-2po<strong>in</strong>t, as we illustrate <strong>in</strong> Fig. 6.45. S<strong>in</strong>ce the model undergoes a saddle-node bifurcationat I = 4.51 and any τ, the straight vertical l<strong>in</strong>e I = 4.51 is the saddle-node bifurcationcurve. The po<strong>in</strong>t τ = 0.17 on this l<strong>in</strong>e separates two cases: When τ > 0.17, theactivation and deactivation of K + current is sufficiently slow so that the membranepotential V undershoots the equilibrium, result<strong>in</strong>g <strong>in</strong> the saddle-node on <strong>in</strong>variantcircle bifurcation. When τ < 0.17, deactivation of K + current is fast, and V overshootsthe saddle-node equilibrium, result<strong>in</strong>g <strong>in</strong> the saddle-node off limit cycle bifurcation.Shaded triangular areas <strong>in</strong> the figures denote the parameter region correspond<strong>in</strong>gto the bistability of stable equilibrium and a limit cycle attractor (rest<strong>in</strong>g and spik<strong>in</strong>gstates). Let us decrease the parameter I and cross such a region from right to left.When I = 4.51, a saddle and a node equilibria appear. Decreas<strong>in</strong>g I further moves thesaddle equilibrium rightward and the limit cycle leftward, until they merge. This occurs


212 BifurcationsK + conductance time constant0.180.1750.170.1650.16rest (excitable)bistabilitysaddle homocl<strong>in</strong>ic orbit bifurcationsaddle-node on <strong>in</strong>variantcircle bifurcationsaddle-node bifurcationperiodic spik<strong>in</strong>gsaddle-nodehomocl<strong>in</strong>ic orbit bifurcation0.1550 1 2 3 4 4.51 5 6 7 8 9<strong>in</strong>jected dc-current IFigure 6.44: Unfold<strong>in</strong>g of saddle-node homocl<strong>in</strong>ic orbit bifurcation <strong>in</strong> the I Na,p +I K -model with parameters as <strong>in</strong> Fig. 6.28.on the saddle homocl<strong>in</strong>ic orbit bifurcation curve, which is determ<strong>in</strong>ed numerically <strong>in</strong>Fig. 6.44.Neuronal models exhibit<strong>in</strong>g saddle-node homocl<strong>in</strong>ic bifurcations can be reduced toa topological normal form˙V = c(b − b sn ) + a(V − V sn ) 2 , if V (t) = V max , then V (t) ← V reset (6.13)which is similar to that for saddle-node bifurcation (6.2) except that there is a resetV ← V reset when the membrane voltage reaches certa<strong>in</strong> large value V max . Any sufficientlylarge V max would work equally well, even V max = +∞, because V reaches +∞<strong>in</strong> a f<strong>in</strong>ite time; see Ex. 3. Us<strong>in</strong>g V max = 30 and results of Sect. 6.1.1, we f<strong>in</strong>d that thetopological normal form for the I Na,p +I K -model is˙V = (I − 4.51) + 0.1887(V + 61) 2 , if V (t) = 30, then V (t) ← V reset .The saddle-node homocl<strong>in</strong>ic bifurcation occurs when I = 4.51 and V reset = −61. Thisnormal form is called quadratic <strong>in</strong>tegrate-and-fire neuron; see Chapters 3 and 8.


Bifurcations 213homocl<strong>in</strong>ic orbit bifurcationsaddle-node bifurcation saddle-node on<strong>in</strong>variant circle bifurcationsaddle-nodehomocl<strong>in</strong>ic orbitbifurcationFigure 6.45: Unfold<strong>in</strong>g of saddle-node homocl<strong>in</strong>ic orbit bifurcation.The topological normal form (6.13) is a useful equation, as we will see <strong>in</strong> the restof the book. It describes quantitative and qualitative features of neuronal dynamicsremarkably well, yet it has only one non-l<strong>in</strong>ear term. This makes it suitable for realtimesimulations of huge numbers of neurons. Its bifurcation structure is studied <strong>in</strong>Ex. 12 (see also Fig. 8.3), and the reader should at least look at the solution at theend of the book.6.3.7 Hard and soft loss of stabilityBifurcation is a qualitative change of the phase portrait of a system. Not all changeshowever are equally dramatic. Some are even hardly noticeable. For example, consideran equilibrium undergo<strong>in</strong>g supercritical Andronov-Hopf bifurcation: As a bifurcationparameter changes, the equilibrium loses stability and a small-amplitude stable limitcycle appears, as <strong>in</strong> Fig. 6.11. The state of the system rema<strong>in</strong>s near the equilibrium; itjust exhibits small-amplitude oscillations around it. We can change the parameter <strong>in</strong>the opposite direction, an then the limit cycle shr<strong>in</strong>ks to a po<strong>in</strong>t and the system returnsto the equilibrium. In neurons, such a bifurcation does not lead to an immediate spike.The neuron rema<strong>in</strong>s quiescent; it just exhibits subthreshold small-amplitude susta<strong>in</strong>edoscillations. Such a loss of stability is called soft: the equilibrium is no longer stable,but its small neighborhood rema<strong>in</strong>s attractive. Supercritical pitchfork, cusp and flipbifurcations correspond to soft loss of stability.


214 BifurcationsIn contrast, if the equilibrium loses stability via subcritical Andronov-Hopf bifurcation,the state of the system diverges from it, which results <strong>in</strong> an immediate spike orsome k<strong>in</strong>d of a large-amplitude jump. Such a loss of stability is called hard: neither theequilibrium nor its neighborhood are attractive. The hard loss of stability usually leadsto noticeable or catastrophic changes <strong>in</strong> systems behavior, and the stability boundaryis called dangerous (Baut<strong>in</strong> 1949). Chang<strong>in</strong>g the bifurcation parameter <strong>in</strong> the oppositedirection will make the equilibrium stable aga<strong>in</strong>, but may not br<strong>in</strong>g the state of thesystem back to it. Saddle-node bifurcation is hard unless it is on an <strong>in</strong>variant circle.In this case, the loss of stability is catastrophic, i.e., lead<strong>in</strong>g to noticeable spikes, butreversible. Saddle homocl<strong>in</strong>ic orbit bifurcation is hard, regardless of whether it is subcriticalor supercritical. In general, most bifurcations <strong>in</strong> neurons or at least <strong>in</strong> neuronalmodels are hard.Review of Important Concepts• Stable equilibrium (rest<strong>in</strong>g state) <strong>in</strong> a typical neuronal model caneither– disappear via saddle-node bifurcation, which can be off or on<strong>in</strong>variant circle, or– lose stability via Andronov-Hopf bifurcation, which can be supercriticalor subcritical.These four cases are summarized <strong>in</strong> Fig. 6.46.• Stable limit cycle (periodic spik<strong>in</strong>g state) <strong>in</strong> a typical twodimensionalneuronal model can either– be cut by saddle-node on <strong>in</strong>variant circle bifurcation,– shr<strong>in</strong>k to a po<strong>in</strong>t via supercritical Andronov-Hopf bifurcation,– disappear via fold limit cycle bifurcation, or– disappear via saddle homocl<strong>in</strong>ic orbit bifurcation.These four cases are summarized <strong>in</strong> Fig. 6.47.• Some atypical (codimension-2) bifurcations may play importantroles <strong>in</strong> neuronal dynamics.• Bogdanov-Takens bifurcation separates <strong>in</strong>tegrators from resonators.Bibliographical NotesThough bifurcation theory can be traced back to Po<strong>in</strong>care and Andronov, it is a relativelynew branch of mathematics. The first attempt to apply it to neuroscience was as


Bifurcations 215nodesaddlesaddle-nodesaddle-node bifurcation<strong>in</strong>variant circlenode saddle saddle-nodesaddle-node on <strong>in</strong>variant circle (SNIC) bifurcationsupercritical Andronov-Hopf bifurcationsubcritical Andronov-Hopf bifurcationFigure 6.46: Summary of all codimension-1 bifurcations of a stable equilibrium (rest<strong>in</strong>gstate).


216 Bifurcations<strong>in</strong>variant circlesaddle-nodenodesaddlesaddle-node on <strong>in</strong>variant circle (SNIC) bifurcationsupercritical Andronov-Hopf bifurcationstableunstablelimit cyclelimitcyclefoldlimitcyclefold limit cycle bifurcationhomocl<strong>in</strong>ic orbitstablelimit cyclesaddle saddle saddlesaddle homocl<strong>in</strong>ic orbit bifurcationFigure 6.47: Summary of all codimension-1 bifurcations of a stable limit cycle (tonicspik<strong>in</strong>g state) on a plane.


Bifurcations 217Figure 6.48: Richard FitzHugh with analog computer, National Institute of Health,Bethesda, Maryland, ca. 1960 (photograph provided by R. FitzHugh <strong>in</strong> 2005).early as 1955, when Richard FitzHugh concluded his paper on mathematical model<strong>in</strong>gof threshold phenomena say<strong>in</strong>g that many neuronal properties“... are <strong>in</strong>variant under cont<strong>in</strong>uous, one-to-one transformations of the coord<strong>in</strong>ates ofphase space and fall with<strong>in</strong> the doma<strong>in</strong> of topology, a branch of mathematics whichmay be <strong>in</strong>tr<strong>in</strong>sically better fitted for the prelim<strong>in</strong>ary description and classification ofbiological systems than analysis, which <strong>in</strong>cludes differential equations. This suggestionis of little practical value at present, s<strong>in</strong>ce too little is known of the topology ofvector fields <strong>in</strong> many-dimensional spaces, at least to those <strong>in</strong>terested <strong>in</strong> theoreticalbiology. Nevertheless, the most logical procedure <strong>in</strong> the description of a complexbiological system might be to characterize the topology of its phase space, then toestablish a set of physically identifiable coord<strong>in</strong>ates <strong>in</strong> the space, and f<strong>in</strong>ally to fitdifferential equations to the trajectories, <strong>in</strong>stead of try<strong>in</strong>g to reach this f<strong>in</strong>al goal atone leap.”It is remarkable that FitzHugh was explicitly talk<strong>in</strong>g about topological equivalence andbifurcations, though never called them such, years before these mathematical notionswere firmly established. This book cont<strong>in</strong>ues the l<strong>in</strong>e of research <strong>in</strong>itiated by FitzHughand further developed by R<strong>in</strong>zel and Ermentrout (1989).In this chapter we provided a fairly detailed exposition of bifurcation theory. Whatwe covered should be sufficient not only for understand<strong>in</strong>g the rest of the book, butalso for navigat<strong>in</strong>g through bifurcation papers concerned with computational neuro-


218 Bifurcationsbifurcationssaddle-nodesaddle-node on <strong>in</strong>variant cicleAndronov-Hopfsaddle homocl<strong>in</strong>ic orbitfold limit cyclesaddle-node homocl<strong>in</strong>ic orbitBogdanov-TakensBaut<strong>in</strong>flipalternative namesfold, limit po<strong>in</strong>t, saddle-node off limit cycleSNIC, saddle-node on limit cycle (SNLC), circle,saddle-node homocl<strong>in</strong>ic, saddle-node central homocl<strong>in</strong>ic,saddle-node <strong>in</strong>f<strong>in</strong>ite period (SNIPer), homocl<strong>in</strong>icHopf, Po<strong>in</strong>care-Andronov-Hopfhomocl<strong>in</strong>ic, saddle-loop, saddle separatrix loop, Andronov-Leontovichsaddle-node of limit cycles, double limit cycle, fold cycle,saddle-node (fold) of periodicssaddle-node noncentral homocl<strong>in</strong>ic, saddle-node separatrix-loopTakens-Bogdanov, double-zerodegenerate Hopf, generalized Hopfperiod doubl<strong>in</strong>gFigure 6.49: Popular alternative names to some of the bifurcations considered <strong>in</strong> thischapter.science. More bifurcation theory, <strong>in</strong>clud<strong>in</strong>g bifurcations <strong>in</strong> mapp<strong>in</strong>gs x n+1 = f(x n , b),can be found <strong>in</strong> the excellent book “Elements of Applied Bifurcation Theory” by YuriKuznetsov (1995, new edition 2004), which, however, might be a bit technical fora non-mathematician. Some of the bifurcations considered <strong>in</strong> this chapter, such asthe blue-sky catastrophe, are classified as “exotic” by Kuznetsov (1995), though thecatastrophe was recently found <strong>in</strong> a model of a leech heart <strong>in</strong>terneuron (Shilnikov andCymbalyuk 2005).There is no unified nam<strong>in</strong>g scheme for the bifurcations, mostly because they werediscovered and rediscovered <strong>in</strong>dependently <strong>in</strong> many fields and <strong>in</strong> many countries. Forexample, the Andronov-Hopf bifurcation was known to Po<strong>in</strong>care, so some scientistsrefer to it as Po<strong>in</strong>care-Andronov-Hopf bifurcation. Many refer to it as just Hopf bifurcationdue to the fault of no lesser men than the famous Russian mathematicianVladimir Igorevich Arnold and famous French mathematician Rene Thom. Accord<strong>in</strong>gto Arnold’s own accounts, he was visited by Thom <strong>in</strong> the 1960s. While discuss<strong>in</strong>g variousbifurcations, Arnold put too much emphasis on the “recent” Hopf (1942) paper.As a result of Arnold’s misattribution, Thom popularized the bifurcation as be<strong>in</strong>g Hopfbifurcation. In Fig. 6.49 we provide some common alternative names to the bifurcationsconsidered <strong>in</strong> this chapter. The complete list of names of known bifurcations is toolong, and it resembles the list of faculty members of the department of RadioPhysics(Gorky State University, now Nizhnii Novgorod, Russia) founded by A.A. Andronov<strong>in</strong> 1945.The division of bifurcations <strong>in</strong>to subcritical and supercritical ones may be confus<strong>in</strong>gfor a novice. For example, some scientists erroneously th<strong>in</strong>k that supercritical bifurcationsresult <strong>in</strong> appearance of attractors (stable equilibria, limit cycles, etc.) andsubcritical bifurcations result <strong>in</strong> their disappearance. Let us emphasize here that theappearance or disappearance of an equilibrium or a limit cycle depends on the directionof change of the bifurcation parameter. For example, the subcritical pitchfork


Bifurcations 219Figure 6.50: The founder of Russian school of nonl<strong>in</strong>eardynamics, Alexander Aleksandovich Andronov (1901-1952) <strong>in</strong> 1950 (picture provided by M.I. Rab<strong>in</strong>ovich).xx=bstablestable0unstablebunstableFigure 6.51: Transcritical bifurcation <strong>in</strong> ẋ = x(b − x).bifurcation <strong>in</strong> Fig. 6.36 could result <strong>in</strong> the appearance of a stable equilibrium x = 0 ifb decreases past 0. Our classification of bifurcations <strong>in</strong>to subcritical and supercriticalis consistent with the follow<strong>in</strong>g widely accepted rule: Let the bifurcation parameterchange <strong>in</strong> the direction lead<strong>in</strong>g to the <strong>in</strong>crease <strong>in</strong> a number of objects (equilibria, limitcycles). The bifurcation is supercritical if stable objects appear, subcritical if unstableobjects appear, and transcritical, such as <strong>in</strong> Fig. 6.51, if equal numbers of stableand unstable objects appear or disappear. The condition for supercritical (subcritical)Andronov-Hopf bifurcation, Eq. (6.7), is taken from Guckenheimer and Holmes (1983).Delayed loss of stability was first described by Shishkova (1973), and then studied <strong>in</strong>detail by Nejshtadt (1985), though many f<strong>in</strong>d his paper difficult to read. An alternativedescription is given by Arnold et al. (1994) and Baer et al. (1989).Canard (French duck) solutions were reported by Benoit et al. (1981). Due to therecent political climate <strong>in</strong> the USA, some refer to “French ducks” as “freedom ducks”,probably to emphasize that “French = freedom”. Canards <strong>in</strong> R 3 were studied by Benoit(1984), Samborskij (1985, <strong>in</strong> R n ), and recently by Szmolyan and Wechselberger (2001,2004) and Wechselberger (2005).


220 BifurcationsExercises1. (Transcritical bifurcation) Justify the bifurcation diagram shown <strong>in</strong> Fig. 6.51.2. Show that the non-degeneracy and transversality conditions are necessary forthe saddle-node bifurcation. That is, present a system that does not exhibitsaddle-node bifurcation, but satisfies(a) the non-hyperbolicity and non-degeneracy conditions, or(b) the non-hyperbolicity and transversality conditions.3. Consider the model˙V = c(b − b sn ) + a(V − V sn ) 2 ,with positive a and c, and b > b sn . Show that the sojourn time <strong>in</strong> a boundedneighborhood of the po<strong>in</strong>t V = V sn scales asT =π√ac(b − bsn )when b is near b sn . (H<strong>in</strong>t: F<strong>in</strong>d the solution that starts at −∞ and term<strong>in</strong>atesat +∞.)4. Show that the two-dimensional systemthe complex-valued systemand the polar-coord<strong>in</strong>ate systemare equivalent.˙u = c(b)u − ω(b)v + (au − dv)(u 2 + v 2 ) , (6.14)˙v = ω(b)u + c(b)v + (du + av)(u 2 + v 2 ) , (6.15)ż = (c(b) + iω(b))z + (a + id)z|z| 2 ,ṙ = c(b)r + ar 3˙ϕ = ω(b) + dr 2 ,5. Show that the non-degeneracy and transversality conditions are necessary forthe Andronov-Hopf bifurcation. That is, present a system that does not exhibitAndronov-Hopf bifurcation, but satisfies(a) the non-hyperbolicity and non-degeneracy conditions, or(b) the non-hyperbolicity and transversality conditions.6. Show that the system (6.14, 6.15) with c(b) = b, ω(b) = 1, a ≠ 0 and d = 0exhibits Andronov-Hopf bifurcation. Check all three conditions.


Bifurcations 221600K + activation, n0.50.40.3n-nullcl<strong>in</strong>eV-nullcl<strong>in</strong>ecurrent, I400200non-monotoneI-V relation0.20.1action potential00-80 -60 -40 -20 0membrane voltage, V (mV)-200-80 -60 -40 -20 0membrane voltage, V (mV)Figure 6.52: Ex. 8: This I Na,p +I K -model has a non-monotonic I-V relation, yet therest<strong>in</strong>g state becomes unstable via Andronov-Hopf bifurcation before disappear<strong>in</strong>g viasaddle-node bifurcation. Parameters as <strong>in</strong> Fig. 4.1a (Chapter 4) except that E leak =−78 mV and n ∞ (V ) has k = 12 mV.7. Determ<strong>in</strong>e the stability of the limit cycle near an Andronov-Hopf bifurcation.(H<strong>in</strong>t: consider the equilibrium r = √ |c/a| <strong>in</strong> the topological normal form (6.8)).8. The model <strong>in</strong> Fig. 6.52 has a non-monotonic I-V relation. Nevertheless, therest state loses stability via Andronov-Hopf bifurcation before disappear<strong>in</strong>g viasaddle-node bifurcation. Draw representative phase portraits of the model. Isthe system near Bogdanov-Takens bifurcation?9. Consider a generic two-dimensional conductance-based model˙V = I − I(V, x) , (6.16)ẋ = (x ∞ (V ) − x)/τ(V ) , (6.17)where V and x are the membrane voltage and a gat<strong>in</strong>g variable, respectively, Iis the <strong>in</strong>jected dc-current, and I(V, x) is the <strong>in</strong>stantaneous I-V relation, which ofcourse depends on the gat<strong>in</strong>g variable x. Here the membrane capacitance C = 1for the sake of simplicity. Show that the eigenvalues at an equilibrium c ± ω aregiven byc = (I V (V, x) + 1/τ(V ))/2andω = √ c 2 − I ′ ∞(V )/τ(V )


222 Bifurcationsv 2v 11a-11-1Figure 6.53: See Ex. 11.where I ∞ (V ) = I(V, x ∞ (V )) is the steady-state I-V relation of the model. Inparticular, the frequency at the Andronov-Hopf bifurcation iswhere C is the membrane capacitance.10. Determ<strong>in</strong>e when the systemundergoes fold limit cycle bifurcation.(frequency) = √ I ′ ∞(V )/(Cτ(V )) ,z ′ = (a + ωi)z + z|z| 2 − z|z| 4 , z ∈ C ,11. Consider a square neighborhood of a saddle equilibrium <strong>in</strong> Fig. 6.53 (comparewith the <strong>in</strong>set <strong>in</strong> Fig. 6.29). Here v 1 and v 2 are eigenvectors with eigenvaluesλ 2 < 0 < λ 1 . Suppose the limit cycle enters the square at the po<strong>in</strong>t a = τ(I −I b ),where τ > 0 is some parameter. Determ<strong>in</strong>e the amount of time the trajectoryspends <strong>in</strong> the square as a function of I.12. Determ<strong>in</strong>e the bifurcation diagram of the topological normal form (6.13) forsaddle-node homocl<strong>in</strong>ic bifurcation.13. Prove that the systemwith a > 0 undergoes˙v = I + v 2 − u ,˙u = a(bv − u)• saddle-node bifurcation when b 2 = 4I,


Bifurcations 223• Andronov-Hopf bifurcation when a < b and a 2 − 2ab + 4I = 0,• Bogdanov-Takens bifurcation when a = b = 2 √ I.Use results of Ex. 15 to prove that the Andronov-Hopf bifurcation <strong>in</strong> the modelabove is always subcritical.14. Use (6.7) to prove that the relaxation oscillator˙v = f(v) − u˙u = µ(v − b)with an N-shaped fast nullcl<strong>in</strong>e u = f(v) undergoes Andronov-Hopf bifurcationwhen f ′ (b) = 0, i.e., at the knee. The bifurcation is supercritical when f ′′′ (b) < 0and subcritical when f ′′′ (b) > 0.15. Prove that the Andronov-Hopf bifurcation po<strong>in</strong>t <strong>in</strong>˙v = F (v) − u˙v = µ(bv − u)satisfies F ′ = µ and b > µ. Use (6.7) to show thata = {F ′′′ + (F ′′ ) 2 /(b − µ)}/16 .16. Prove that the Andronov-Hopf bifurcation po<strong>in</strong>t <strong>in</strong>˙v = F (v) − u˙u = µ(G(v) − u)satisfies F ′ = µ and G ′ > µ. Use (6.7) to show thata = {F ′′′ + F ′′ (F ′′ − G ′′ )/(G ′ − µ)}/16 .17. Prove that the Andronov-Hopf bifurcation po<strong>in</strong>t <strong>in</strong>˙v = F (v) − (v + 1)u˙u = µ(G(v) − u)satisfies F ′ = µ and G ′ > µ. Use (6.7) to show thata = {F ′′′ + µ − (F ′′ − µ)(1 + µ[G ′′ − F ′′ + 2µ]/ω 2 )}/16 .18. Use (6.7) to show that a two-dimensional relaxation oscillator˙v = F (v, u)˙u = µG(v, u)at an Andronov-Hopf bifurcation po<strong>in</strong>t hasa = 1 { [Fvv G u − F u G vvF vvv + F vv16F u G v− F vuF u]}+ O( √ µ) .


224 Bifurcations19. [M.S.] A leaky <strong>in</strong>tegrate-and-fire model has the same asymptotic fir<strong>in</strong>g rate(1/ln) as a system near saddle homocl<strong>in</strong>ic orbit bifurcation. Explore the possibilitythat <strong>in</strong>tegrate-and-fire models describe neurons near such a bifurcation.20. [M.S.] (blue-sky catastrophe) Prove that˙ϕ = ω, ẋ = a + x 2 , if x = +∞, then x ← −∞, and ϕ ← 0 ,is the canonical model (see Sect. 8.1.5) for blue-sky catastrophe. This modelwithout the reset of ϕ is canonical for the fold limit cycle on homocl<strong>in</strong>ic torusbifurcation. The model with the reset x ← b+s<strong>in</strong> ϕ is canonical for the Lukyanov-Shilnikov bifurcation of a fold limit cycle with non-central homocl<strong>in</strong>ics (Shilnikovand Cymbalyuk 2004, Shilnikov et al. 2005). Here, ϕ is the phase variable onthe unit circle and a and b are bifurcation parameters.21. [M.S.] Def<strong>in</strong>e topological equivalence and the notion of a bifurcation for piecewisecont<strong>in</strong>uous flows.22. [Ph.D.] Use the def<strong>in</strong>ition above to classify codimension-1 bifurcations <strong>in</strong> piecewisecont<strong>in</strong>uous flows.23. [M.S.] The bifurcation sequence <strong>in</strong> Fig. 6.40 seems to be typical <strong>in</strong> 2-dimensionalneuronal models. Develop the theory of Bogdanov-Takens bifurcation with aglobal reentrant orbit.24. [Ph.D.] Develop an automated dynamic clamp protocol (Sharp et al. 1993) thatanalyzes bifurcations <strong>in</strong> neurons <strong>in</strong> vitro, similar to what AUTO, XPPAUT, orMATCONT do <strong>in</strong> models.


Chapter 7Neuronal ExcitabilityNeurons are excitable <strong>in</strong> the sense that they are typically at rest but can fire spikes<strong>in</strong> response to certa<strong>in</strong> forms of stimulation. What k<strong>in</strong>d of stimulation is needed tofire a given neuron? What is the evoked fir<strong>in</strong>g pattern? These are the questionsconcern<strong>in</strong>g the neuron’s computational properties, e.g., whether they are <strong>in</strong>tegrators orresonators, their fir<strong>in</strong>g frequency range, the spike latencies (delays), the co-existence ofrest<strong>in</strong>g and spik<strong>in</strong>g states, etc. From the dynamical systems po<strong>in</strong>t of view, neurons areexcitable because they are near a bifurcation from rest to spik<strong>in</strong>g activity. The type ofbifurcation, and not the ionic currents per se, determ<strong>in</strong>es the computational propertiesof neurons. In this chapter we cont<strong>in</strong>ue our effort to understand the relationshipbetween bifurcations of the rest<strong>in</strong>g state and the neuro-computational properties ofexcitable systems.7.1 ExcitabilityA textbook def<strong>in</strong>ition of neuronal excitability is that a “subthreshold” synaptic <strong>in</strong>putevokes a small graded post-synaptic potential (PSP), while a “superthreshold” <strong>in</strong>putevokes a large all-or-none action potential, which is an order of magnitude larger thanthe amplitude of the subthreshold response. Unfortunately, we cannot adopt thisdef<strong>in</strong>ition to def<strong>in</strong>e excitability of dynamical systems because many systems, <strong>in</strong>clud<strong>in</strong>gsome neuronal models discussed <strong>in</strong> Chap. 4, have neither all-or-none action potentialsnor fir<strong>in</strong>g thresholds. Instead, we employ a purely geometrical def<strong>in</strong>ition.From the geometrical po<strong>in</strong>t of view, a dynamical system with a stable equilibrium isexcitable if there is a large-amplitude piece of trajectory that starts <strong>in</strong> a small neighborhoodof the equilibrium, leaves the neighborhood, and then returns to the equilibrium,as we illustrate <strong>in</strong> Fig. 7.1, left.In the context of neurons, the equilibrium corresponds to the rest<strong>in</strong>g state. Becauseit is stable, all trajectories start<strong>in</strong>g <strong>in</strong> a sufficiently small region of the equilibrium, muchsmaller than the shaded neighborhood <strong>in</strong> the figure, converge back to the equilibrium.Such trajectories correspond to subthreshold PSPs. In contrast, the large trajectory<strong>in</strong> the figure corresponds to fir<strong>in</strong>g a spike. Therefore, superthreshold PSPs are those225


226 Excitabilityspike?spikerestrestexcitableoscillatoryFigure 7.1: Left: An abstract def<strong>in</strong>ition of excitability. There is a spike trajectory thatstarts near a stable equilibrium and returns to it. Right: Excitable systems are nearbifurcations. A modification of the vector field <strong>in</strong> the small shaded region can result<strong>in</strong> a periodic trajectory.that push the state of the neuron to or near the beg<strong>in</strong>n<strong>in</strong>g of the large trajectory(small square <strong>in</strong> Fig. 7.1), thereby <strong>in</strong>itiat<strong>in</strong>g the spike. These <strong>in</strong>puts can be <strong>in</strong>jected byexperimenter via an attached electrode, or they can represent the total synaptic <strong>in</strong>putfrom the other neurons <strong>in</strong> the network, or both.7.1.1 BifurcationsThe def<strong>in</strong>ition <strong>in</strong> Fig. 7.1 is quite general, and it does not make any assumptions regard<strong>in</strong>gthe details of the vector field <strong>in</strong>side or outside of the small shaded neighborhood.Us<strong>in</strong>g the theory presented <strong>in</strong> the previous chapter, we can show that such an excitablesystem is near a bifurcation from rest<strong>in</strong>g to oscillatory dynamics.• Bifurcation of a limit cycle. The vector field <strong>in</strong> the small shaded neighborhoodof the equilibrium can be modified slightly so that the spike trajectory enters thesquare and becomes periodic, as <strong>in</strong> Fig. 7.1, right. That is, the dynamical systemgoes through a bifurcation result<strong>in</strong>g <strong>in</strong> the appearance of a limit cycle.What happens to the stable equilibrium, denoted as “?” <strong>in</strong> the figure? Depend<strong>in</strong>g onthe type of the bifurcation of the limit cycle, the equilibrium may disappear or may losestability. This happens when the limit cycle appears via saddle-node on <strong>in</strong>variant circleor supercritical Andronov-Hopf bifurcations, respectively. Both cases are depicted <strong>in</strong>Fig. 7.2.


Excitability 227excitable bistable oscillatorysaddle-node on <strong>in</strong>variantcircle bifurcationsaddle-homocl<strong>in</strong>icorbit bifurcationsaddlenodebifurcationsupercritical Andronov-Hopfbifurcationfold limit cyclebifurcationsubcriticalAndronov-HopfbifurcationFigure 7.2: Excitable dynamical systems bifurcate <strong>in</strong>to oscillatory ones either directlyor <strong>in</strong>directly, via bistable systems.


228 ExcitabilityAlternatively, the equilibrium may rema<strong>in</strong> stable and co-exist with the newly bornlimit cycle, as it happens dur<strong>in</strong>g saddle homocl<strong>in</strong>ic orbit or fold limit cycle bifurcations<strong>in</strong> Fig. 7.2. The dynamical system is no longer excitable, but bistable, though manyscientists still treat bistable systems as excitable. An appropriate synaptic <strong>in</strong>put canswitch the behavior from rest<strong>in</strong>g to spik<strong>in</strong>g and back. Notice that we considered onlybifurcations of a limit cycle so far.• Bifurcation of the equilibrium. Suppose the system is bistable, as <strong>in</strong> Fig. 7.2.S<strong>in</strong>ce the equilibrium is near the cycle, a small modification of the vector field <strong>in</strong>the shaded neighborhood can make it disappear via saddle-node bifurcation, orlose stability via subcritical Andronov-Hopf bifurcation.In any case, excitable dynamical systems can bifurcate <strong>in</strong>to oscillatory systems eitherdirectly or <strong>in</strong>directly through bistable systems. All these cases are summarized <strong>in</strong>Fig. 7.2.7.1.2 Hodgk<strong>in</strong>’s classificationAs we mentioned <strong>in</strong> the <strong>in</strong>troduction chapter, the first one to study bifurcation mechanismsof excitability (years before mathematicians discovered such bifurcations) wasHodgk<strong>in</strong> (1948), who <strong>in</strong>jected steps of currents of various amplitudes <strong>in</strong>to excitablemembranes and looked at the result<strong>in</strong>g spik<strong>in</strong>g behavior. We illustrate his experiments<strong>in</strong> Fig. 7.3 us<strong>in</strong>g record<strong>in</strong>gs of rat neocortical and bra<strong>in</strong>stem neurons. When the currentstrength is small, the neurons are quiescent. When the current is strong, the neuronsfire tra<strong>in</strong>s of action potentials. Depend<strong>in</strong>g on the average frequency of such fir<strong>in</strong>g,Hodgk<strong>in</strong> identified two major classes of excitability:• Class 1 neural excitability. Action potentials can be generated with arbitrarilylow frequency, depend<strong>in</strong>g on the strength of the applied current.• Class 2 neural excitability. Action potentials are generated <strong>in</strong> a certa<strong>in</strong>frequency band that is relatively <strong>in</strong>sensitive to changes <strong>in</strong> the strength of theapplied current.Class 1 neurons, sometimes called type I neurons, fire with a frequency that may varysmoothly over a broad range of about 2 to 100 Hz or even higher. The important observationhere is that the frequency can be changed tenfold. In contrast, the frequencyband of Class 2 neurons is quite limited, e.g., 150−200 Hz, but it can vary from neuronto neuron. The exact numbers are not important to us here. The qualitative dist<strong>in</strong>ctionbetween the classes noticed by Hodgk<strong>in</strong> is that the frequency-current relation (theF-I curve <strong>in</strong> Fig. 7.3, bottom) starts from zero and cont<strong>in</strong>uously <strong>in</strong>creases for Class 1neurons, but is discont<strong>in</strong>uous for Class 2 neurons.Obviously, the two classes of excitability have different neuro-computational properties.Class 1 excitable neurons can smoothly encode the strength of an <strong>in</strong>put, e.g.,


Excitability 229Layer 5 pyramidal cellBra<strong>in</strong>stem mesV cellI=320I=1000I=80I=600I=60I=40I=20100 ms 40 mVI=500I=400I=20020 ms 20 mVI (pA)I (pA)0 pA0 pAClass 1 excitabilityClass 2 excitabilityasymptotic fir<strong>in</strong>g frequency, Hz40302010F-I curve00 100 200 300<strong>in</strong>jected dc-current, I (pA)asymptotic fir<strong>in</strong>g frequency, Hz25020015010050F-I curve00 500 1000 1500<strong>in</strong>jected dc-current, I (pA)Figure 7.3: Top: Typical responses of membrane potentials of two neurons to steps ofdc-current of various magnitudes I. Bottom: Correspond<strong>in</strong>g frequency-current (F-I)relations are qualitatively different. Shown are record<strong>in</strong>gs of layer 5 pyramidal neuronsfrom rat primary visual cortex (left) and mesV neuron from rat bra<strong>in</strong>stem (right). Theasymptotic frequency is 1000/T ∞ , where T ∞ is taken to be the <strong>in</strong>terval between thelast two spikes <strong>in</strong> a long spike tra<strong>in</strong>.


230 Excitability50 ms 20 mV 1 ms1000 pA-60 mV0 pA 500 pAFigure 7.4: Class 3 excitability of a mesV neuron of rat bra<strong>in</strong>stem (contrast withFig. 7.3).20 mV100 ms-20mV50 ms 3 mVsubthreshold oscillations700 pA100 pAFigure 7.5: Class 3 excitability of a layer 5 pyramidal neuron of rat visual cortex. The<strong>in</strong>set shows subthreshold oscillations of membrane potential.the strength of the applied dc-current or the strength of the <strong>in</strong>com<strong>in</strong>g synaptic bombardment,<strong>in</strong>to the frequency of their spik<strong>in</strong>g output. Class 2 neurons cannot do that.Instead, they can act as threshold elements report<strong>in</strong>g when the strength of <strong>in</strong>put isabove a certa<strong>in</strong> value. Both properties are important <strong>in</strong> neural computations.Hodgk<strong>in</strong> also observed that axons left <strong>in</strong> oil or sea water for long periods exhibited• Class 3 neural excitability. A s<strong>in</strong>gle action potential is generated <strong>in</strong> responseto a pulse of current. Repetitive (tonic) spik<strong>in</strong>g can be generated only forextremely strong <strong>in</strong>jected currents or not at all.Two examples of Class 3 excitable systems are depicted <strong>in</strong> Fig. 7.4 and Fig. 7.5. ThemesV neuron <strong>in</strong> the figure fires a phasic spike at the onset of the pulse of current, andthen rema<strong>in</strong>s quiescent. Even <strong>in</strong>ject<strong>in</strong>g pulses as high as 1000 pA, which result <strong>in</strong> spike


Excitability 231layer 5 pyramidal cellbra<strong>in</strong>stem mesV celltransition20 mVtransition-60 mV-50 mV200 pA3000 pA0 pA500 ms0 pA500 msFigure 7.6: As the magnitude of <strong>in</strong>jected dc-current <strong>in</strong>creases, the neurons bifurcatesfrom rest<strong>in</strong>g to repetitive spik<strong>in</strong>g behavior. Shown are record<strong>in</strong>gs of the same neuronsas <strong>in</strong> Fig. 7.3. Notice that the ratio of the first and last <strong>in</strong>terspike <strong>in</strong>tervals of thepyramidal cell is much greater than that <strong>in</strong> mes V neuron.tra<strong>in</strong>s <strong>in</strong> another mesV neuron <strong>in</strong> Fig. 7.3, cannot evoke multiple spikes <strong>in</strong> this neuron.Similarly, the pyramidal neuron <strong>in</strong> Fig. 7.5 cannot susta<strong>in</strong> tonic spik<strong>in</strong>g even when the<strong>in</strong>jected current is ten times stronger than the neuron’s rheobase. Ironically, neuronsexhibit<strong>in</strong>g such a behavior would most likely be discarded as “sick” or “unhealthy”,though the neurons analyzed <strong>in</strong> the figures looked normal from any other po<strong>in</strong>t of view.We will study the dynamic mechanism of this class of excitability and show that it mayhave noth<strong>in</strong>g to do with sickness.It will be clear shortly that this classification is of limited value except that itpo<strong>in</strong>ts to the fact that neurons should be dist<strong>in</strong>guished accord<strong>in</strong>g not only to ionicmechanisms of excitability, but also to dynamic mechanisms, <strong>in</strong> particular, to the typeof bifurcation of the rest state.7.1.3 Classes 1 and 2Let us consider the strength of the applied current <strong>in</strong> Hodgk<strong>in</strong>’s experiments as be<strong>in</strong>ga bifurcation parameter. Instead of chang<strong>in</strong>g the parameter abruptly, as <strong>in</strong> Fig. 7.3,we change it slowly <strong>in</strong> Fig. 7.6 us<strong>in</strong>g record<strong>in</strong>gs of the same neurons as <strong>in</strong> the previousfigure. In Sect. 7.1.5 we expla<strong>in</strong> the fundamental difference between these two protocols.When the current ramps up, the rest potential <strong>in</strong>creases until a bifurcation occurs,result<strong>in</strong>g <strong>in</strong> loss of stability or disappearance of the equilibrium correspond<strong>in</strong>g to therest state, and the neuron activity becomes oscillatory. Notice that the pyramidalneuron <strong>in</strong> Fig. 7.6 starts to fire with a small frequency, which then <strong>in</strong>creases accord<strong>in</strong>gto the F-I curve <strong>in</strong> Fig. 7.3 (a slower current ramp is needed to span the entire frequencyrange of the F-I curve). In contrast, the bra<strong>in</strong>stem neuron starts to fire with a highfrequency that rema<strong>in</strong>s relatively constant even though the magnitude of the <strong>in</strong>jectedcurrent <strong>in</strong>creases.


232 ExcitabilityClass 3 excitable neuron0 mV10 mV50 ms4,000 pA-60 mV0 pAFigure 7.7: A Class 3 excitable bra<strong>in</strong>stem mesV neuron does not fire <strong>in</strong> response to aramp current, even though the <strong>in</strong>jected current is stronger than the one <strong>in</strong> Fig. 7.4.Among all four codimension-1 bifurcations of equilibrium, discussed <strong>in</strong> the previouschapter and mentioned <strong>in</strong> Fig. 7.2, only saddle-node on <strong>in</strong>variant circle bifurcationresults <strong>in</strong> a limit cycle attractor with arbitrary small frequency and cont<strong>in</strong>uous F-Icurve. The other three bifurcations result <strong>in</strong> limit cycle attractors with relatively largefrequencies and discont<strong>in</strong>uous F-I curves. Therefore,• Class 1 neural excitability corresponds to the rest state disappear<strong>in</strong>g viasaddle-node on <strong>in</strong>variant circle bifurcation.• Class 2 neural excitability corresponds to the rest state disappear<strong>in</strong>g viasaddle-node (off <strong>in</strong>variant circle) bifurcation or los<strong>in</strong>g stability via subcritical orsupercritical Andronov-Hopf bifurcations.Of course, the rest state can lose stability or disappear via other bifurcations hav<strong>in</strong>ghigher codimension, sometimes lead<strong>in</strong>g to counter-<strong>in</strong>tuitive results (e.g., Class 1 excitabilitynear Andronov-Hopf bifurcation; see Ex. 6 and Sect. 7.2.11). In this chapterwe concentrate on the four bifurcations above because they have the lowest codimensionand hence are the most likely to be seen experimentally.7.1.4 Class 3In Fig. 7.7 we <strong>in</strong>ject a slow ramp current <strong>in</strong>to the Class 3 excitable system. In contrastto Fig. 7.6, no spik<strong>in</strong>g and no bifurcation occurs <strong>in</strong> this experiment despite the factthat the membrane potential goes all the way to 0 mV. Therefore,• Class 3 neural excitability occurs when the rest<strong>in</strong>g state rema<strong>in</strong>s stablefor any fixed I <strong>in</strong> a biophysically relevant range.Then, why are there s<strong>in</strong>gle spikes <strong>in</strong> Fig. 7.4? Their existence <strong>in</strong> the figure and theirabsence <strong>in</strong> the ramp experiment is related to the phenomenon of accommodation thatwe show now describe.


Excitability 2330.250.20.15w0.10.05w-nullcl<strong>in</strong>eI=0.03V-nullcl<strong>in</strong>eI=0I=00.060.0450.030.0150.0100.4 0.2 0 0.2 0.4 0.6 0.8VFigure 7.8: Class 3 excitability <strong>in</strong> FitzHugh-Nagumo model (4.11, 4.12) with a =0.1, b = 0.01, c = 0. The model fires a s<strong>in</strong>gle spike for any pulse of current.Let us consider a neuron hav<strong>in</strong>g a transient Na + current with relatively fast <strong>in</strong>activation.If a sufficiently slow ramp of current is <strong>in</strong>jected, the current has enough timeto <strong>in</strong>activate and no action potentials could be generated. Such a neuron accommodatesto the slow ramp. In contrast, a quick membrane depolarization due to a strongstep of current does not give enough time for Na + <strong>in</strong>activation, thereby result<strong>in</strong>g <strong>in</strong> aspike. Dur<strong>in</strong>g the spike, the current <strong>in</strong>activates quickly and precludes any further actionpotentials. Instead of <strong>in</strong>activat<strong>in</strong>g Na + current, we could have used low-thresholdpersistent K + current, or any other resonant current, to illustrate the phenomenon ofaccommodation.From the dynamical systems po<strong>in</strong>t of view, slow ramp results <strong>in</strong> quasi-static dynamicsso that all gat<strong>in</strong>g variables follow their steady-state values, x = x ∞ (V ), and themembrane potential follows its I-V curve. As long as the equilibrium correspond<strong>in</strong>g tothe rest<strong>in</strong>g state is stable, the neuron is at rest. Even global bifurcations result<strong>in</strong>g <strong>in</strong>the appearance of stable limit cycles do not change that. Only when the equilibriumbifurcates (loses stability or disappears), the neuron changes its behavior, e.g., jumpsto a limit cycle attractor and starts to fire spikes. Class 3 excitable systems do not fire<strong>in</strong> response to slow ramps because the rest<strong>in</strong>g state does not bifurcate.In contrast, a pulse of current changes the phase portrait <strong>in</strong> a rather abrupt manner,as we illustrate <strong>in</strong> Fig. 7.8 us<strong>in</strong>g the FitzHugh-Nagumo model with vertical slownullcl<strong>in</strong>e. Though no bifurcation can occur <strong>in</strong> the model, and the rest<strong>in</strong>g state is stablefor any value of I, its location suddenly shifts when I jumps. The trajectory fromthe old equilibrium, (0, 0), to the new one goes through the right branch of the cubicV -nullcl<strong>in</strong>e thereby result<strong>in</strong>g <strong>in</strong> a s<strong>in</strong>gle spike. S<strong>in</strong>ce the new equilibrium (0, 0.03) isa global attractor and no limit cycles exist, periodic spik<strong>in</strong>g cannot be generated. InEx. 7 we explore the relationship between Class 3 excitability and Andronov-Hopf bifurcation(notice the subthreshold oscillations of membrane potential of the pyramidalneuron <strong>in</strong> Fig. 7.5). We see that <strong>in</strong>ject<strong>in</strong>g ramps of current is not equivalent to <strong>in</strong>ject-


234 Excitability0.20.150.10.05I 1I 1I 1K + gat<strong>in</strong>g variable, nI 0I 00-70 -60 -50 -40 -70 -60 -50 -40 -70 -60 -50 -40membrane potential, V (mV) membrane potential, V (mV) membrane potential, V (mV)0-20-40-60-80V(t)I 1 I 1II(t)1I 0I 0slow ramp from I 0 to I 1step from I 0 to I 1 shock pulse at I 1(a) (b) (c)Figure 7.9: The difference between ramp, step, and shock stimulation is <strong>in</strong> the resett<strong>in</strong>gof <strong>in</strong>itial condition.<strong>in</strong>g pulses of current. The system goes through a bifurcation of the equilibrium <strong>in</strong> theformer, but may bypass it and jump somewhere else <strong>in</strong> the latter.7.1.5 Ramps, steps, and shocksIn Fig. 7.9 we elaborate the difference between <strong>in</strong>ject<strong>in</strong>g slow ramps, steps, and shocks(i.e., brief pulses) of current. In the first two cases the magnitude of the <strong>in</strong>jected currentchanges from I 0 to I 1 , while <strong>in</strong> the third case the current is I 1 except the <strong>in</strong>f<strong>in</strong>itesimallybrief moment when it has an <strong>in</strong>f<strong>in</strong>itely large strength. In all three cases the dynamicsof the model can be understood via analysis of its phase portrait at I = I 1 . The keydifference among the stimulation protocols is how they reset the <strong>in</strong>itial condition.At the beg<strong>in</strong>n<strong>in</strong>g of the slow ramp <strong>in</strong> Fig. 7.9a, the state of the neuron is at thestable equilibrium. As the current slowly <strong>in</strong>creases, the equilibrium slowly moves, andthe trajectory follows it. When the current reaches I = I 1 , the trajectory is at the newequilibrium, so no response is evoked because the equilibrium is stable. In contrast,when the current is stepped from I 0 to I 1 <strong>in</strong> Fig. 7.9b, the location of the equilibriumchanges <strong>in</strong>stantaneously, but the membrane potential and the gat<strong>in</strong>g variables do nothave the time to catch up. To understand the response of the model to the step, we


Excitability 235(a)(b)frequencypulsesrampfrequency-current (F-I) relation3.88subcriticalAndronov-Hopfbifurcationhomocl<strong>in</strong>ic5.25Andronov-Hopfcurrent, I(c)(d)0.7I=5.25near subcriticalAndronov-Hopfbifurcation(e)0.7I=3.8866near saddlehomocl<strong>in</strong>ic orbitbifurcationK+ gat<strong>in</strong>g variable, nK+ gat<strong>in</strong>g variable, n0.1-80 membrane potential, V (mV) 00.1-80 membrane potential, V (mV) 0Figure 7.10: The I Na,p +I K -model undergoes subcritical Andronov-Hopf bifurcation yetcan exhibit low-frequency fir<strong>in</strong>g when pulses (but not ramps) of current are <strong>in</strong>jected.Parameters: C = 1, I = 0, E L = −66.2, g L = 2, g Na = 5, g K = 4.5, m ∞ (V ) hasV 1/2 = −30 and k = 10, n ∞ (V ) has V 1/2 = −34 and k = 13, and τ(V ) = 1, E Na = 60mV and E K = −90 mV. The shaded region denotes the attraction doma<strong>in</strong> of the rest<strong>in</strong>gstate. The <strong>in</strong>set shows a distorted draw<strong>in</strong>g of the phase portrait.need to consider its dynamics at I = I 1 with the <strong>in</strong>itial condition set to the locationof the old equilibrium marked by the white square <strong>in</strong> the figure. Such a step evokes aspike response even though the new equilibrium is stable. F<strong>in</strong>ally, shock<strong>in</strong>g the neuronresults <strong>in</strong> an <strong>in</strong>stantaneous <strong>in</strong>crease of its membrane potential to a new value. As anexercise, prove that the magnitude of the <strong>in</strong>crease equals the product of pulse widthand pulse height divided by the membrane capacitance. This shifts the <strong>in</strong>itial conditionhorizontally to a new po<strong>in</strong>t, marked by the white square <strong>in</strong> Fig. 7.9c, and results <strong>in</strong> aspike response.Now, let us revisit the Hodgk<strong>in</strong> experiments and demonstrate the fundamental differencebetween the stimulation protocols. In Fig. 7.10a,b,c we simulate the I Na,p +I K -model and show that it is Class 2 excitable <strong>in</strong> response to ramps of current but Class1 excitable <strong>in</strong> response to steps of current. The apparent contradiction is resolved <strong>in</strong>


236 Excitabilitysaddle-node bifurcationsubcritical Andronov-Hopf bifurcation<strong>in</strong>hibitorypulseattractiondoma<strong>in</strong>n-nullcl<strong>in</strong>espik<strong>in</strong>g limit cycleexcitatorypulsenV-nullcl<strong>in</strong>eVV-nullcl<strong>in</strong>eexc.pulsespik<strong>in</strong>g limit cycleunstablen-nullcl<strong>in</strong>e<strong>in</strong>hibitorypulseattractiondoma<strong>in</strong>of spik<strong>in</strong>glimit cycleFigure 7.11: Co-existence of stable equilibrium and spik<strong>in</strong>g limit cycle attractor <strong>in</strong> theI Na,p +I K -model. Left: The rest state is about to disappear via saddle-node bifurcation.Right: The rest state is about to lose stability via subcritical Andronov-Hopfbifurcation. Right (left) arrows denote the location and the direction of an excitatory(<strong>in</strong>hibitory) pulse that switches spik<strong>in</strong>g behavior to rest<strong>in</strong>g.Fig. 7.10d and e, where we consider the model’s phase portraits. Notice the co-existenceof the rest<strong>in</strong>g state and a limit cycle attractor. The rest<strong>in</strong>g state loses stability via subcriticalAndronov-Hopf bifurcation at I = 5.25, so the emerg<strong>in</strong>g spik<strong>in</strong>g has non-zerofrequency at I ≈ 5.25. However, <strong>in</strong>ject<strong>in</strong>g steps of current results <strong>in</strong> transitions to thelimit cycle even before the rest<strong>in</strong>g state loses its stability. The limit cycle <strong>in</strong> the modelappears via saddle homocl<strong>in</strong>ic orbit bifurcation at I ≈ 3.8866, its period is quite largeresult<strong>in</strong>g <strong>in</strong> the Class 1 response to steps of current. The F-I curves for homocl<strong>in</strong>icbifurcations have logarithmic scal<strong>in</strong>g, so small frequency oscillations are difficult tocatch numerically let alone experimentally.The surpris<strong>in</strong>g discrepancy <strong>in</strong> Fig. 7.10a occurs because the rest<strong>in</strong>g state of theI Na,p +I K -model is near the Bogdanov-Takens bifurcation, i.e., the model is near atransition from resonator to <strong>in</strong>tegrator. Such a bifurcation was recorded, though <strong>in</strong>directly,<strong>in</strong> some neocortical pyramidal neurons, as we will show later <strong>in</strong> this chapter and<strong>in</strong> Chap. 8. Another surpris<strong>in</strong>g example of Andronov-Hopf bifurcation with Class 1excitability is presented <strong>in</strong> Ex. 6. To avoid such surprises, we adopt the ramp def<strong>in</strong>itionof excitability throughout the book.7.1.6 BistabilityWhen transition from rest<strong>in</strong>g to spik<strong>in</strong>g states occurs via saddle-node (off <strong>in</strong>variantcircle) or subcritical Andronov-Hopf bifurcation, there is a co-existence of a stableequilibrium and a stable limit cycle attractor just before the bifurcation, as we illustrate<strong>in</strong> Fig. 7.11. We refer to such systems as bistable. They have a remarkable


Excitability 23720 mV50 msaverage fir<strong>in</strong>g frequency, Hz25020015010050F-I curveClass 2 excitabilitynoisewithwithout noise00 500 1000 1500<strong>in</strong>jected dc-current, I (pA)Figure 7.12: Examples of noise-<strong>in</strong>duced low-frequency fir<strong>in</strong>gs of Class 2 excitable system.The F-I curve may look like the one for Class 1 excitability. Shown are record<strong>in</strong>gsof bra<strong>in</strong>stem mesV neuron.neuro-computational property: Bistable systems can be switched from one state tothe other by an appropriately timed brief stimulus. R<strong>in</strong>zel (1978) predicted such abehavior <strong>in</strong> the Hodgk<strong>in</strong>-Huxley model, and then bistability and hysteresis were foundexperimentally <strong>in</strong> the squid axon (Guttman et al. 1980). What was really surpris<strong>in</strong>gfor many neuroscientists is that neurons can be switched from repetitive spik<strong>in</strong>g torest<strong>in</strong>g by brief depolariz<strong>in</strong>g shock-stimuli.This phenomenon is illustrated <strong>in</strong> Fig. 7.11. Each shaded area <strong>in</strong> the figure denotesthe attraction doma<strong>in</strong> of a spik<strong>in</strong>g limit cycle attractor. Obviously, the state of therest<strong>in</strong>g neuron must be pushed <strong>in</strong>to the shaded area to <strong>in</strong>itiate periodic spik<strong>in</strong>g. Similarly,the state of the periodically spik<strong>in</strong>g neuron must be pushed out of the shadedarea to stop the spik<strong>in</strong>g. As the arrows <strong>in</strong> the figure <strong>in</strong>dicate, both excitatory and <strong>in</strong>hibitorystimuli can do that, depend<strong>in</strong>g on their tim<strong>in</strong>g relative to the phase of spik<strong>in</strong>goscillation. This protocol can be used to test bistability experimentally.Bistable behavior reveals itself <strong>in</strong>directly when a neuron is kept close to the bifurcation,e.g., when the <strong>in</strong>jected dc-current is just below the rheobase. Noisy perturbationscan switch the neuron from rest<strong>in</strong>g to spik<strong>in</strong>g states thereby creat<strong>in</strong>g an irregular spiketra<strong>in</strong> consist<strong>in</strong>g of short bursts of spikes. Such stutter<strong>in</strong>g spik<strong>in</strong>g have been observed<strong>in</strong> many neurons, <strong>in</strong>clud<strong>in</strong>g some regular spik<strong>in</strong>g (RS) and fast spik<strong>in</strong>g (FS) neocorticalneurons, as we discuss <strong>in</strong> Chap. 8. The mean fir<strong>in</strong>g frequency dur<strong>in</strong>g stutter<strong>in</strong>gis proportional to the amplitude of the <strong>in</strong>jected current and it can be quite low evenfor Class 2 excitable system, as we illustrate <strong>in</strong> Fig. 7.12. Thus, caution should be


238 Excitability(a) rest<strong>in</strong>g spik<strong>in</strong>g (b) spik<strong>in</strong>g rest<strong>in</strong>g (c)Class 2 Class 1800F-I curveI=4.51spik<strong>in</strong>gI=3.08frequency, Hz(b)Class 1Class 2(a)rest<strong>in</strong>gspik<strong>in</strong>grest<strong>in</strong>g02 4 6current, IFigure 7.13: (a) The frequency of emerg<strong>in</strong>g oscillations at the transition “rest<strong>in</strong>g →spik<strong>in</strong>g” def<strong>in</strong>es the class of excitability. (b) The frequency of disappear<strong>in</strong>g oscillationsat the transition “spik<strong>in</strong>g → rest<strong>in</strong>g def<strong>in</strong>es the class of spik<strong>in</strong>g. (c) The I Na,p +I K -model with high-threshold K + current exhibits class 2 excitability but class 1 spik<strong>in</strong>g.Its F-I curve has a hysteresis.used when determ<strong>in</strong><strong>in</strong>g experimentally the class of excitability; only spike tra<strong>in</strong>s withregular <strong>in</strong>terspike periods should be accepted to measure the F-I relations.7.1.7 Class 1 and 2 spik<strong>in</strong>gThe class of excitability is determ<strong>in</strong>ed by the frequency of emerg<strong>in</strong>g oscillations atthe transition “rest<strong>in</strong>g → spik<strong>in</strong>g”, as <strong>in</strong> Fig. 7.13a. Let us look at the frequencyof disappear<strong>in</strong>g oscillations at the transition “spik<strong>in</strong>g → rest<strong>in</strong>g”. To <strong>in</strong>duce such atransition, we <strong>in</strong>ject a strong pulse of dc-current of slowly decreas<strong>in</strong>g amplitude, as <strong>in</strong>Fig. 7.13b. Similarly to the Hodgk<strong>in</strong>’s classification of excitability, we say that a neuronhas Class 1 spik<strong>in</strong>g if the frequency-current (F-I) curve at the transition “spik<strong>in</strong>g →rest<strong>in</strong>g” decreases to zero, as <strong>in</strong> Fig. 7.13c, and Class 2 spik<strong>in</strong>g if it stops at a certa<strong>in</strong>non-zero value.The class of excitability co<strong>in</strong>cides with the class of spik<strong>in</strong>g when the transitions“rest<strong>in</strong>g ↔ spik<strong>in</strong>g” occur via saddle-node on <strong>in</strong>variant circle bifurcation or supercriticalAndronov-Hopf bifurcation. Indeed, if the current ramps are sufficiently slow, theneuron as a dynamical system goes through the same bifurcation, just <strong>in</strong> the oppositedirections. The classes may differ when the bifurcation is of the saddle-node (off <strong>in</strong>variantcircle) type or subcritical Andronov-Hopf type because of the bistability of therest<strong>in</strong>g and spik<strong>in</strong>g states. Such a bistability results <strong>in</strong> the hysteresis behavior of thesystem when the <strong>in</strong>jected current I <strong>in</strong>creases and decreases slowly, which may result<strong>in</strong> the hysteresis of the F-I curve. For example, the transition “rest<strong>in</strong>g → spik<strong>in</strong>g” <strong>in</strong>Fig. 7.13a occurs via saddle-node bifurcation at I = 4.51, and the frequency of spik<strong>in</strong>gequals the frequency of the limit cycle attractor, which is non-zero at this value of I.Decreas<strong>in</strong>g I results <strong>in</strong> the transition “spik<strong>in</strong>g → rest<strong>in</strong>g” via the saddle homocl<strong>in</strong>ic


Excitability 239co-existence of rest<strong>in</strong>g and spik<strong>in</strong>g statesYES(bistable)NO(monostable)subthreshold oscillationsNO(<strong>in</strong>tegrator)YES(resonator)saddle-nodesubcriticalAndronov-Hopfsaddle-node on<strong>in</strong>variant circlesupercriticalAndronov-HopfFigure 7.14: Classification of neurons <strong>in</strong>tomonostable/bistable <strong>in</strong>tegrators/resonatorsaccord<strong>in</strong>g to the bifurcation of the rest<strong>in</strong>gstate.orbit bifurcation <strong>in</strong> Fig. 7.13b, and <strong>in</strong> the oscillations with zero frequency at I = 3.08.Thus, the F-I behavior of the model <strong>in</strong> this figure (and <strong>in</strong> Fig. 7.10) exhibits Class 2excitability but Class 1 spik<strong>in</strong>g. Because of the logarithmic scal<strong>in</strong>g of the F-I curve atthe saddle homocl<strong>in</strong>ic bifurcation (see Sect. 6.2.4), estimat<strong>in</strong>g experimentally the zerovalue of the F-I curves is challeng<strong>in</strong>g.Interest<strong>in</strong>gly, steps of <strong>in</strong>jected dc-current, as <strong>in</strong> Fig. 7.10c, <strong>in</strong>duce the transition“rest<strong>in</strong>g → spik<strong>in</strong>g”. But because the model <strong>in</strong> the figure is near codimension-2Bogdanov-Takens bifurcation, the steps test the frequency of the limit cycle attractorat the bifurcation “spik<strong>in</strong>g → rest<strong>in</strong>g”, as <strong>in</strong> Fig. 7.10e; that is, they test the classof spik<strong>in</strong>g! The F-I curve <strong>in</strong> response to steps <strong>in</strong> the figure is the same as the F-I curve<strong>in</strong> response to a slowly decreas<strong>in</strong>g current ramp. As an exercise, expla<strong>in</strong> why this istrue for Fig. 7.10 but not for Fig. 7.13.To summarize, we def<strong>in</strong>e the class of excitability accord<strong>in</strong>g to the frequency ofemerg<strong>in</strong>g spik<strong>in</strong>g of a neuron <strong>in</strong> response to a slowly <strong>in</strong>creas<strong>in</strong>g current ramp. The classof excitability corresponds to a bifurcation of the rest<strong>in</strong>g state (equilibrium) result<strong>in</strong>g<strong>in</strong> the transition “rest<strong>in</strong>g → spik<strong>in</strong>g”. We def<strong>in</strong>e the class of spik<strong>in</strong>g accord<strong>in</strong>g to thefrequency of disappear<strong>in</strong>g spik<strong>in</strong>g of a neuron <strong>in</strong> response to a slowly decreas<strong>in</strong>g currentramp. The class of spik<strong>in</strong>g corresponds to the bifurcation of the limit cycle result<strong>in</strong>g <strong>in</strong>the transition “spik<strong>in</strong>g → rest<strong>in</strong>g”. Stimulat<strong>in</strong>g a neuron with the ramps (and pulses)is the first step <strong>in</strong> explor<strong>in</strong>g the bifurcations <strong>in</strong> the neuron dynamics. Comb<strong>in</strong>ed withthe test for the existence of subthreshold oscillations of the membrane potential, it tellswhether the neuron is an <strong>in</strong>tegrator or a resonator, and whether it is monostable orbistable, as we discuss next.7.2 Integrators vs. ResonatorsIn this book we classify excitable systems based on two features: the co-existence ofrest<strong>in</strong>g and spik<strong>in</strong>g states and the existence of subthreshold oscillations. The formerfeature divides all systems <strong>in</strong>to monostable and bistable. The latter feature dividesall systems <strong>in</strong>to <strong>in</strong>tegrators (no oscillations) and resonators. These features determ<strong>in</strong>e


240 Excitabilityproperties <strong>in</strong>tegrators resonatorsbifurcationsaddle-node on<strong>in</strong>variant circlesaddle-nodesubcriticalAndronov-HopfsupercriticalAndronov-Hopfexcitability class 1class 2 class 2 class 2oscillatorypotentialsfrequencypreferencenonoyesyesI-V relationat restnon-monotonemonotonespike latency large smallthresholdand rheobasewell-def<strong>in</strong>edmay not be def<strong>in</strong>edall-or-noneaction potentials yes noco-existence ofrest<strong>in</strong>g and spik<strong>in</strong>gpost-<strong>in</strong>hibitory spikeor facilitation(brief stimuli)<strong>in</strong>hibition-<strong>in</strong>ducedspik<strong>in</strong>gno yes yes nonoyesnopossibleFigure 7.15: Summary of neuro-computational properties.uniquely the type of bifurcation of the rest<strong>in</strong>g state, as we summarize <strong>in</strong> Fig. 7.14. Forexample, a bistable <strong>in</strong>tegrator corresponds to a saddle-node bifurcation whereas monostableresonator corresponds to a supercritical Andronov-Hopf bifurcation. Integratorsand resonators have drastically different neuro-computational properties, summarized<strong>in</strong> Fig. 7.15 and discussed next (the I-V curves are discussed <strong>in</strong> the previous chapter).7.2.1 Fast subthreshold oscillationsAccord<strong>in</strong>g to the def<strong>in</strong>ition, resonators have oscillatory potentials whereas <strong>in</strong>tegratorsdo not. This feature is so important that many of the other neuronal properties discussedlater are just mere consequences of the existence or absence of such oscillations.Fast subthreshold oscillations, as <strong>in</strong> Fig. 7.16, are typically due to a fast lowthresholdpersistent K + current. At rest, there is a balance of all <strong>in</strong>ward currentsand this partially activated K + current. A brief depolarization further activates K +current and results <strong>in</strong> fast after-hyperpolarization. While the cell is hyperpolarized,the current de-activates below its steady state level, the balance is shifted toward the<strong>in</strong>ward currents, and the membrane potential depolarizes aga<strong>in</strong>, and so on.


Excitability 241K + activation, n10.80.60.4damped oscillationsn-nullcl<strong>in</strong>eV-nullcl<strong>in</strong>e0.20-80 -60 -40 -20 0membrane potential, V (mV)bra<strong>in</strong>stem (mes V neurons)Hodgk<strong>in</strong>-Huxley model9 ms2 mV12.5 ms1 mV-49 mVbra<strong>in</strong>stem (mes V neurons)susta<strong>in</strong>ed noisy oscillationsmammalian cortical layer 4 cell-40 mV-46 mV10 ms-43 mV10 mV100 ms-50 mV1 mV-47 mV-68 mVmitral cells <strong>in</strong> olfactory bulbthalamocortical neurons-59 mV-62 mV-65 mV5 mV20 msFigure 7.16: Examples of fast damped (top) or susta<strong>in</strong>ed (bottom) subthreshold oscillationsof membrane potential <strong>in</strong> neurons and their voltage dependence. Modified fromIzhikevich et al. (2003).


242 ExcitabilityThe existence of fast subthreshold oscillatory potentials is a dist<strong>in</strong>guishable featureof neurons near Andronov-Hopf bifurcation. Indeed, the rest<strong>in</strong>g state of such a neuronis a stable focus. When stimulated by a brief synaptic <strong>in</strong>put or an <strong>in</strong>jected pulse ofcurrent, the state of the system deviates from the focus equilibrium, and then returnsto the equilibrium along a spiral trajectory, as depicted <strong>in</strong> Fig. 7.16, top, therebyproduc<strong>in</strong>g a damped oscillation. The frequency of such an oscillation is the imag<strong>in</strong>arypart of the complex-conjugate eigenvalues at the equilibrium (see Sect. 6.1.3), and itcan be as large as 200 Hz <strong>in</strong> mammalian neurons.In Ex. 3 we prove that noise can make such oscillations susta<strong>in</strong>ed. While the stateof the system is perturbed and returns to the focus equilibrium, another strong randomperturbation may push it away from the equilibrium, thereby start<strong>in</strong>g a new dampedoscillation. As a result, persistent noisy perturbations create a random sequence ofdamped oscillations and do not let the neuron rest. The membrane potential of sucha neuron exhibits noisy susta<strong>in</strong>ed oscillations of small amplitude depicted <strong>in</strong> Fig. 7.16and discussed <strong>in</strong> Sect. 6.1.4.Injected dc-current or background synaptic noise <strong>in</strong>crease the rest potential, changeits eigenvalues and, hence change the frequency and amplitude of noisy oscillations.Fig. 7.16 depicts typical cases when the frequency and the amplitude <strong>in</strong>crease as therest<strong>in</strong>g state becomes more depolarized.One should be careful to dist<strong>in</strong>guish fast and slow subthreshold oscillations of membranepotential. Fast oscillations, as <strong>in</strong> Fig. 7.16, are those hav<strong>in</strong>g period comparablewith the membrane time constant or with the period of repetitive spik<strong>in</strong>g. In contrast,some neurons found <strong>in</strong> entorh<strong>in</strong>al cortex, <strong>in</strong>ferior olive, hippocampus, thalamus, andmany other bra<strong>in</strong> regions can exhibit slow subthreshold oscillations with the periodof 100 ms and more. These oscillations reflect the <strong>in</strong>terplay between fast and slowmembrane currents, e.g., I h or I T , and may be irrelevant to the bifurcation mechanismof excitability. We will discuss this issue <strong>in</strong> detail <strong>in</strong> Sect. 7.3.3 and <strong>in</strong> Chap. 9. Amaz<strong>in</strong>gly,such neurons still possess many neuro-computational properties of resonators,such as frequency preference and rebound spik<strong>in</strong>g, but exhibit these properties on aslower time scale.7.2.2 Frequency preference and resonanceA standard experimental procedure to test the propensity of a neuron to subthresholdoscillations is to stimulate it with a s<strong>in</strong>usoidal current hav<strong>in</strong>g slowly <strong>in</strong>creas<strong>in</strong>g frequency,called a zap current, as <strong>in</strong> Fig. 7.17. The amplitude of the evoked oscillationsof the membrane potential, normalized by the amplitude of stimulat<strong>in</strong>g oscillatory current,is called the neuronal impedance — a frequency-doma<strong>in</strong> extension of the conceptof resistance. The impedance profile of <strong>in</strong>tegrators is decreas<strong>in</strong>g while that of resonatorshas a peak correspond<strong>in</strong>g to the frequency of subthreshold oscillations, around140 Hz <strong>in</strong> the mesV neuron <strong>in</strong> the figure. Thus, <strong>in</strong>tegrators act as low-pass filters whileresonators act as band-pass filters to periodic signals.Instead of s<strong>in</strong>usoidal stimulation, consider more biological stimulation with pulses


Excitability 243membrane potential response, mV-40 mV10 Hz0 pA5 mV1 sec200 pA zap current, pA140 Hzspikes cut200 Hzamplitude of the response<strong>in</strong>tegratorsresonatorsresonancefrequency of stimulationFigure 7.17: Response of the mesV neuron to <strong>in</strong>jected zap current sweep<strong>in</strong>g through arange of frequencies. Integrators and resonators have different responses.-42<strong>in</strong>tegrators-42-42-42-42-42potential (mV)potential (mV)potential (mV)potential (mV)potential (mV)potential (mV)-525 ms-5210 ms-5215 ms0 10 20 30time (ms)0 10 20 30time (ms)0 10 20 30time (ms)resonators-525 ms-5210 ms-5215 ms0 10 20 30time (ms)0 10 20 30time (ms)0 10 20 30time (ms)Figure 7.18: Responses of <strong>in</strong>tegrators (top) and resonators (bottom) to <strong>in</strong>put pulseshav<strong>in</strong>g various <strong>in</strong>ter-pulse periods.


244 Excitability1co<strong>in</strong>cidencedetection<strong>in</strong>tegratorresonatorcomb<strong>in</strong>ed PSP amplitude(normalized)resonantfrequency0.30 3 6 9 12 15 18<strong>in</strong>terpulse period, msFigure 7.19: Dependence of comb<strong>in</strong>ed PSP amplitude on the <strong>in</strong>ter-pulse period; seeFig. 7.18.of current simulat<strong>in</strong>g synaptic bombardment. The response of any neuron to <strong>in</strong>putpulses depends on the frequency content of these pulses. In Fig. 7.18 we use tripletswith various <strong>in</strong>ter-pulse periods to illustrate the issue. The pulses may arrive fromthree different presynaptic neurons or from a s<strong>in</strong>gle presynaptic neuron fir<strong>in</strong>g shortbursts.In Fig. 7.18, top, we show that <strong>in</strong>tegrators prefer high-frequency <strong>in</strong>puts. The firstpulse <strong>in</strong> each triplet evokes a post-synaptic potential (PSP) that decays exponentially.The PSP evoked by the second pulse adds to the first one, and so on. The dependenceof the comb<strong>in</strong>ed PSP amplitude on the <strong>in</strong>ter-pulse period is shown <strong>in</strong> Fig. 7.19. Apparently,the <strong>in</strong>tegrator acts as a co<strong>in</strong>cidence detector because it is most sensitive tothe pulses arriv<strong>in</strong>g simultaneously.Resonators can also detect co<strong>in</strong>cidences, as one can see <strong>in</strong> Fig. 7.19. In addition,they can detect resonant <strong>in</strong>puts. Indeed, the first pulse <strong>in</strong> each triplet <strong>in</strong> Fig. 7.18,bottom, evokes a damped oscillation of the membrane potential, which results <strong>in</strong> anoscillation of the fir<strong>in</strong>g probability. The natural period of such an oscillation is around9 ms for the mesencephalic V neuron used <strong>in</strong> the figure. The effect of the second pulsedepends on its tim<strong>in</strong>g relative to the first pulse: If the <strong>in</strong>terval between the pulses isnear the natural period, e.g., 10 ms <strong>in</strong> Fig. 7.18 and Fig. 7.20, the second pulse arrivesdur<strong>in</strong>g the ris<strong>in</strong>g phase of oscillation, and it <strong>in</strong>creases the amplitude of oscillation evenfurther. In this case the effects of the pulses add up. The third pulse <strong>in</strong>creases theamplitude of oscillation even further thereby <strong>in</strong>creas<strong>in</strong>g the probability of an actionpotential, as <strong>in</strong> Fig. 7.20.If the <strong>in</strong>terval between pulses is near half the natural period, e.g., 5 ms <strong>in</strong> Fig. 7.18and Fig. 7.20, the second pulse arrives dur<strong>in</strong>g the fall<strong>in</strong>g phase of oscillation, and itleads to a decrease <strong>in</strong> oscillation amplitude. The spikes effectively cancel each other


Excitability 2455ms 10ms 15msnon-resonant burstresonant burstnon-resonant burstFigure 7.20: Experimental observations of selective response to a resonant (10 ms<strong>in</strong>terspike period) burst <strong>in</strong> mesencephalic V neurons <strong>in</strong> bra<strong>in</strong>stem hav<strong>in</strong>g subthresholdmembrane oscillations with natural period around 9 ms; see also Fig. 7.18. Threeconsecutive voltage traces are shown to demonstrate some variability of the result.Modified from Izhikevich et al. (2003).<strong>in</strong>hibitory <strong>in</strong>put5ms 15ms 10msnon-resonant burstnon-resonant burst resonant burstFigure 7.21: Experimental observations of selective response to <strong>in</strong>hibitory resonantburst <strong>in</strong> mesencephalic V neurons <strong>in</strong> bra<strong>in</strong>stem hav<strong>in</strong>g oscillatory potentials with thenatural period around 9 ms. Modified from Izhikevich et al. (2003).


246 Excitabilityspikeexcitatory pulsesspike<strong>in</strong>hibitory pulsesspikesmallPSPHodgk<strong>in</strong>-Huxley modeln+m+hspikeVresonantnon-resonant21211 22121Figure 7.22: (Left) Projection of trajectories of the Hodgk<strong>in</strong>-Huxley model on a plane.(Right) Phase portrait and typical trajectories dur<strong>in</strong>g resonant and non-resonant responseof the model to excitatory and <strong>in</strong>hibitory doublets of spikes. Modified fromIzhikevich (2000).out <strong>in</strong> this case. Similarly, the spikes cancel each other when the <strong>in</strong>terpulse period is15 ms, which is 60 % greater than the natural period. The same phenomenon occursfor <strong>in</strong>hibitory synapses, as we illustrate <strong>in</strong> Fig. 7.21. Here the second pulse <strong>in</strong>creases(decreases) the amplitude of oscillation if it arrives dur<strong>in</strong>g the fall<strong>in</strong>g (ris<strong>in</strong>g) phase.We study the mechanism of such frequency preference <strong>in</strong> Ex. 4, and present itsgeometrical illustration <strong>in</strong> Fig. 7.22. There, we depict a projection of the phase portraitof the Hodgk<strong>in</strong>-Huxley model hav<strong>in</strong>g a stable focus equilibrium. The model does nothave a true threshold, as we discuss <strong>in</strong> Sect. 7.2.4. To fire a spike, a perturbationmust push the state of the model beyond the shaded figure that is bounded by twotrajectories, one of which corresponds to a small post-synaptic potential (PSP), whilethe other corresponds to a spike.Fig. 7.22, right. depicts responses of the model to pairs of pulses, called doublets.Pulse 1 <strong>in</strong> the excitatory doublet shifts the membrane potential from the equilibriumto the right, thereby <strong>in</strong>itiat<strong>in</strong>g a subthreshold oscillation. The effect of pulse 2 dependson its tim<strong>in</strong>g: If it arrives when the trajectory f<strong>in</strong>ishes one full rotation around theequilibrium, then it pushes the voltage variable even more to te right, beyond theshaded area <strong>in</strong>to the spik<strong>in</strong>g zone, and the neuron fires an action potential. In contrast,if it arrives too soon, the trajectory does not f<strong>in</strong>ish the rotation, and it is still to theleft of the equilibrium. In this case, pulse 2 pushes the state of the model closer tothe equilibrium, thereby cancel<strong>in</strong>g the effect of pulse 1. Similarly, the effect of an<strong>in</strong>hibitory doublet depends on the <strong>in</strong>terspike period between the <strong>in</strong>hibitory pulses. Ifthe <strong>in</strong>terpulse period is near the natural period of damped oscillations, pulse 2 arriveswhen the trajectory f<strong>in</strong>ishes one full rotation, and it adds to pulse 1, thereby fir<strong>in</strong>g theneuron. If it arrives too soon or too late, it cancels the effect of pulse 1.Quite often, the frequency of subthreshold oscillations depends on their amplitudes,e.g., oscillations <strong>in</strong> the Hodgk<strong>in</strong>-Huxley model slow down as they become larger. In this


Excitability 247AResonant for BResonant for C12 ms 18 msPeriod12 msB2 mVCPeriod18 ms2 mV20 mVFigure 7.23: Selective communication via bursts: Neuron A sends bursts of spikes toneurons B and C that have different natural periods (12 ms and 18 ms, respectively;both are simulations of the Hodgk<strong>in</strong>-Huxley model). As a result of chang<strong>in</strong>g the <strong>in</strong>terspikefrequency, neuron A can selectively affect either B or C without chang<strong>in</strong>g theefficacy of synapses. Modified from Izhikevich (2002).case, the most optimal <strong>in</strong>put is a resonant burst with a slowly decreas<strong>in</strong>g (adapt<strong>in</strong>g)<strong>in</strong>terspike frequency. We will see many examples of such bursts <strong>in</strong> Chap. 9.The fact that resonator neurons prefer <strong>in</strong>puts with “resonant” frequencies is not<strong>in</strong>terest<strong>in</strong>g by itself. What makes it <strong>in</strong>terest<strong>in</strong>g is the observation that the same <strong>in</strong>putcan be resonant for one neuron and non-resonant for another, depend<strong>in</strong>g on theirnatural periods. For example, <strong>in</strong> Fig. 7.23 neurons B and C have different periods ofsubthreshold oscillations: 12 and 18 ms, respectively. By send<strong>in</strong>g a burst of spikeswith <strong>in</strong>terspike <strong>in</strong>terval of 12 ms, neuron A can elicit a response <strong>in</strong> neuron B, butnot <strong>in</strong> C. Similarly, the burst with <strong>in</strong>terspike <strong>in</strong>terval of 18 ms elicits a response <strong>in</strong>neuron C, but not <strong>in</strong> B. Thus, neuron A can selectively affect either neuron B or Cby merely chang<strong>in</strong>g the <strong>in</strong>tra-burst frequency without chang<strong>in</strong>g the efficacy of synapticconnections. In contrast, <strong>in</strong>tegrators do not have this property.7.2.3 Frequency preference <strong>in</strong> vivoFigure 7.20 and 7.21 demonstrate conv<strong>in</strong>c<strong>in</strong>gly the essence of frequency preference andresonance phenomenon <strong>in</strong> vitro, i.e., when the neuron is quiescent and “wait<strong>in</strong>g” forthe resonant burst to come. What if the neuron is under a constant bombardment ofsynaptic <strong>in</strong>put, as it happens <strong>in</strong> vivo, fir<strong>in</strong>g 10 or so spikes per second; Would it beable to tell the difference between the resonant and non-resonant <strong>in</strong>puts?To address this question, we performed a frozen-noise experiment pioneered byBryant and Segundo (1976) and depicted <strong>in</strong> Fig. 7.24. We generated a noisy signal


248 Excitability(g)9 ms(f)(e)(d)(c)(b)(a)membranepotential (mV) spike raster<strong>in</strong>jectedcurrent (pA)7 ms6 ms4 msno burst<strong>in</strong>put200-20-40-602000frozen noiseburst <strong>in</strong>put2 mV6.7 ms50 100 150 200 250 300 350 400 450 500time (ms)Figure 7.24: Frozen-noise experiments demonstrate frequency preference and resonanceto embedded bursts. (a) A random signal (frozen noise) is <strong>in</strong>jected <strong>in</strong>to a neuron <strong>in</strong>vitro to simulate the <strong>in</strong> vivo conditions (b). The neuron responds with some spiketim<strong>in</strong>gvariability depicted <strong>in</strong> (c). (d-g) Burst <strong>in</strong>put is added to the frozen noise. Noticethat the neuron is most sensitive to the <strong>in</strong>put hav<strong>in</strong>g the resonant period 7 ms, whichis near the period of subthreshold oscillation (6.7 ms). Shown are <strong>in</strong> vitro responses ofmesencephalic V neuron of rats bra<strong>in</strong>stem recorded by the author, Niraj S. Desai, andBetsy C. Walcott. The order of stimulation was first l<strong>in</strong>e of c,d,e,f,g, then second l<strong>in</strong>eof c,d,e,f,g, then third l<strong>in</strong>e, etc., to avoid slow artifacts.(frozen noise <strong>in</strong> Fig. 7.24a) and saved it <strong>in</strong>to the memory of the program that <strong>in</strong>jectscurrent <strong>in</strong>to a neuron. Then we <strong>in</strong>jected the stored signal <strong>in</strong>to the neuron 50 times tosee how reliable its spike response is. Despite the <strong>in</strong> vivo-like activity <strong>in</strong> Fig. 7.24b,the spike raster <strong>in</strong> Fig. 7.24c shows vertical clusters <strong>in</strong>dicat<strong>in</strong>g that the neuron prefersto fire at certa<strong>in</strong> “scheduled” moments of time correspond<strong>in</strong>g to certa<strong>in</strong> features of thefrozen-noise <strong>in</strong>put.In Fig. 7.24d-g, we added bursts of 3 spikes to the frozen noise. The amplitudesof the bursts were constant (less than 10% of the frozen noise amplitude), but the<strong>in</strong>terspike periods were different. The idea is to see whether the response of the neuronwould be any different when the burst period is near the neuronal <strong>in</strong>tr<strong>in</strong>sic period of 6.7ms (see the <strong>in</strong>set <strong>in</strong> Fig. 7.24b). As one expects, the non-resonant bursts with 4 ms and9 ms periods rema<strong>in</strong>ed undetected by the neuron, s<strong>in</strong>ce the spike rasters <strong>in</strong> Fig. 7.24d


Excitability 2491ms10 mVthreshold ?Figure 7.25: F<strong>in</strong>d<strong>in</strong>g threshold <strong>in</strong> Hodgk<strong>in</strong>-Huxleymodel.and g are essentially the same as <strong>in</strong> Fig. 7.24c. The resonant burst with 7 ms period <strong>in</strong>Fig. 7.24f produced the most significant deviation from Fig. 7.24c marked by the blackarrow, <strong>in</strong>dicat<strong>in</strong>g that the neuron is most sensitive to the resonant <strong>in</strong>put. Typically,the resonant burst does not make the neuron fire extra spikes, but only changes thetim<strong>in</strong>g of “scheduled” spikes. Inject<strong>in</strong>g resonant bursts at different moments results<strong>in</strong> other <strong>in</strong>terest<strong>in</strong>g phenomena, such as extra spikes or the omission of “scheduled”spikes, not shown here, or no effect at all. F<strong>in</strong>ally, there is a subtle but noticeableeffect of the resonant (7 ms) and nearly-resonant (6 ms) bursts even 100 ms after thestimulation (white arrows <strong>in</strong> the figure), for which we have no explanation.7.2.4 Thresholds and action potentialsA common misconception is that all neurons have fir<strong>in</strong>g thresholds. Moreover, greateffort has been made to determ<strong>in</strong>e such thresholds experimentally. Typically, a neuronis stimulated with brief current pulses of various amplitudes to elicit various degreesof depolarization of the membrane potential, as we illustrate <strong>in</strong> Fig. 7.25 us<strong>in</strong>g theHodgk<strong>in</strong>-Huxley model. Small “subthreshold” depolarizations decay while large “superthreshold”or “suprathreshold” depolarizations result <strong>in</strong> action potentials. Themaximal value of the subthreshold depolarization is taken to be the fir<strong>in</strong>g thresholdvalue for that neuron. Indeed, the neuron will fire a spike if depolarized just abovethat value.The notion of a fir<strong>in</strong>g threshold is simple and attractive, especially when we teachneuroscience to undergraduates. Everybody, <strong>in</strong>clud<strong>in</strong>g the author of this book, usesit to describe neuronal properties. Unfortunately, it is wrong. First, the problem is<strong>in</strong> the def<strong>in</strong>ition of an action potential. Are the two dashed curves <strong>in</strong> Fig. 7.26 actionpotentials? What about a curve <strong>in</strong>-between (not shown <strong>in</strong> the figure)? Suppose wedef<strong>in</strong>e an action potential to be any deviation from the rest<strong>in</strong>g potential, say by 20 mV.Is the concept of fir<strong>in</strong>g threshold well-def<strong>in</strong>ed <strong>in</strong> this case? Unfortunately, the answeris still NO.The membrane potential value that separates subthreshold depolarizations fromaction potentials (whatever the def<strong>in</strong>ition of an action potential is) depends on theprior activity of the neuron. For example, if a neuron hav<strong>in</strong>g transient Na + currentjust fired an action potential, the current is partially <strong>in</strong>activated, and a subsequentdepolarization above the fir<strong>in</strong>g threshold might not evoke another action potential.


250 Excitabilitysquid axonmodel0 mV0 mV-50 mV-50 mV5 ms5 msFigure 7.26: Variable-size action potentials <strong>in</strong> squid giant axon and revised Hodgk<strong>in</strong>-Huxley model (Clay 1998) <strong>in</strong> response to brief steps of currents of variable magnitudes(data was k<strong>in</strong>dly provided by John Clay).Conversely, if the neuron was briefly hyperpolarized and then released from hyperpolarization,it could fire a rebound post-<strong>in</strong>hibitory spike, as we discuss later <strong>in</strong> thischapter (see Fig. 7.29). Apparently, releas<strong>in</strong>g from hyperpolarization does not qualifyas a superthreshold stimulation. Why then the neuron fired?7.2.5 Threshold manifoldsThe problem of formulat<strong>in</strong>g a mathematical def<strong>in</strong>ition of fir<strong>in</strong>g thresholds was firsttackled by FitzHugh (1955). Us<strong>in</strong>g geometrical analysis of neural models, he noticedthat thresholds, if exist, are never numbers but manifolds, e.g., curves <strong>in</strong> twodimensionalsystems. We illustrate his concept <strong>in</strong> Fig. 7.27 us<strong>in</strong>g phase plane analysisof the I Na,p +I K -model.Integrators do have well-def<strong>in</strong>ed threshold manifolds. S<strong>in</strong>ce an <strong>in</strong>tegrator neuron isnear a saddle-node bifurcation, whether on or off an <strong>in</strong>variant circle, there is a saddlepo<strong>in</strong>t with its stable manifold; see Fig. 7.27a. This manifold separates two regions ofthe phase space, and for this reason often called a separatrix. Depend<strong>in</strong>g on the prioractivity of the neuron and the size of the <strong>in</strong>put, its state can end up <strong>in</strong> the shadedarea and generate a subthreshold potential, or <strong>in</strong> the white area and generate an actionpotential. An <strong>in</strong>termediate-size <strong>in</strong>put cannot reduce the size of the action potential,but can only delay its occurrence. In the extreme case, a perturbation can put the statevector precisely on the threshold manifold, and the system converges to the saddle, atleast <strong>in</strong> theory. S<strong>in</strong>ce the saddle is unstable, small noise present <strong>in</strong> neurons pushes thestate either to the left or to the right, result<strong>in</strong>g <strong>in</strong> either a long subthreshold potentialor a large-amplitude spike with a long latency, as we discuss <strong>in</strong> Sect. 7.2.9 and show<strong>in</strong> Fig. 7.34. F<strong>in</strong>ally, notice that a neuron has a s<strong>in</strong>gle threshold value of membranepotential only when its threshold manifold is a straight vertical l<strong>in</strong>e.Resonators may or may not have well-def<strong>in</strong>ed threshold manifolds, depend<strong>in</strong>g onthe type of bifurcation. Consider a resonator neuron <strong>in</strong> the bistable regime; that is,sufficiently near a subcritical Andronov-Hopf bifurcation with an unstable limit cycle


Excitability 251K + activationK + activation0.060(a)-70 -60 -50 -40(c)<strong>in</strong>tegratorsmall EPSPsmallEPSPthreshold manifoldresonatorhalf-amplitudespikespikespike0.50.210.80.60.4(b)threshold manifoldsmallEPSP-70 -60 -50 -40(d)resonatorresonatorcanardtrajectoryPspikespikes0.1thresholdset0.20-60 -50 -40 -30membrane potential, mV-70 -60 -50 -40 -30 -20membrane potential, mVFigure 7.27: Threshold manifolds and sets <strong>in</strong> the I Na,p +I K -model. Parameters <strong>in</strong> (a)as <strong>in</strong> Fig. 4.1a, <strong>in</strong> (b), (c), and (d) as <strong>in</strong> Fig. 6.16 with I = 45 (b) and I = 42 (c andd).separat<strong>in</strong>g the rest<strong>in</strong>g and the spik<strong>in</strong>g states, as <strong>in</strong> Fig. 7.27b. Such an unstable cycleacts as a threshold manifold. Any perturbation that leaves the state of the neuron<strong>in</strong>side the attraction doma<strong>in</strong> of the rest<strong>in</strong>g state, which is the shaded region boundedby the unstable cycle, results <strong>in</strong> subthreshold potentials. Any perturbation that pushesthe state of the neuron outside the shaded region results <strong>in</strong> an action potential. In theextreme case, a perturbation may put the state right on the unstable limit cycle. Then,the neuron exhibits unstable ”threshold” oscillations, at least <strong>in</strong> theory. In practice,such oscillations cannot be susta<strong>in</strong>ed because of noise, and they will either subside orresult <strong>in</strong> spikes.The bistable regime near subcritical Andronov-Hopf bifurcation is the only casewhen a resonator can have a well-def<strong>in</strong>ed threshold manifold. In all other cases, <strong>in</strong>clud<strong>in</strong>gthe supercritical Andronov-Hopf bifurcation, resonators do not have well-def<strong>in</strong>edthresholds. We illustrate this <strong>in</strong> Fig. 7.27c. A small deviation from the rest<strong>in</strong>g stateproduces a trajectory correspond<strong>in</strong>g to a “subthreshold” potential. A large deviationproduces a trajectory correspond<strong>in</strong>g to an action potential. We refer to the shadedregion between the two trajectories as a threshold set. It consists of trajectories corre-


252 Excitability(a) <strong>in</strong>tegrator(b) resonator?new rest<strong>in</strong>gstatethresholdspikethresholdsetnew rest<strong>in</strong>gstatePSPspikeABold rest<strong>in</strong>gstate, I=0old rest<strong>in</strong>gstate, I=0Figure 7.28: Integrators have well-def<strong>in</strong>ed rheobase current while resonators may not.spond<strong>in</strong>g to partial-amplitude action potentials, such as those <strong>in</strong> Fig. 7.26. No s<strong>in</strong>glecurve separates small potentials from action potentials, so there is no well-def<strong>in</strong>edthreshold manifold.FitzHugh (1955) noticed that the threshold set can be quite th<strong>in</strong> <strong>in</strong> some models,<strong>in</strong>clud<strong>in</strong>g the Hodgk<strong>in</strong>-Huxley model. In particular, the difference between the trajectoriescorrespond<strong>in</strong>g to small potentials and action potentials can be as small as 0.0001mV, which is smaller than the noisy fluctuations of the membrane potential. Thus,to observe an <strong>in</strong>termediate-amplitude spike <strong>in</strong> such models, one needs to simulate themodels with accuracy beyond the limits of uncerta<strong>in</strong>ty which appear when the physical<strong>in</strong>terpretation of the model is considered. As a result, for any practical purposesuch models exhibit all-or-none behavior with the threshold set look<strong>in</strong>g like a thresholdmanifold. FitzHugh referred to this as be<strong>in</strong>g a quasithreshold phenomenon.Quasi-thresholds are related to the special canard trajectory depicted <strong>in</strong> Fig. 7.27d.The trajectory follows the unstable branch of the cubic nullcl<strong>in</strong>e all the way to the rightknee po<strong>in</strong>t P. The flow near the trajectory is highly unstable; any small perturbationpushes the state of the system to the left or to the right, result<strong>in</strong>g <strong>in</strong> a “subthreshold”or “superthreshold” response. The solutions depicted <strong>in</strong> Fig. 7.26, right, try to followsuch a trajectory. An easy way to compute the trajectory <strong>in</strong> two-dimensional relaxationoscillators is to start with the po<strong>in</strong>t P and <strong>in</strong>tegrate the system backwards (t → −∞).We discuss canard (French duck) solutions <strong>in</strong> detail <strong>in</strong> Sect. 6.3.4.7.2.6 RheobaseNeuronal rheobase, i.e., the m<strong>in</strong>imal amplitude of a current of <strong>in</strong>f<strong>in</strong>ite duration thatmakes the neuron fire, measures the “current threshold” of the neuron. Integrators havea well-def<strong>in</strong>ed rheobase while resonators may not. To see this, consider an <strong>in</strong>tegratorneuron <strong>in</strong> Fig. 7.28a receiv<strong>in</strong>g a current step that changes <strong>in</strong>stantly its phase portrait.In particular, the current moves the equilibrium from the old location correspond<strong>in</strong>g


Excitability 25310 ms10 mV-45 mV0 pA-100 pAFigure 7.29: Rebound spikes <strong>in</strong> responseto a brief hyperpolariz<strong>in</strong>g pulse<strong>in</strong> a bra<strong>in</strong>stem mesV neuron hav<strong>in</strong>g fastsubthreshold oscillations of membranepotential.to I = 0 (white square <strong>in</strong> the figure) to a new location (black circle). Whether theneuron fires or not depends on the location of the old equilibrium relative to the stablemanifold to the saddle, which plays the role of the new threshold. In case A theneuron does not fire, <strong>in</strong> case B it fires even though the rest<strong>in</strong>g state is still stable. Theneuronal rheobase is the amplitude of the current I that puts the threshold exactlyon the location of the old equilibrium. Such a value of I always exists, and it oftencorresponds to the saddle-node bifurcation value. Notice that the rheobase currentresults <strong>in</strong> a spike with <strong>in</strong>f<strong>in</strong>ite latency, at least theoretically.A resonator neuron may not have well-def<strong>in</strong>ed rheobase simply because it may nothave well-def<strong>in</strong>ed threshold. Indeed, the dotted l<strong>in</strong>e <strong>in</strong> Fig. 7.28b may correspond toa subthreshold or superthreshold response depend<strong>in</strong>g on where it is <strong>in</strong> the thresholdset. Stimulat<strong>in</strong>g such a neuron with “rheobase” current produces spikes with f<strong>in</strong>itelatencies but partial amplitudes. A bistable resonator (near subcritical Andronov-Hopfbifurcation) may have a well-def<strong>in</strong>ed rheobase because it has a well-def<strong>in</strong>ed threshold— the small-amplitude unstable limit cycle.7.2.7 Post-<strong>in</strong>hibitory spikeProlonged <strong>in</strong>jection of a hyperpolariz<strong>in</strong>g current and then sudden release from hyperpolarizationcan produce a rebound post-<strong>in</strong>hibitory response <strong>in</strong> many neurons. Thehyperpolariz<strong>in</strong>g current is often called anodal current, release from the hyperpolarizationis called anodal break, so rebound spik<strong>in</strong>g is called anodal break excitation(FitzHugh 1976). Notice that fir<strong>in</strong>g of a neuron follows a sudden <strong>in</strong>crease of <strong>in</strong>jectedcurrent, whether it is a positive step or release from a negative step.Often, post-<strong>in</strong>hibitory responses are caused by the “hyperpolarization-activated” h-current, which slowly builds up and upon term<strong>in</strong>ation of the hyperpolarization drivesthe membrane potential over the threshold manifold (or threshold set). Alternatively,the rebound response can be caused by slow de-<strong>in</strong>activation of Na + or Ca 2+ currents,or slow de-activation of a K + current that is partially activated at rest and preventsfir<strong>in</strong>g. In any case, such a rebound response relies on slow currents and long or strong


254 Excitability(a) <strong>in</strong>tegrator(b) resonatorK + activation gate<strong>in</strong>hibitionexcitation-60 -40 -20membrane potential, mV<strong>in</strong>hibitionexcitation-60 -40 -20membrane potential, mVFigure 7.30: Direction of excitatory and <strong>in</strong>hibitory <strong>in</strong>put <strong>in</strong> <strong>in</strong>tegrators (a) and resonators(b).hyperpolariz<strong>in</strong>g steps, it does not depend on the bifurcation mechanism of excitability,and it can occur <strong>in</strong> <strong>in</strong>tegrators or resonators.Some neurons can exhibit rebound spikes after short and relatively weak hyperpolariz<strong>in</strong>gcurrents, as we illustrate <strong>in</strong> Fig. 7.29. The negative pulse deactivates afast low-threshold resonant current, e.g., K + current, which is partially activated atrest. Upon release from the hyperpolarization, there is a deficit of the outward currentand the net membrane current results <strong>in</strong> rebound depolarization and possibly a spike.Such a response occurs on the fast time scale and it does depend on the bifurcationmechanism of excitability.In Fig. 7.30 we show why <strong>in</strong>tegrators cannot fire rebound spikes to short stimulation,while resonators typically can. A brief excitatory pulse of current depolarizes themembrane and br<strong>in</strong>gs it closer to the threshold manifold, as <strong>in</strong> Fig. 7.30a. Consequently,an <strong>in</strong>hibitory pulse hyperpolarizes the membrane and <strong>in</strong>creases the distance to thethreshold manifold. The dynamics of such a neuron is consistent with our <strong>in</strong>tuitionthat excitation facilitates spik<strong>in</strong>g and <strong>in</strong>hibition prevents it.Contrary to our <strong>in</strong>tuition, <strong>in</strong>hibition can also facilitate spik<strong>in</strong>g <strong>in</strong> resonator neuronsbecause the threshold set may wrap around the rest<strong>in</strong>g state, as <strong>in</strong> Fig. 7.30b. A sufficientlystrong <strong>in</strong>hibitory pulse can push the state of the neuron beyond the thresholdset thereby evok<strong>in</strong>g a rebound action potential. If the <strong>in</strong>hibitory pulse is not strong, itstill can have an excitatory effect, s<strong>in</strong>ce it br<strong>in</strong>gs the state of the system closer to thethreshold set. For example, it can enhance the effect of subsequent excitatory pulses,as we illustrate <strong>in</strong> Fig. 7.31. The excitatory pulse here is subthreshold if applied alone.However, it becomes superthreshold if preceded by an <strong>in</strong>hibitory pulse. The tim<strong>in</strong>g ofpulses is important here, as we discussed <strong>in</strong> Sect. 7.2.2. John R<strong>in</strong>zel suggested to callthis phenomenon a post-<strong>in</strong>hibitory facilitation.


Excitability 2557.2.8 Inhibition-<strong>in</strong>duced spik<strong>in</strong>gIn Fig. 7.32, top, we use the I Na,t -model <strong>in</strong>troduced <strong>in</strong> Chap. 5 to illustrate an <strong>in</strong>terest<strong>in</strong>gproperty of some resonators — <strong>in</strong>hibition-<strong>in</strong>duced spik<strong>in</strong>g. Recall, that the modelconsists of an Ohmic leak current and a transient Na + current with <strong>in</strong>stantaneous activationand relatively slow <strong>in</strong>activation k<strong>in</strong>etics. It can generate action potentials dueto the <strong>in</strong>terplay between the amplify<strong>in</strong>g gate m and the resonant gate h.We widened the activation function h ∞ (V ) so that Na + current is largely <strong>in</strong>activatedat the rest<strong>in</strong>g state; see the <strong>in</strong>set <strong>in</strong> Fig. 7.32. Indeed, h = 0.27 when I = 0. Eventhough such a system is excitable, it cannot fire repetitive action potentials when apositive step of current, e.g., I = 10, is <strong>in</strong>jected. Depolarization produced by the<strong>in</strong>jected current <strong>in</strong>activates Na + current so much that no repetitive spikes are possible.Such a system is Class 3 excitable.Remarkably, <strong>in</strong>jection of a negative step of current, e.g., I = −15 <strong>in</strong> the figure,results <strong>in</strong> a periodic tra<strong>in</strong> of action potentials! How is it possible? Inhibition-<strong>in</strong>ducedspik<strong>in</strong>g or burst<strong>in</strong>g are possible <strong>in</strong> neurons hav<strong>in</strong>g slow h-current or T-current, such asthe thalamo-cortical relay neurons. (We discuss these and other examples <strong>in</strong> the nextchapter.) The I Na,t -model does not have such currents, yet it can fire <strong>in</strong> response to<strong>in</strong>hibition.Figure 7.33 summarizes the ionic mechanism of <strong>in</strong>hibition-<strong>in</strong>duced spik<strong>in</strong>g. Therest<strong>in</strong>g state <strong>in</strong> the model corresponds to the balance of the outward leak current anda partially activated, partially <strong>in</strong>activated <strong>in</strong>ward Na + current. When the membranepotential is hyperpolarized by the negative <strong>in</strong>jected current, two processes take place:Na + current de<strong>in</strong>activates (variable h <strong>in</strong>creases), and deactivates (variable m = m ∞ (V )decreases). S<strong>in</strong>ce m ∞ (V ) is flatter than h ∞ (V ), de<strong>in</strong>activation is stronger than deacti-K + activation gate10 mV1 ms<strong>in</strong>hibitorypulseexcitatorypulse<strong>in</strong>hibitorypulseexcitatorypulse-60 -40 -20membrane potential, mVFigure 7.31: Post-<strong>in</strong>hibitory facilitation: A subthreshold excitatory pulse can becomesuperthreshold if it is preceded by an <strong>in</strong>hibitory pulse.


256 Excitabilitymembrane potential, V (mV)200-20-40Na+ <strong>in</strong>activation gate, h-600.5I=-20h (V)I=10 0.60I=-15-80 mV rest 20 mV0.70 50 100 150-60 -40 -20 0 20 40time (ms)membrane potential, V (mV)00.10.20.30.4I=-10I=-15I=5I=0I=-5I=10h-nullcl<strong>in</strong>e1m (V)V-nullcl<strong>in</strong>esFigure 7.32: Inhibition-<strong>in</strong>duced spik<strong>in</strong>g <strong>in</strong> the I Na,t -model. Parameters are the sameas <strong>in</strong> Fig. 5.6b, except g leak = 1.5 and m ∞ (V ) has k = 27.hyperpolarizationde<strong>in</strong>activationof I Naexcessof I Nanegative<strong>in</strong>jecteddc-currentrest<strong>in</strong>gpotential<strong>in</strong>activationof I Naactivationof I NaspikeFigure 7.33: Mechanism of <strong>in</strong>hibition<strong>in</strong>ducedspik<strong>in</strong>g <strong>in</strong> the I Na,t -model.vation and the Na + conductance, g Na mh, <strong>in</strong>creases. This leads to an imbalance of the<strong>in</strong>ward current and to the generation of the first spike. Dur<strong>in</strong>g the spike, the current<strong>in</strong>activates completely, and the leak and negative <strong>in</strong>jected currents repolarize and thenhyperpolarize the membrane. Dur<strong>in</strong>g the hyperpolarization, clearly seen <strong>in</strong> the figure,Na + current de<strong>in</strong>activates and is ready for the generation of the next spike.To understand the dynamic mechanism of such an <strong>in</strong>hibition-<strong>in</strong>duced spik<strong>in</strong>g, weneed to consider the geometry of the nullcl<strong>in</strong>es of the model, depicted <strong>in</strong> Fig. 7.32,bottom. Notice how the position of the V -nullcl<strong>in</strong>e depends on I. Negative I shifts thenullcl<strong>in</strong>e down and leftward so that the vertex of its left knee, marked by a dot, movesto the left. As a result, the equilibrium of the system, which is the <strong>in</strong>tersection of theV - and h-nullcl<strong>in</strong>es, moves toward the middle branch of the cubic V -nullcl<strong>in</strong>e. WhenI = −2, the equilibrium loses stability via supercritical Andronov-Hopf bifurcation,and the model exhibits periodic activity.Instead of the I Na,t -model, we could have used the I Na + I K -model or any othermodel with a low-threshold resonant gat<strong>in</strong>g variable. The key po<strong>in</strong>t here is not the ionicbasis of the spike-generation mechanism, but its dynamic attribute — the Andronov-Hopf bifurcation. Even the FitzHugh-Nagumo model (4.11, 4.12) can exhibit thisphenomenon; see Ex. 1.


Excitability 25720 mV50 ms-75 mV0 pA20 pAFigure 7.34: Long latencies and threshold cross<strong>in</strong>g of layer 5 neuron recorded <strong>in</strong> vitro<strong>in</strong> rat motor cortex.7.2.9 Spike latencyIn Fig. 7.34 we illustrate an <strong>in</strong>terest<strong>in</strong>g neuronal property - latency-to-first-spike. Abarely superthreshold stimulation evokes action potentials with a significant delay,which could be as large as a second <strong>in</strong> some cortical neurons. Usually, such a delay isattributed to slow charg<strong>in</strong>g of the dendritic tree or to the action of the A-current, whichis a voltage-gated transient K + current with fast activation and slow <strong>in</strong>activation. Thecurrent activates quickly <strong>in</strong> response to a depolarization and prevents the neuron fromimmediate fir<strong>in</strong>g. With time, however, the A-current <strong>in</strong>activates and eventually allowsfir<strong>in</strong>g. (A low-threshold slowly activat<strong>in</strong>g Na + or Ca 2+ current would achieve a similareffect.)In Fig. 7.35 we expla<strong>in</strong> the latency mechanism from dynamical systems po<strong>in</strong>t ofview. Long latencies arise when neurons undergo saddle-node bifurcation depicted<strong>in</strong> Fig. 7.35, left. When a step of current is delivered, the V -nullcl<strong>in</strong>e moves up sothat the saddle and node equilibria that existed when I = 0 coalesce and annihilateeach other. Although there are no equilibria, the vector field rema<strong>in</strong>s small <strong>in</strong> theshaded neighborhood, as if there were still a ghost of the rest<strong>in</strong>g equilibrium there(see Sect. 3.3.5). The voltage variable <strong>in</strong>creases and passes that neighborhood. Aswe discussed <strong>in</strong> Ex. 3 at the end of the previous chapter, the passage time scales as1/ √ I − I b , where I b is the bifurcation po<strong>in</strong>t, see Fig. 6.8. Hence, the spike is generatedwith a significant latency. If the bifurcation is on an <strong>in</strong>variant circle, then the stateof the neuron returns to the shaded neighborhood after each spike result<strong>in</strong>g <strong>in</strong> fir<strong>in</strong>gwith small frequency, characteristic of Class 1 excitability; see Fig. 7.3. In contrast, ifthe saddle-node bifurcation is off an <strong>in</strong>variant circle, then the state does not returnto the neighborhood, and the fir<strong>in</strong>g frequency can be large, as <strong>in</strong> Fig. 7.34 or <strong>in</strong> theneostriatal and basal ganglia neurons reviewed <strong>in</strong> Sect. 8.4.2.We see that the existence of long spike latencies is an <strong>in</strong>nate neuro-computationalproperty of <strong>in</strong>tegrators. It is still not clear how or when the bra<strong>in</strong> is us<strong>in</strong>g it. Two mostplausible hypotheses are 1) Neurons encode the strength of <strong>in</strong>put <strong>in</strong>to spik<strong>in</strong>g latency.


258 Excitability0.1saddle-node bifurcationV-nullcl<strong>in</strong>e when I>I bV-nullcl<strong>in</strong>e when I=0n-nullcl<strong>in</strong>eK + activation gate10.5Andronov-Hopf bifurcationn-nullcl<strong>in</strong>eI=0V-nullcl<strong>in</strong>e when I>I bV-nullcl<strong>in</strong>e when0-70 -60 -50 -40membrane potential, mV-70 -60 -50 -40 -30membrane potential, mVFigure 7.35: Bifurcation mechanism of latency to first spike when the <strong>in</strong>jected dccurrentsteps from I = 0 to I > I b , where I b is a bifurcation value. The shadedcircle denotes the region where vector field is small. Shown are phase portraits of theI Na,p +I K -model.


Excitability 2592) Neuronal responses become less sensitive to noise, s<strong>in</strong>ce only prolong <strong>in</strong>puts cancause spikes.Interest<strong>in</strong>gly, but resonators do not exhibit long latencies even though there is aneighborhood where the vector field is small and even zero, as we show <strong>in</strong> Fig. 7.35,right. When the current pulse is applied, the V -nullcl<strong>in</strong>e moves up and the voltagevariable accelerates. However, it misses the shaded neighborhood, and the neuronfires an action potential practically without any latency. In Ex. 5 we discuss whysome models near Andronov-Hopf bifurcation, <strong>in</strong>clud<strong>in</strong>g the Hodgk<strong>in</strong>-Huxley model<strong>in</strong> Fig. 7.26, seem to exhibit small but noticeable latencies. In Sect. 8.2.7 we showthat latencies could result from slow charg<strong>in</strong>g of the dendritic compartment. In thiscase, <strong>in</strong>tegrators neurons exhibit latency to the first spike, while resonator neurons mayexhibit latency to the second spike (after they fire the first, transient spike).7.2.10 Flipp<strong>in</strong>g from an <strong>in</strong>tegrator to a resonatorOne of the reasons we provided so many examples of neuronal systems <strong>in</strong> Chap. 5is to conv<strong>in</strong>ce the reader that all neuronal models can exhibit both saddle-node andAndronov-Hopf bifurcations, depend<strong>in</strong>g on the parameters describ<strong>in</strong>g the ionic currents.S<strong>in</strong>ce the k<strong>in</strong>etics of ionic currents <strong>in</strong> neurons could change dur<strong>in</strong>g developmentor due to the action of neuromodulators, neurons could switch from be<strong>in</strong>g <strong>in</strong>tegratorsto be<strong>in</strong>g resonators.In Fig. 7.36 we illustrate an <strong>in</strong>terest<strong>in</strong>g case: Mitral cells <strong>in</strong> rat ma<strong>in</strong> olfactory bulbcan exhibit bistability of membrane potential. That is, the potential can be <strong>in</strong> twostates: down-state around −60 mV, and up-state around −50 mV (Heyward et al.2001). A sufficiently strong synaptic <strong>in</strong>put can shift the cell between these states <strong>in</strong>a matter of milliseconds. An amaz<strong>in</strong>g observation is that the down-state is a stablenode and the up-state is a stable focus, as we illustrate at the bottom of the figure andstudy <strong>in</strong> detail <strong>in</strong> Sect. 8.4.5. As a result, mitral cells can be quickly switched frombe<strong>in</strong>g <strong>in</strong>tegrators to be<strong>in</strong>g resonators by synaptic <strong>in</strong>put.Similar phenomenon was observed <strong>in</strong> a cerebella Purk<strong>in</strong>je neuron <strong>in</strong> Fig. 7.37. It actsas an <strong>in</strong>tegrator <strong>in</strong> the down-state, but has fast (> 100 Hz) subthreshold oscillations<strong>in</strong> the up-state, and hence can act as a resonator.Cortical pyramidal neurons can also exhibit up- and down-states, though the statesare not <strong>in</strong>tr<strong>in</strong>sic, but <strong>in</strong>duced by the synaptic activity. S<strong>in</strong>ce the neurons are depolarized<strong>in</strong> the up-state, there is an <strong>in</strong>terest<strong>in</strong>g possibility that fast K + conductancesare partially activated and the fast Na + <strong>in</strong>activation gate is partially <strong>in</strong>activated sothat the neuron exhibits fast subthreshold oscillations and acts as a resonator. Thatis, <strong>in</strong>tegrator neurons could switch to the resonator mode when <strong>in</strong> the up-state. Thispossibility needs to be tested experimentally.


260 Excitability100 ms10 mVoscillations200 ms20 mV-60 mVup-statedown-state<strong>in</strong>tegratorresonatorn-nullcl<strong>in</strong>eV-nullcl<strong>in</strong>eup-state(focus)down-state(node)Figure 7.36: Bistability of the up- and down-state of mitral cells <strong>in</strong> rat ma<strong>in</strong> olfactorybulb. The cells are <strong>in</strong>tegrators <strong>in</strong> the down-state and resonators <strong>in</strong> the up-state. Membranepotential record<strong>in</strong>gs are modified from Heyward et al. (2001). The shaded areadenotes the attraction doma<strong>in</strong> of the up-state.1 mV20 msfast oscillations-60 mV5 mV100 msFigure 7.37: Fast subthreshold oscillations dur<strong>in</strong>g complex spikes of cerebella Purk<strong>in</strong>jeneuron of a gu<strong>in</strong>ea-pig (data was provided by Yonatan Loewenste<strong>in</strong>).


Excitability 261<strong>in</strong>tegrator(saddle-node bifurcation)?resonator(Andronov-Hopf bifurcation)V-nullcl<strong>in</strong>en-nullcl<strong>in</strong>e?Figure 7.38: Is there an <strong>in</strong>termediate mode between <strong>in</strong>tegrators and resonators?7.2.11 Transition between <strong>in</strong>tegrators and resonatorsConsider the I Na +I K -model or any other m<strong>in</strong>imal model from Chap. 5 that can exhibitsaddle-node or Andronov-Hopf bifurcation, depend<strong>in</strong>g on the parameter values. Let usstart with the I Na +I K -model near saddle-node bifurcation, and hence <strong>in</strong> the <strong>in</strong>tegratormode. The <strong>in</strong>tersection of its nullcl<strong>in</strong>es at the left knee is similar to the one <strong>in</strong> Fig. 7.38,left. Now, let us slowly change the parameters toward the values correspond<strong>in</strong>g to theAndronov-Hopf bifurcation with the nullcl<strong>in</strong>es <strong>in</strong>tersect<strong>in</strong>g as <strong>in</strong> Fig. 7.38, right. Atsome po<strong>in</strong>t, the behavior of the model must change from <strong>in</strong>tegrator to a resonatormode. Is the change sudden, or is it gradual?Any qualitative change of the behavior of the system is a bifurcation. Such abifurcation should somehow comb<strong>in</strong>e the saddle-node and the Andronov-Hopf cases;That is, it should have a zero eigenvalue, and a pair of complex-conjugate eigenvalueswith zero real part. S<strong>in</strong>ce the I Na +I K -model is two-dimensional, the only way these twoconditions are satisfied is when the model undergoes the Bogdanov-Takens bifurcationconsidered <strong>in</strong> Sect. 6.3.3. This bifurcation has codimension 2, that is, it can be reliablyobserved when two parameters are changed, <strong>in</strong> our case E leak and the half-voltage, V 1/2 ,of n ∞ (V ).The top of Fig. 7.39 depicts the phase portrait of the I Na + I K -model at theBogdanov-Takens bifurcation. Notice that the nullcl<strong>in</strong>es are tangent near the left knee,but the tangency is degenerate. A small change of the parameter V 1/2 can result either<strong>in</strong> a saddle and a node (middle of the figure) or <strong>in</strong> a focus equilibrium (bottom of thefigure). The neuron acts as an <strong>in</strong>tegrator <strong>in</strong> the former case and as a resonator <strong>in</strong> thelatter case.Due to the proximity to a codimension-2 bifurcation, the behavior of the I Na +I K -model is quite degenerate. That is, it could exhibit features that are normallynot observed. For example, the <strong>in</strong>tegrator can exhibit post-<strong>in</strong>hibitory spik<strong>in</strong>g, as <strong>in</strong>Fig. 7.40. This occurs because the shaded region <strong>in</strong> the figure, bounded by the stablemanifold of the saddle, goes to the left of the rest<strong>in</strong>g state. An <strong>in</strong>hibitory pulse ofcurrent that hyperpolarizes the membrane potential to V < −65 mV and deactivates


262 ExcitabilityBogdanov-Takens bifurcationK + activation gate, n K + activation gate, n K + activation gate, n0.50.40.30.20.100.50.40.30.20.100.60.50.40.30.20.1n-nullcl<strong>in</strong>eV-nullcl<strong>in</strong>e-60 -40 -20 0V-nullcl<strong>in</strong>e-60 -40 -20 0n-nullcl<strong>in</strong>en-nullcl<strong>in</strong>eK + activation gate, n K + activation gate, n K + activation gate, n<strong>in</strong>tegrator (near saddle-node bifurcation)resonator (near Andronov-Hopf bifurcation)V-nullcl<strong>in</strong>e0.050.040.030.020.010-65 -60 -55 -500.050.040.030.020.01threshold0-65 -60 -55 -500.050.040.030.020.010-60 -40 -20 0membrane potential, V (mV)0-65 -60 -55 -50membrane potential, V (mV)Figure 7.39: Bogdanov-Takens bifurcation <strong>in</strong> the I Na +I K -model (4.1, 4.2). Parametersas <strong>in</strong> Fig. 4.1a, except n ∞ (V ) has k = 7 mV and V 1/2 = −31.64 mV, E leak = −79.42and I = 5. Integrator: V 1/2 = −31 mV and I = 4.3. Resonator: V 1/2 = −34 mV andI = 7.


Excitability 263membrane potential, V (mV)0-20-40-60-800.050.040.030.02I=-100.01I=-10 I=4.30 10 20 30 40 50time (ms)K + activation gate, n0V-nullcl<strong>in</strong>eI=4.3-65 -60 -55 -50membrane potential, V (mV)Figure 7.40: Post-<strong>in</strong>hibitory spike of an <strong>in</strong>tegrator neuron near Bogdanov-Takens bifurcation;see Fig. 7.39.n-nullcl<strong>in</strong>eresonator<strong>in</strong>tegratorspikespike1 2focus1node212Figure 7.41: Post-<strong>in</strong>hibitory facilitation — enhancement of subthreshold depolariz<strong>in</strong>gpulse (2) by preced<strong>in</strong>g <strong>in</strong>hibitory pulse (1) — can occur <strong>in</strong> <strong>in</strong>tegrator neurons nearBogdanov-Takens bifurcation.the K + current to n < 0.005 pushes the po<strong>in</strong>t (V, n) to the shaded region, i.e., beyondthe threshold. Upon release from <strong>in</strong>hibition, the <strong>in</strong>tegrator neuron produces a reboundspike and then returns to the rest<strong>in</strong>g state.Integrator neurons can also exhibit frequency preference and resonance, illustrated<strong>in</strong> Fig. 7.41. The post-<strong>in</strong>hibitory facilitation <strong>in</strong> resonator neurons <strong>in</strong> Fig. 7.41a wasdescribed <strong>in</strong> Sect. 7.2.7. It may happen <strong>in</strong> <strong>in</strong>tegrator neurons when the node equilibriumhas nearly equal eigenvalues and nearly parallel eigenvectors, as <strong>in</strong> Fig. 7.41b. Theformer are about to become complex-conjugate result<strong>in</strong>g <strong>in</strong> rotation of the vector fieldaround the equilibrium, and hence <strong>in</strong> the post-<strong>in</strong>hibitory rebound response to the first(<strong>in</strong>hibitory) pulse.Resonator neurons near Bogdanov-Takens bifurcation can fire spikes with noticeablelatencies. This occurs because the V -nullcl<strong>in</strong>e follows the n-nullcl<strong>in</strong>e at the focusequilibrium <strong>in</strong> Fig. 7.39, bottom. Such a proximity creates a “tunnel” with smallvector field that slows down the spik<strong>in</strong>g trajectory. F<strong>in</strong>ally, the neuron can exhibitan oscillation (marked as P <strong>in</strong> Fig. 7.42, bottom) before fir<strong>in</strong>g a spike <strong>in</strong> response to


264 Excitabilitya pulse of current. Of course, these behaviors are difficult to catch experimentally,because the system must be near a codimension-2 bifurcation.200 ms20 mVSSAAAPSAPSFigure 7.42: Proximity to Bogdanov-Takens bifurcation <strong>in</strong> layer 5 pyramidal neuron ofrat primary visual cortex results <strong>in</strong> slow subthreshold oscillation before a spike. Shownare hand-drown phase portrait and <strong>in</strong> vitro record<strong>in</strong>gs obta<strong>in</strong>ed while an automatedprocedure was test<strong>in</strong>g the neuronal rheobase.7.3 Slow ModulationSo far we have considered neuronal models hav<strong>in</strong>g voltage- or Ca 2+ -gated conductancesoperat<strong>in</strong>g on a fast time scale comparable with the duration of a spike. Such conductancesparticipate directly or <strong>in</strong>directly <strong>in</strong> the generation of each spike and subsequentrepolarization of the membrane potential. In addition, neurons have dendritic trees andsome slow conductances and currents that may not be <strong>in</strong>volved <strong>in</strong> the spike-generationmechanism directly , but rather modulate may it. For example, some cortical pyramidalneurons have I h , all thalamocortical neurons have I h and I Ca(T) . Activation and<strong>in</strong>activation k<strong>in</strong>etics of these current is too slow to participate <strong>in</strong> the generation ofup-stroke or down-stroke of a spike, but the currents can modulate spik<strong>in</strong>g pattern,e.g., they can transform it <strong>in</strong>to burst<strong>in</strong>g.To illustrate the phenomenon of slow modulation, we use the I Na,p +I K +I K(M) -


Excitability 265modelI K(M)I Na,p +I K -modelC ˙V{ }} { { }} {= I −g L (V −E L )−g Na m ∞ (V )(V −E Na )−g K n(V −E K ) − g M n M (V −E K )ṅ = (n ∞ (V ) − n)/τ(V )ṅ M = (n ∞,M (V ) − n M )/τ M (V ) (slow K + M-current)(7.1)whose excitable and spik<strong>in</strong>g properties are similar to those of the I Na,p +I K -submodelon a short time scale. However, the long-term behavior of the two models may be quitedifferent. For example, the K + M-current may result <strong>in</strong> frequency adaptation dur<strong>in</strong>g along tra<strong>in</strong> of action potentials. It can change the shape of the I-V relation of the modeland result <strong>in</strong> slow oscillations, post-<strong>in</strong>hibitory spikes, and other resonator propertieseven when the I Na,p +I K -submodel is an <strong>in</strong>tegrator. All these <strong>in</strong>terest<strong>in</strong>g phenomenaare discussed <strong>in</strong> this section.In general, models hav<strong>in</strong>g fast and slow currents, such as (7.1), can be written <strong>in</strong>the fast-slow formẋ = f(x, u) (fast spik<strong>in</strong>g),˙u = µg(x, u) (slow modulation),(7.2)where the vector x ∈ R m describes fast variables responsible for spik<strong>in</strong>g. It <strong>in</strong>cludes themembrane potential V , activation and <strong>in</strong>activation gat<strong>in</strong>g variables for fast currents,etc. The vector u ∈ R k describes relatively slow variables that modulate fast spik<strong>in</strong>g,e.g., the gat<strong>in</strong>g variable of a slow K + current, the <strong>in</strong>tracellular concentration of Ca 2+ions, etc. The small parameter µ represents the ratio of time scales between spik<strong>in</strong>gand its modulation. Such systems often result <strong>in</strong> burst<strong>in</strong>g activity, and we study them<strong>in</strong> detail <strong>in</strong> Chap. 9.7.3.1 Spike-frequency modulationSlow currents can modulate the <strong>in</strong>stantaneous spik<strong>in</strong>g frequency of a long tra<strong>in</strong> ofaction potentials, as we illustrate <strong>in</strong> Fig. 7.43a us<strong>in</strong>g record<strong>in</strong>gs of a layer 5 pyramidalneuron. The neuron generates a tra<strong>in</strong> of spikes with <strong>in</strong>creas<strong>in</strong>g <strong>in</strong>terspike <strong>in</strong>terval (see<strong>in</strong>set <strong>in</strong> the figure) <strong>in</strong> response to a long pulse of <strong>in</strong>jected dc-current. In Fig. 7.43bwe plot the <strong>in</strong>stantaneous <strong>in</strong>terspike <strong>in</strong>tervals T i , i.e., the time <strong>in</strong>tervals between spikesnumber i and i + 1, as a function of the magnitude of <strong>in</strong>jected current I. Notice thatT i (I) < T i+1 (I), mean<strong>in</strong>g that the <strong>in</strong>tervals <strong>in</strong>crease with each spike. The functionT 0 (I) describes the latency of the first spike, and T ∞ (I) describes the steady-state(asymptotic) <strong>in</strong>terspike period. The <strong>in</strong>stantaneous frequencies are def<strong>in</strong>ed as F i (I) =1000/T i (I) (Hz), and they are depicted <strong>in</strong> Fig. 7.43c. S<strong>in</strong>ce the neuron is Class 1excitable, the F-I curves are square-root parabolas (see Sect. 6.1.2). Notice that F 0 (I)is nearly a straight l<strong>in</strong>e, probably reflect<strong>in</strong>g the passive charg<strong>in</strong>g of the dendritic tree.Decrease of the <strong>in</strong>stantaneous spik<strong>in</strong>g frequency, as <strong>in</strong> Fig. 7.43, is referred to asspike-frequency adaptation. This is a prom<strong>in</strong>ent feature of cortical pyramidal neuronsof the regular spik<strong>in</strong>g (RS) type (Connors and Gutnick 1990), as well as many


266 Excitabilitymembrane potential (mV)<strong>in</strong>terspike period, Ti (ms)40200-20-40-60(a)T 0 T 1 T 2 T 3 T 4 T 5I=300 pA0 pA spike number, i-800 50 time (ms) 100 150(b)<strong>in</strong>terspike frequency, Fi=1000/Ti (Hz)150F120010010080F 160rampF 250 T 0 T40F 3,4,5T 3FT 2 20rampT0T 1 0 00 100 200 3000 100 200 300<strong>in</strong>jected current, I (pA)<strong>in</strong>jected current, I (pA)(c)T i402000 5 10 15Figure 7.43: Spike-frequency adaptation <strong>in</strong> layer 5 pyramidal cell (see Fig. 7.3). Rampdata is from Fig. 7.6.other types of neurons discussed <strong>in</strong> the next chapter. In contrast, cortical fast spik<strong>in</strong>g(FS) <strong>in</strong>terneurons (Gibson et al. 1999) exhibit spike-frequency acceleration, depicted<strong>in</strong> Fig. 7.44, i.e., the <strong>in</strong>stantaneous <strong>in</strong>terspike <strong>in</strong>tervals decrease, and the frequency<strong>in</strong>creases with each spike.Whether a neuron exhibits spike frequency adaptation or acceleration depends onthe nature of the slow current or currents and how they affect the spik<strong>in</strong>g limit cycle ofthe fast subsystem. At the first glance, a resonant slow current, e.g., slowly activat<strong>in</strong>gK + or slowly <strong>in</strong>activat<strong>in</strong>g Na + current, builds up dur<strong>in</strong>g each spike and provides anegative feedback that should slow down spik<strong>in</strong>g of the fast subsystem. Buildup ofa slow amplify<strong>in</strong>g current, e.g., slowly activat<strong>in</strong>g Na + or <strong>in</strong>activat<strong>in</strong>g K + current, orslow charg<strong>in</strong>g of the dendritic tree should have the opposite effect. In Chap. 9, devotedto burst<strong>in</strong>g, we will show that this simple rule works for many models, but there arealso many exceptions. To understand how the slow subsystem modulates repetitivespik<strong>in</strong>g, we need to consider bifurcations of the fast subsystem <strong>in</strong> (7.2), treat<strong>in</strong>g the


Excitability 267membrane potential (mV)0-20-40-60-800-20-40-60-800-20-40-60-800 50 100 150 200time (ms)<strong>in</strong>terspike <strong>in</strong>tervals (ms)1510501510501510500 10 20 30spike numberFigure 7.44: Spike-frequency acceleration of a cortical fast spik<strong>in</strong>g (FS) <strong>in</strong>terneuron.Data k<strong>in</strong>dly provided by Barry Connors.slow variable u as a bifurcation parameter.7.3.2 I-V relationSlow currents and conductances, though not responsible for the generation of spikes,can mask the true I-V relation of the fast subsystem <strong>in</strong> (7.2) responsible for spik<strong>in</strong>g.Take, for example, the I Na,p +I K -model with parameters as <strong>in</strong> Fig. 4.1a (high-thresholdK + current), so that its I-V curve is non-monotonic with a region of negative slopedepicted <strong>in</strong> Fig. 7.45a. Such a system is near saddle-node on <strong>in</strong>variant circle bifurcationand it acts as an <strong>in</strong>tegrator. Now add a slow K + M-current with I-V relation depictedas a dashed curve <strong>in</strong> the figure and a time constant τ M = 100 ms. The spike-generat<strong>in</strong>gmechanism of the comb<strong>in</strong>ed I Na,p +I K +I K(M) -model is described by the fast I Na,p +I K -submodel, so that the neuron cont<strong>in</strong>ue to have <strong>in</strong>tegrator properties, at least on themillisecond time scale. However, the asymptotic I-V relation I ∞ (V ) is dom<strong>in</strong>ated by thestrong I K(M) (V ) and it is monotonic, as if the I Na,p +I K +I K(M) -model is a resonator. Themodel can <strong>in</strong>deed exhibit some resonance properties, such as post-<strong>in</strong>hibitory (rebound)responses, but only on the long time scale of hundreds of milliseconds, i.e., on the timescale of the slow K + M-current.Similarly, we can take a resonator model with a monotonic I-V relation and add aslow amplify<strong>in</strong>g current or a gat<strong>in</strong>g variable to get a non-monotonic I ∞ (V ), as if themodel becomes an <strong>in</strong>tegrator. For example, <strong>in</strong> Fig. 7.45b we use the I Na,p +I K -modelwith parameters as <strong>in</strong> Fig. 4.1b (low-threshold K + current) and add an <strong>in</strong>activation


268 Excitability(a)150100I K(M)(b)300200I fast (V)=I L +I Na,p +I Kcurrent, I500-50I (V)current, I1000-100I (V)-100I fast (V)=I L +I Na,p +I K-200-150-100 -50 0membrane potential, V (mV)-300-100 -50 0 50membrane potential, V (mV)Figure 7.45: Slow conductances can mask the true I-V relation of the spike-generat<strong>in</strong>gmechanism. (a) The I Na,p +I K -model with parameters as <strong>in</strong> Fig. 4.1a has a nonmonotonicI-V curve I fast (V ). Addition of the slow K + M-current with parametersas <strong>in</strong> Sect. 2.3.5 and g M = 5 (dashed curve) makes the asymptotic I-V relation, I ∞ (V ),of the full I Na,p +I K +I K(M) -model monotonic. (b) Addition of a slow <strong>in</strong>activation gateto the K + current of the I Na,p +I K -model with parameters as <strong>in</strong> Fig. 4.1b results <strong>in</strong> anon-monotonic asymptotic I-V relation of the full I Na,p +I A -model.gate to the persistent K + current, effectively transform<strong>in</strong>g it <strong>in</strong>to transient A-current.If the <strong>in</strong>activation k<strong>in</strong>etics is sufficiently slow, the I Na,p +I A -model reta<strong>in</strong>s resonatorproperties on the millisecond time scale, i.e., on the time scale of <strong>in</strong>dividual spikes.However, its asymptotic I-V relation, depicted <strong>in</strong> Fig. 7.45b, becomes non-monotonic.Besides spike-frequency acceleration, the model acquires another <strong>in</strong>terest<strong>in</strong>g property— bistability. A s<strong>in</strong>gle spike does not <strong>in</strong>activate I A significantly. A burst of spikescould <strong>in</strong>activate the K + A-current to such a degree that repetitive spik<strong>in</strong>g becomessusta<strong>in</strong>ed. Slow <strong>in</strong>activation of the A-current is believed to facilitate the transitionfrom down- to up-states <strong>in</strong> neocortical and neostriatal neurons.When a neuronal model consists of conductances operat<strong>in</strong>g on drastically differenttime scales, it has multiple I-V relations, one for each time scale. We illustrate thisphenomenon <strong>in</strong> Fig. 7.46 us<strong>in</strong>g the I Na,p +I K +I K(M) -model with activation time constantof 0.01 ms for I Na,p , 1 ms for I K , and 100 ms for I K(M) . The up-stroke of an actionpotential is described only by leak and persistent Na + currents, s<strong>in</strong>ce the K + currentsdo not have enough time to activate dur<strong>in</strong>g such a short event. Dur<strong>in</strong>g the up-stroke,the model can be reduced to a one-dimensional system (see Chap. 3) with <strong>in</strong>stantaneousI-V relation I 0 (V ) = I leak + I Na,p (V ) depicted <strong>in</strong> Fig. 7.46a. The dynamics dur<strong>in</strong>g andimmediately after the action potential is described by the fast I Na,p +I K -subsystemwith its I-V relation I fast (V ) = I 0 (V ) + I K (V ). F<strong>in</strong>ally, the asymptotic I-V relation,I ∞ (V ) = I fast (V ) + I K(M) (V ), takes <strong>in</strong>to account all currents <strong>in</strong> the model.The three I-V relations determ<strong>in</strong>e the fast, medium, and asymptotic behavior ofa neuron <strong>in</strong> a voltage-clamp experiment. If the time scales are well separated (theyare <strong>in</strong> Fig. 7.46), all three I-V relations can be measured from a simple voltage-clamp


Excitability 269(a)current, I3002001000-100-200I (V)I fast (V)I 0 (V)-300-100 -50 0 50membrane potential, V (mV)(b)current, I2000150010005000-500current response to voltage-clamp stepsI 0 (V)I fast (V)I (V)-2 -1 0 1 2 3time, (logarithmic scale, log 10 ms)(c)current, I10005000<strong>in</strong>stantaneous I-V (d) fast I-V relation (e) steady-state I-V15002000I 0 (V)1000current, I5000I fast (V)current, I150010005000I (V)-5000 0.1 0.2time (ms)-5000 10 20time (ms)-5000 500 1000time (ms)Figure 7.46: (a) The I Na,p +I K +I K(M) -model <strong>in</strong> Fig. 7.45a has three I-V relations: InstantaneousI 0 (V ) = I leak (V ) + I Na,p (V ) describes spike upstroke dynamics. The curveI fast (V ) = I 0 (V ) + I K (V ) is the I-V relation of the fast I Na,p +I K -subsystem responsiblefor spike-generat<strong>in</strong>g mechanism. The curve I ∞ (V ) = I fast (V ) + I K(M) (V ) is thesteady-state (asymptotic) I-V relation of the full model. Dots denote values obta<strong>in</strong>edfrom a simulated voltage-clamp experiment <strong>in</strong> (b); notice the logarithmic time scale.Magnifications of current responses are shown <strong>in</strong> (c,d,e). Simulated time constants:τ Na,p (V ) = 0.01 ms, τ K (V ) = 1 ms, τ M (V ) = 100 ms.experiment, depicted <strong>in</strong> Fig. 7.46b. We hold the model at V = −70 mV and stepthe command voltage to various values. The values of the current, taken at t = 0.05ms, t = 5 ms, and t = 500 ms <strong>in</strong> Fig. 7.46b, result <strong>in</strong> the <strong>in</strong>stantaneous, fast, andsteady-state I-V curves, respectively. Notice that the data <strong>in</strong> Fig. 7.46b are plottedon the logarithmic time scale. Various magnifications us<strong>in</strong>g the l<strong>in</strong>ear time scale aredepicted <strong>in</strong> Fig. 7.46c,d, and e. Numerically obta<strong>in</strong>ed values of the three I-V relationsare depicted as dots <strong>in</strong> Fig. 7.46a. They approximate the theoretical values quite wellbecause there is a 100-fold separation of time scales <strong>in</strong> the model.


270 Excitability50 mV300 msFigure 7.47: Slow subthreshold oscillation of membrane potential of cat thalamocorticalneuron evoked by slow hyperpolarization (modified from Roy et al. 1984).7.3.3 Slow Subthreshold oscillationInteractions between fast and slow conductances can result <strong>in</strong> low-frequency subthresholdoscillation of membrane potential, such as the one <strong>in</strong> Fig. 7.47, even when the fastsubsystem is near a saddle-node bifurcation, acts as an <strong>in</strong>tegrator, and cannot havesubthreshold oscillations. The oscillation <strong>in</strong> Fig. 7.47 is caused by the <strong>in</strong>terplay betweenactivation and <strong>in</strong>activation of the slow Ca 2+ T-current and <strong>in</strong>ward h-current,and it is a precursor of burst<strong>in</strong>g activity, which we consider <strong>in</strong> detail <strong>in</strong> Chap. 9.There exist three different mechanisms of slow subthreshold oscillations of membranepotential of a neuron.• The fast subsystem responsible for spik<strong>in</strong>g has a small-amplitude subthresholdlimit cycle attractor. The period of the limit cycle may be much larger than thetime scale of the slowest variable of the fast subsystem when the cycle is nearsaddle-node on <strong>in</strong>variant circle, saddle homocl<strong>in</strong>ic orbit bifurcation, or Bogdanov-Takens bifurcation considered <strong>in</strong> Chap. 6. In this case, no slow currents or conductancesmodulat<strong>in</strong>g the fast subsystem are needed. However, such a cycle mustbe near the bifurcation, hence the low-frequency subthreshold oscillation exists<strong>in</strong> a narrow parameter range and it is difficult to be seen experimentally.• The I-V relation of the fast subsystem has N-shape <strong>in</strong> the subthreshold voltagerange, so that there are two stable equilibria correspond<strong>in</strong>g to two rest<strong>in</strong>g states,as e.g., <strong>in</strong> Fig. 7.36. A slow resonant variable switches the fast subsystem betweenthose two states via a hysteresis loop result<strong>in</strong>g <strong>in</strong> a subthreshold slow relaxationoscillation.• If the fast subsystem has a monotonic I-V relation, then a stable subthresholdoscillation can result from the <strong>in</strong>terplay between slow variables.The first cases does not need any slow variables, the second case needs only one slowvariable, whereas the third case needs at least two. Indeed, one slow variable is notenough because the entire system can be reduced to a one-dimensional slow equation.7.3.4 Rebound response and voltage sagA slow resonant current can cause a neuron to fire a rebound spike or a burst <strong>in</strong> responseto a sufficiently long hyperpolariz<strong>in</strong>g current, even when the spike-generat<strong>in</strong>g mecha-


Excitability 271Figure 7.48: Rebound responses to long <strong>in</strong>hibitory pulses <strong>in</strong> (a) pyramidal neuron ofsensorimotor cortex of juvenile rat (modified from Hutcheon et al. 1996) and (b) ratauditory thalamic neurons (modified from Tennigkeit et al. 1997).(a)(b)-50 mV25 mV100 mssubthresholdsuperthreshold100 pA0 pA 0 pAFigure 7.49: Post-<strong>in</strong>hibitory facilitation (a) and post-excitatory depression (b) <strong>in</strong> alayer 5 pyramidal neuron (IB type) of rat visual cortex recorded <strong>in</strong> vitro <strong>in</strong> responseto a long hyperpolariz<strong>in</strong>g pulse.nism of the neuron is near a saddle-node bifurcation and hence has neuro-computationalproperties of an <strong>in</strong>tegrator. For example, the cortical pyramidal neuron <strong>in</strong> Fig. 7.48ahas a slow resonant current I h , which opens by hyperpolarization. A short pulse ofcurrent does not open enough of I h and results only <strong>in</strong> a small subthreshold reboundpotential. In contrast, a long pulse of current opens enough I h , result<strong>in</strong>g <strong>in</strong> a strong<strong>in</strong>ward current that produces the voltage sag and, upon term<strong>in</strong>ation of stimulation,drives the membrane potential over the threshold.Similarly, the thalamocortical neuron <strong>in</strong> Fig. 7.48b has a low-threshold Ca 2+ T-current I Ca(T) that is partially activated but completely <strong>in</strong>activated at rest. A negativepulse of current hyperpolarizes the neuron and de<strong>in</strong>activates the T-current, therebymak<strong>in</strong>g it available to generate a spike. Notice that there is no voltage sag <strong>in</strong> Fig. 7.48bbecause the T-current is deactivated at low membrane potentials. Upon term<strong>in</strong>ation ofthe long pulse of current, the membrane potential returns to the rest<strong>in</strong>g state around−68 mV, the Ca 2+ T-current activates and drives the neuron over the threshold. Adist<strong>in</strong>ctive feature of thalamocortical neurons is that they fire a rebound burst of spikes<strong>in</strong> response to strong negative currents.Even when the rebound depolarization is not strong enough to elicit a spike, it


272 Excitability20 mVADP100 msAHP-60 mVFigure 7.50: Afterhyperpolarizations (AHP) and afterdepolarizations (ADP) <strong>in</strong> <strong>in</strong>tr<strong>in</strong>sicallyburst<strong>in</strong>g (IB) pyramidal neurons of the rat motor cortex recorded <strong>in</strong> vitro.Figure 7.51: Rebound spikes and afterdepolarization(marked ADP) at the break of hyperpolariz<strong>in</strong>g current<strong>in</strong> thalamocortical neurons of the cat dorsal lateralgeniculate nucleus. (Data modified from Pirchioet al. 1997, rest<strong>in</strong>g potential is −56 mV, hold<strong>in</strong>g potentialis −67 mV).may <strong>in</strong>crease the excitability of the neuron, so that it fires a spike to an otherwisesubthreshold stimulus, as <strong>in</strong> Fig. 7.49a. This type of post-<strong>in</strong>hibitory facilitation relieson the slow currents, and not on the resonant properties of the spike-generation mechanism(as <strong>in</strong> Fig. 7.31). Figure 7.49b demonstrates the <strong>in</strong>verse property, post-excitatorydepression, i.e., a decreased excitability after a transient depolarization. In this seem<strong>in</strong>glycounter<strong>in</strong>tuitive case, a superthreshold stimulation becomes subthreshold when itis preceded by a depolarized pulse, because the pulse partially <strong>in</strong>activates Na + currentand/or activates K + current.7.3.5 AHP and ADPThe membrane potential may undergo negative and positive deflections right afterthe spike, as illustrated <strong>in</strong> Fig. 7.50 and Fig. 7.51. These are known as afterhyperpolarizations(AHP) and afterdepolarizations (ADP). The latter are sometimes calleddepolariz<strong>in</strong>g after-potentials (DAP). A great effort is usually made to determ<strong>in</strong>e theionic basis of AHPs and ADPs, s<strong>in</strong>ce it is often implicitly assumed that they are generatedby slow currents that turn on right after the spike, such as slow Ca 2+ -gatedK + current I AHP or slow persistent Na + current, respectively. Below we discuss theseand other alternative mechanisms.Let us consider the AHP first. Each spike <strong>in</strong> the <strong>in</strong>itial burst <strong>in</strong> Fig. 7.50 presumably


Excitability 273(a)(b)(c)spikes cut at 0 mVdendriticspikeADP20 mV25 ms-40 mV0 pAADP10 mV10 msADP120 pA60 pA10 mV10 ms-40 mVADPsFigure 7.52: (a) Somatic spike evokes dendritic spike, which <strong>in</strong> turn produces afterdepolarization(ADP) <strong>in</strong> the soma of the pyramidal neuron of rat somatosensory cortex(<strong>in</strong> vitro record<strong>in</strong>g was provided Greg Stuart and Maarten Kole). (b) and (c) Increasedlevel of depolarization <strong>in</strong> another neuron (the same as <strong>in</strong> Fig. 7.49) converts ADP to asecond spike.activates a slow voltage- or Ca 2+ -dependent outward K + current, which eventuallystops the burst and hyperpolarizes the membrane potential. Dur<strong>in</strong>g the AHP period,the slow outward current deactivates and the neuron can fire aga<strong>in</strong>. The neuron canswitch from burst<strong>in</strong>g to tonic spik<strong>in</strong>g mode due to the <strong>in</strong>complete deactivation of theslow current. The same explanation holds if we substitute “activation of outward” by“<strong>in</strong>activation of <strong>in</strong>ward” current.Similarly, slow <strong>in</strong>activation of the transient Ca 2+ T-current expla<strong>in</strong>s the reboundresponse and the long afterdepolarization (marked ADP) <strong>in</strong> Fig. 7.51: The currentwas de<strong>in</strong>activated by the preced<strong>in</strong>g hyperpolarization, so upon release from the hyperpolarization,it quickly activates and slowly <strong>in</strong>activates, thereby produc<strong>in</strong>g a slowdepolariz<strong>in</strong>g wave on which fast spikes can ride. The ADP seen <strong>in</strong> the figure is the tailof the wave.Probably the most common mechanism of ADPs is due to the dendritic spikes,at least <strong>in</strong> pyramidal neurons of neocortex and hippocampus considered <strong>in</strong> the nextchapter. In Fig. 7.52a we depict dual somatic/dendritic record<strong>in</strong>g of membrane potentialof a pyramidal neuron. The somatic spike backpropagates <strong>in</strong>to the dendritictree, activates voltage-gates conductances there, and results <strong>in</strong> a slower dendritic spike.The latter depolarizes the soma and produces a noticeable ADP. Record<strong>in</strong>gs of anotherneuron <strong>in</strong> Fig. 7.52b and c show that if there is an additional source of depolarization,such as the <strong>in</strong>jected dc-current, the ADPs can grow and result <strong>in</strong> a second spike. Thismay evoke another dendritic spike, another ADP or spike, etc., result<strong>in</strong>g <strong>in</strong> a burst<strong>in</strong>gactivity discussed <strong>in</strong> Sect. 8.2.2.Slow ADPs can also be generated due to a nonl<strong>in</strong>ear <strong>in</strong>terplay of fast currentsresponsible for spik<strong>in</strong>g, rather than due to slow currents or dendritic spikes. Oneobvious example is the damped oscillation of membrane potential of the I Na,p +I K -model <strong>in</strong> Fig. 7.53 right after the spike, with the trough and the peak correspond<strong>in</strong>g toan AHP and an ADP, respectively. Notice that the duration of the ADP is ten timesthe duration of the spike even though the model does not have any slow currents. Such


274 ExcitabilityK + activation gate, n0.10.05V-nullcl<strong>in</strong>eseparatrixn-nullcl<strong>in</strong>esaddleADP-65 -60 -55 -50 -45membrane potential, V (mV)membrane potential, V (mV)0-20-40-60ADP-800 10 20 30 40time (ms)Figure 7.53: A long afterdepolarization (ADP) <strong>in</strong> the I Na,p +I K -model without any slowcurrents. Parameters as <strong>in</strong> Fig. 6.52.a long-last<strong>in</strong>g effect appears because the trajectory follows the separatrix, comes closeto the saddle po<strong>in</strong>t, and spends some time there before return<strong>in</strong>g to the stable rest<strong>in</strong>gstate.An example <strong>in</strong> Fig. 7.54 shows the membrane potential of a model neuron slowlypass<strong>in</strong>g through a saddle-node on <strong>in</strong>variant circle bifurcation. Because the vector fieldis small at the bifurcation, which takes place around t = 70 ms, the membrane potentialis slowly <strong>in</strong>creas<strong>in</strong>g along the limit cycle and then slowly decreas<strong>in</strong>g along the locus ofstable node equilibria, thereby produc<strong>in</strong>g a slow ADP. In Chap. 9 we will show that suchADPs exist <strong>in</strong> 4 out of 16 basic types of burst<strong>in</strong>g neurons, <strong>in</strong>clud<strong>in</strong>g thalamocorticalrelay neurons and R 15 burst<strong>in</strong>g cells <strong>in</strong> the abdom<strong>in</strong>al ganglion of the mollusk Aplysia.


Excitability 275membrane potential, V (mV)0-10-20-30-40-50-60saddle-node on <strong>in</strong>variant circlebifurcation<strong>in</strong>variant circleADPsaddlenode-70-800 20 40 60 80 100 120 140 160 180 200time (ms)Figure 7.54: Afterdepolarization <strong>in</strong> the I Na,p +I K -model pass<strong>in</strong>g slowly through saddlenodeon <strong>in</strong>variant circle bifurcation, as the magnitude of the <strong>in</strong>jected current rampsdown.Review of Important Concepts• A neuron is excitable because, as a dynamical system, it is near abifurcation from rest<strong>in</strong>g to spik<strong>in</strong>g activity.• The type of bifurcation determ<strong>in</strong>es the neuron’s computational properties,summarized <strong>in</strong> Fig. 7.15.• Saddle-node on <strong>in</strong>variant circle bifurcation results <strong>in</strong> Class 1 excitability:the neuron can fire with arbitrary small frequency andencode the strength of <strong>in</strong>put <strong>in</strong>to the fir<strong>in</strong>g rate.• Saddle-node off <strong>in</strong>variant circle and Andronov-Hopf bifurcations result<strong>in</strong> Class 2 excitability: the neuron can fire only with<strong>in</strong> a certa<strong>in</strong>frequency range.• Neurons near saddle-node bifurcation are <strong>in</strong>tegrators: they preferhigh-frequency excitatory <strong>in</strong>put, have well-def<strong>in</strong>ed thresholds, andfire all-or-none spikes with some latencies.• Neurons near Andronov-Hopf bifurcation are resonators: they haveoscillatory potentials, prefer resonant-frequency <strong>in</strong>put, and can easilyfire post-<strong>in</strong>hibitory spikes.


276 ExcitabilityBibliographical NotesThere is no universally accepted def<strong>in</strong>ition of excitability. Our def<strong>in</strong>ition is consistentwith the one <strong>in</strong>volv<strong>in</strong>g ε-pseudo-orbits (Izhikevich 2000). R. FitzHugh (1955, 1960,1976) pioneered geometrical analyses of phase portraits of neuronal models with theview to understand their neuro-computational properties. It is amaz<strong>in</strong>g that such importantneuro-computational properties as all-or-none action potentials, fir<strong>in</strong>g thresholds,and <strong>in</strong>tegration of EPSPs are still <strong>in</strong>troduced and illustrated us<strong>in</strong>g the Hodgk<strong>in</strong>-Huxley model, which accord<strong>in</strong>g to FitzHugh, cannot have these properties. Throughoutthis chapter we follow Izhikevich (2000) to compare and contrast neuro-computationalproperties of <strong>in</strong>tegrators and resonators.The frozen noise experiment <strong>in</strong> Fig. 7.24 was pioneered by Bryant and Segundo <strong>in</strong>1976, but due to an <strong>in</strong>terest<strong>in</strong>g quirk of history, it is better known at present as theMa<strong>in</strong>en-Sejnowski (1995) experiment (despite the fact that the latter paper refers tothe former). Post-<strong>in</strong>hibitory facilitation was po<strong>in</strong>ted out by Luk and Aihara (2000),Izhikevich (2001). John R<strong>in</strong>zel suggested to call it “post-<strong>in</strong>hibitory exaltation” (<strong>in</strong>a similar va<strong>in</strong>, the phenomenon <strong>in</strong> Fig. 7.49b may be called “post-excitatory hesitation”).Richardson et al. (2003) po<strong>in</strong>ted out that frequency preference and resonanceoccurs without subthreshold oscillations when the system is near the transition froman <strong>in</strong>tegrator to a resonator.The Hodgk<strong>in</strong>’s classification of neuronal excitability can be applied to classify anyrhythmic system, e.g., contractions of uterus dur<strong>in</strong>g labor. Typically, the contractionsstart with low frequency that gradually <strong>in</strong>creases — Class 1 excitability. The author’swife had to be <strong>in</strong>duced pharmacologically to evoke labor contractions, which is a typicalmedical <strong>in</strong>tervention when the baby is overdue. The contraction monitor showed as<strong>in</strong>usoidal signal with constant period, around 2 m<strong>in</strong>utes, but slowly grow<strong>in</strong>g amplitude— Class 2 excitability via supercritical Andronov-Hopf bifurcation! S<strong>in</strong>ce the motherhad an advance degree <strong>in</strong> applied mathematics, the author waited for a 1-m<strong>in</strong>ute periodof quiescence between the contractions and managed to expla<strong>in</strong> to the deliver<strong>in</strong>g motherthe basic relationship between bifurcations and excitability. Five years later, <strong>in</strong>duceddelivery of the author’s second daughter resulted <strong>in</strong> the same supercritical Andronov-Hopf bifurcation. The author rem<strong>in</strong>ded this concept to the mother and expla<strong>in</strong>ed it tothe obstetrician m<strong>in</strong>utes after the delivery.Exercises1. When can the FitzHugh-Nagumo model (4.11, 4.12) exhibit <strong>in</strong>hibition-<strong>in</strong>ducedspik<strong>in</strong>g, such as the one <strong>in</strong> Fig. 7.32?2. (French ducks) Numerically <strong>in</strong>vestigate the quasi-threshold <strong>in</strong> the FitzHugh-Nagumo model (4.11, 4.12). How is it related to the French duck (canard; seeEckhaus 1983) limit cycles discussed <strong>in</strong> Sect. 6.3.4?


Excitability 277Figure 7.55: Ex. 6: Zero frequency fir<strong>in</strong>g near subcritical Andronov-Hopf bifurcation<strong>in</strong> the I Na,p +I K -model with parameters as <strong>in</strong> Fig. 6.16 and a high-threshold slow K +current (g K,slow = 25, τ K,slow = 10 ms, n ∞,slow (V ) has V 1/2 = −10 mV and k = 5 mV.)3. (Noise-<strong>in</strong>duced oscillations) Consider the systemż = (−ε + iω)z + εI(t) , z ∈ C (7.3)which has a stable focus equilibrium z = 0 and is subject to a weak noisy <strong>in</strong>putεI(t). Show that the system exhibits susta<strong>in</strong>ed noisy oscillations with an averageamplitude |I ∗ (ω)|, whereI ∗ 1(ω) = limT →∞ T∫ T0e −iωt I(t) dtis the Fourier coefficient of I(t) correspond<strong>in</strong>g to the frequency ω.4. (Frequency preference) Show that a system exhibit<strong>in</strong>g damped oscillation withfrequency ω is sensitive to an <strong>in</strong>put hav<strong>in</strong>g frequency ω <strong>in</strong> its power spectrum.(H<strong>in</strong>t: use Ex. 3.)5. (Rush and R<strong>in</strong>zel 1996) Use the phase portrait of the reduced Hodgk<strong>in</strong>-Huxleymodel <strong>in</strong> Fig. 5.21 to expla<strong>in</strong> some small but noticeable latencies <strong>in</strong> Fig. 7.26.6. The neuronal model <strong>in</strong> Fig. 7.55 has a high-threshold slow persistent K + current.Its rest<strong>in</strong>g state undergoes a subcritical Andronov-Hopf bifurcation, yet it canfire low-frequency spikes, and hence exhibits Class 1 excitability. Expla<strong>in</strong>. (H<strong>in</strong>t:Show numerically that the model is near a certa<strong>in</strong> codimention-2 bifurcation<strong>in</strong>volv<strong>in</strong>g a homocl<strong>in</strong>ic orbit).


278 Excitability7. Show that the rest<strong>in</strong>g state of a Class 3 excitable conductance-based model isnear an Andronov-Hopf bifurcation if some other variable, not I, is used as abifurcation parameter.


Chapter 8Simple ModelsThe advantage of us<strong>in</strong>g conductance-based models, such as the I Na +I K -model, is thateach variable and parameter has a well-def<strong>in</strong>ed biophysical mean<strong>in</strong>g. In particular, theycould be measured experimentally. The drawback is that the measurement proceduresmay not be accurate, that the parameters are usually measured <strong>in</strong> different neurons,averaged, and then f<strong>in</strong>e-tuned (a fancy word mean<strong>in</strong>g “to make arbitrary choices”).As a result, the model does not have the same behavior as one sees <strong>in</strong> experiments.And even if it “looks” OK, there is no guarantee that the model is accurate from thedynamical systems po<strong>in</strong>t of view, i.e., that it exhibits the same k<strong>in</strong>d of bifurcations asthe type of neuron under consideration.Sometimes we do not need or cannot afford to have a biophysically detailed conductancebasedmodel. Instead, we want a simple model that faithfully reproduces all the neurocomputationalfeatures of the neuron. In this chapter we review salient features ofcortical, thalamic, and other neurons, and we present simple models that capture theessence of their behavior from the dynamical systems po<strong>in</strong>t of view.8.1 Simplest ModelsLet us start with review<strong>in</strong>g the simplest possible models of neurons. As one canguess from their names, the <strong>in</strong>tegrate-and-fire and resonate-and-fire neurons capturethe essence of <strong>in</strong>tegrators and resonators. The models are similar <strong>in</strong> many respects:both are described by l<strong>in</strong>ear differential equations, both have a hard fir<strong>in</strong>g thresholdand a reset, both have a unique stable equilibrium at rest. The only difference is thatthe equilibrium is a node <strong>in</strong> the <strong>in</strong>tegrate-and-fire case, but it is a focus <strong>in</strong> the resonateand-firecase. One can model the former us<strong>in</strong>g only one equation, and the latter us<strong>in</strong>gonly two equations, though multi-dimensional extensions are straightforward. Bothmodels are useful from the analytical po<strong>in</strong>t of view, i.e., to prove theorems.Many scientists, <strong>in</strong>clud<strong>in</strong>g the author of this book, refer to these neural modelsas be<strong>in</strong>g “spik<strong>in</strong>g models”’. The models have a threshold, but they lack any spikegenerationmechanism, i.e., they cannot produce a brief regenerative depolarizationof membrane potential correspond<strong>in</strong>g to the spike up-stroke. Therefore, they are not279


280 Simple Modelsspikemembrane potential, VE threshE leakE KtimeFigure 8.1: Leaky <strong>in</strong>tegrate-and-fire neuron with noisy <strong>in</strong>put. The spike is addedmanually for aesthetic purposes and to fool the reader <strong>in</strong>to believ<strong>in</strong>g that this is aspik<strong>in</strong>g neuron.“spik<strong>in</strong>g models”; the spikes <strong>in</strong> the next two figures, as well as <strong>in</strong> hundreds of scientificpapers devoted to these models, are drawn by hand. The quadratic <strong>in</strong>tegrate-and-firemodel is the simplest truly spik<strong>in</strong>g model.8.1.1 Integrate-and-fireThe leaky <strong>in</strong>tegrate-and-fire model (Lapicque 1907, Ste<strong>in</strong> 1967, Tuckwell 1988) is anidealization of a neuron hav<strong>in</strong>g Ohmic leakage current and a number of voltage-gatedcurrents that are completely de-activated at rest. Subthreshold behavior of such aneuron can be described by the l<strong>in</strong>ear differential equationC ˙V = I −Ohmic leakage{ }} {g leak (V − E leak ) ,where all parameters have the same biophysical mean<strong>in</strong>gs as <strong>in</strong> the previous chapters.When the membrane potential V reaches the threshold value E thresh , the voltagesensitivecurrents <strong>in</strong>stantaneously activate, the neuron is said to fire an action potential,and V is reset to E K , as <strong>in</strong> Fig. 8.1. After appropriate re-scal<strong>in</strong>g, the leaky <strong>in</strong>tegrateand-firemodel can be written <strong>in</strong> the form˙v = b − v , if v = 1, then v ← 0, (8.1)where the rest<strong>in</strong>g state is v = b, the threshold value is v = 1 and the reset value isv = 0. Apparently, the neuron is excitable when b < 1 and fires a periodic spike tra<strong>in</strong>when b > 1 with period T = − ln(1 − 1/b) (verify that).The <strong>in</strong>tegrate-and-fire neuron illustrates a number of important neuro-computationalproperties:• All-or-none spikes. S<strong>in</strong>ce the shape of the spike is not simulated, all spikes areimplicitly assumed to be identical <strong>in</strong> size and duration.


Simple Models 281• Well-def<strong>in</strong>ed threshold. A stereotypical spike is fired as soon as V = E thresh ,leav<strong>in</strong>g no room for any ambiguity (see though Ex. 1).• Relative refractory period. When E K < E leak , then neuron is less excitable rightafter the spike.• Dist<strong>in</strong>ction between excitation and <strong>in</strong>hibition. Excitatory <strong>in</strong>puts (I > 0) br<strong>in</strong>gthe membrane potential closer to the threshold and hence facilitate fir<strong>in</strong>g, while<strong>in</strong>hibitory <strong>in</strong>puts (I < 0) do the opposite.• Class 1 excitability. The neuron can cont<strong>in</strong>uously encode the strength of an <strong>in</strong>put<strong>in</strong>to the frequency of spik<strong>in</strong>g.In summary, the neuron seems to be a good model for an <strong>in</strong>tegrator.However, a closer look reveals that the <strong>in</strong>tegrate-and-fire neuron has flaws. Thetransition from rest<strong>in</strong>g to repetitive spik<strong>in</strong>g occurs neither via saddle-node nor viaAndronov-Hopf bifurcation, but via some other weird type of a bifurcation that canbe observed only <strong>in</strong> piece-wise cont<strong>in</strong>uous systems. As a result, the F-I curve has logarithmicscal<strong>in</strong>g and not the expected square-root scal<strong>in</strong>g of a typical Class 1 excitablesystem (see though Ex. 6.19). The <strong>in</strong>tegrate-and-fire model cannot have spike latencyto a transient <strong>in</strong>put because superthreshold stimuli evoke immediate spikes withoutany delays (compare with Fig. 8.8(I)). In addition, the model has some weird mathematicalproperties, such as non-uniqueness of solutions, as we show <strong>in</strong> Ex. 1. F<strong>in</strong>ally,the <strong>in</strong>tegrate-and-fire model is not a spik<strong>in</strong>g model: Technically, it did not fire a spike<strong>in</strong> Fig. 8.1, it was only “said to fire a spike”, which was added manually afterwards tofool the reader.Despite all these drawbacks, the <strong>in</strong>tegrate-and-fire model is an acceptable sacrificefor a mathematician who wants to prove theorems and derive analytical expressions.However, us<strong>in</strong>g this model might be a waste of time for a computational neuroscientistwho wants to simulate large-scale networks. At the end of this section we presentalternative models that are as computationally efficient as the <strong>in</strong>tegrate-and-fire neuron,yet as biophysically plausible as Hodgk<strong>in</strong>-Huxley-type models.8.1.2 Resonate-and-fireThe resonate-and-fire model is a two-dimensional extension of the <strong>in</strong>tegrate-and-firemodel that <strong>in</strong>corporates an additional low-threshold persistent K + current or h-current,or any other resonant current that is partially activated at rest. Let W denote themagnitude of such a current. In the l<strong>in</strong>ear approximation, the conductance-basedequations describ<strong>in</strong>g neuronal dynamics can be written <strong>in</strong> the formC ˙V = I − g leak (V − E leak ) − W ,Ẇ = (V − V 1/2 )/k − W .known as Young (1937) model (see also eq. 2-1 <strong>in</strong> FitzHugh 1969). Whenever themembrane potential reaches the threshold value, V thresh , the neurons is said to fire a


282 Simple Models1thresholdyz resetx0-1x-1 0 1100 20 40 60 80 100time, msFigure 8.2: Resonate-and-fire model with b = −0.05, ω = 0.25 and z reset = i. The spikewas added manually.spike. Young did not specify what happens after the spike. The resonate-and-firemodel is the Young model with the resett<strong>in</strong>g: if V ≥ V thresh , then V ← V reset andW ← W reset , where V rest and W reset are some parameters.When the rest<strong>in</strong>g state is a stable focus, the model can be recast <strong>in</strong> complex coord<strong>in</strong>atesasż = (b + iω)z + I ,where b + iω ∈ C is the complex eigenvalue of the rest<strong>in</strong>g state, z = x + iy ∈ C is thecomplex-valued variable describ<strong>in</strong>g damped oscillations with frequency ω around therest<strong>in</strong>g state. The real part, x, is a current-like variable. It describes the dynamicsof the resonant current and synaptic currents. The imag<strong>in</strong>ary part, y is a voltage-likevariable. The neuron is said to fire a spike when y reaches the threshold y = 1. Thus,the threshold is a horizontal l<strong>in</strong>e on the complex plane that passes through i ∈ C, as<strong>in</strong> Fig. 8.2, though other choices are also possible. After fir<strong>in</strong>g the spike, the variablez is reset to z reset .The resonate-and-fire model illustrates the most important features of resonators:damped oscillations, frequency preference, post-<strong>in</strong>hibitory (rebound) spikes, and Class2 excitability. It cannot have susta<strong>in</strong>ed subthreshold oscillations of membrane potential.Integrate-and-fire and resonate-and-fire neurons do not contradict, but complementeach other. Both are l<strong>in</strong>ear, and hence are useful when we prove theorems and deriveanalytical expressions. They have the same flaws limit<strong>in</strong>g their applicability, which wediscussed earlier. In contrast, two simple models described below are difficult to treatanalytically, but because of their universality they should be the models of choice whenlarge-scale simulations are concerned.8.1.3 Quadratic <strong>in</strong>tegrate-and-fireSubstitut<strong>in</strong>g −v by +v 2 <strong>in</strong> (8.1) results <strong>in</strong> the quadratic <strong>in</strong>tegrate-and-fire model˙v = b + v 2 , if v = v peak , then v ← v reset , (8.2)


Simple Models 283which we considered <strong>in</strong> Sect. 3.3.8. Here v peak is not a threshold, but the peak (cut off)of a spike, as we expla<strong>in</strong> below. It is useful to use v peak = +∞ <strong>in</strong> analytical studies. Insimulations, the peak value is assumed to be large but f<strong>in</strong>ite, so it can be normalizedto v peak = 1.Notice that ˙v = b + v 2 is a topological normal form for the saddle-node bifurcation.That is, it describes dynamics of any Hodgk<strong>in</strong>-Huxley-type system near that bifurcation,as we discuss <strong>in</strong> Chap. 3 and 6. There we derived the normal form (6.3) for theI Na,p +I K -model and showed that the two systems agree quantitatively <strong>in</strong> a reasonablybroad voltage range. By resett<strong>in</strong>g v to v reset , the quadratic <strong>in</strong>tegrate-and-fire modelcaptures the essence of recurrence when the saddle-node bifurcation is on an <strong>in</strong>variantcircle.When b > 0, the right-hand side of the model is strictly positive, and the neuronfires a periodic tra<strong>in</strong> of action potentials. Indeed, v <strong>in</strong>creases, reaches the peak, resetsto v reset , and then <strong>in</strong>creases aga<strong>in</strong>, as we show <strong>in</strong> Fig. 3.35, top. In Ex. 3 we prove thatthe period of such spik<strong>in</strong>g activity isT = √ 1 (atan v peak√ − atan v )reset√ < √ π ,b b b bso that the frequency scales as √ b, as <strong>in</strong> Class 1 excitable systems.When b < 0, the parabola b + v 2 has two zeroes, ± √ |b|. One corresponds to thestable node equilibrium (rest<strong>in</strong>g state), the other corresponds to the unstable node(threshold state); see Ex. 2. Subthreshold perturbations are those that keep v belowthe unstable node. Superthreshold perturbations are those that push v beyond theunstable node, result<strong>in</strong>g <strong>in</strong> the <strong>in</strong>itiation of an action potential, reach<strong>in</strong>g the peakvalue v peak , and then resett<strong>in</strong>g to v reset . If, <strong>in</strong> addition, v reset > √ |b|, then there is aco-existence of rest<strong>in</strong>g and periodic spik<strong>in</strong>g states, as <strong>in</strong> Fig. 3.35, bottom. The periodof the spik<strong>in</strong>g state is provided <strong>in</strong> Ex. 4. A two-parameter bifurcation diagram of (8.2)is depicted <strong>in</strong> Fig. 8.3.Unlike its l<strong>in</strong>ear predecessor, the quadratic <strong>in</strong>tegrate-and-fire neuron is a genu<strong>in</strong>e<strong>in</strong>tegrator. It exhibits saddle-node bifurcation, it has a soft threshold, and it generatesspikes with latencies, like many mammalian cells do. Besides, the model is canonical<strong>in</strong> the sense that the entire class of neuronal models near saddle-node on <strong>in</strong>variant circlebifurcation can be transformed <strong>in</strong>to this model by a piece-wise cont<strong>in</strong>uous changeof variables (see Sect. 8.1.5 and the Ermentrout-Kopell theorem <strong>in</strong> Hoppensteadt andIzhikevich 1997). In conclusion, the quadratic and not the leaky <strong>in</strong>tegrate-and-fireneuron should be used <strong>in</strong> simulations of large-scale networks of <strong>in</strong>tegrators. A generalizationof this model is discussed next.8.1.4 Simple model of choiceA strik<strong>in</strong>g similarity among many spik<strong>in</strong>g models, discussed <strong>in</strong> Chap. 5, is that theycan be reduced to two-dimensional systems hav<strong>in</strong>g a fast voltage variable and a slower“recovery” variable, which may describe activation of K + current or <strong>in</strong>activation of Na +


284 Simple Models1v reset =|b| 1/2bistabilitysaddle-node bifurcationparameter v reset0saddle homocl<strong>in</strong>ic orbit bifurcationv resetv resetv resetsaddle-node homocl<strong>in</strong>icorbit bifurcationexcitablev resetv resetv resetsaddle-node on <strong>in</strong>variantcircle bifurcationperiodicv reset-1-1 0 1parameter bFigure 8.3: Bifurcation diagram of the quadratic <strong>in</strong>tegrate-and-fire neuron (8.2).current or their comb<strong>in</strong>ation. Typically, the fast variable has an N-shaped nullcl<strong>in</strong>e andthe slower variable has a sigmoid-shaped nullcl<strong>in</strong>e. The rest<strong>in</strong>g state <strong>in</strong> such modelsis the <strong>in</strong>tersection of the nullcl<strong>in</strong>es near the left knee, as we illustrate <strong>in</strong> Fig. 8.4a.There, V and u denote the fast and the slow variables, respectively. In Chap. 7 weshowed that many computational properties of biological neurons could be expla<strong>in</strong>edby consider<strong>in</strong>g dynamics at the left knee.In Sect. 5.2.4 we derive a simple model that captures the subthreshold behavior <strong>in</strong>a small neighborhood of the left knee conf<strong>in</strong>ed to the shaded square <strong>in</strong> Fig. 8.4 andthe <strong>in</strong>itial segment of the up-stroke of an action potential. In many cases, especially<strong>in</strong>volv<strong>in</strong>g large-scale simulations of spik<strong>in</strong>g models, the shape of the action potentialis less important than the subthreshold dynamics lead<strong>in</strong>g to this action potential. So,reta<strong>in</strong><strong>in</strong>g detailed <strong>in</strong>formation about the left knee and its neighborhood and simplify<strong>in</strong>gthe vector field outside the neighborhood is justified.The simple model˙v = I + v 2 − u if v ≥ 1, then (8.3)˙u = a(bv − u) v ← c, u ← u + d. (8.4)has only four dimensionless parameters. Depend<strong>in</strong>g on the values of a and b, it can


Simple Models 2851recovery variable, u0.80.60.40.2u-nullcl<strong>in</strong>eV-nullcl<strong>in</strong>e0.20.10-80 -60 -40 -20 0 20membrane potential, V (mV)a0-70 -60 -50 -40membrane potential, V (mV)bFigure 8.4: Phase portrait (a) and its magnification (b) of a typical neuronal modelhav<strong>in</strong>g voltage variable V and a recovery variable u.<strong>in</strong>tegratorresonator0.0250thresholdmanifoldsmallEPSPu=I+v2u=bvspikerecovery variable, u0.10smallEPSPu=bvu=I+v2spike-0.1thresholdset-0.2 -0.1 0 0.1 0.2 0.3membrane potential, v-0.5 -0.25 0 0.25 0.5 0.75membrane potential, vFigure 8.5: The simple model (8.3, 8.4) can be an <strong>in</strong>tegrator or a resonator. Comparewith Fig. 7.27.


286 Simple Modelsbe an <strong>in</strong>tegrator or a resonator, as we illustrate <strong>in</strong> Fig. 8.5. The parameters c and ddo not affect steady-state subthreshold behavior. Instead, they take <strong>in</strong>to account theaction of high-threshold voltage-gated currents activated dur<strong>in</strong>g the spike and affectonly the after-spike transient behavior. If there are many currents with diverse timescales, then u, a, b, and d are vectors, and the equation (8.3) conta<strong>in</strong>s ∑ u <strong>in</strong>stead ofu.The simple model may be treated as quadratic <strong>in</strong>tegrate-and-fire neuron with adaptation<strong>in</strong> the simplest case b = 0. When b < 0, the model can be treated as quadratic<strong>in</strong>tegrate-and-fire neuron with a passive dendritic compartment (see Ex. 10). Whenb > 0, the connection to the quadratic <strong>in</strong>tegrate-and-fire neuron is lost, and the simplemodel represents a novel class of spik<strong>in</strong>g models.In the rest of this chapter we tune the simple model to reproduce spik<strong>in</strong>g andburst<strong>in</strong>g behavior of many known types of neurons. It is convenient to use it <strong>in</strong> theformC ˙v = k(v − v r )(v − v t ) − u + I if v ≥ v peak , then (8.5)˙u = a{b(v − v r ) − u} v ← c, u ← u + d (8.6)where v is the membrane potential, u is the recovery current, C is the membranecapacitance, v r is the rest<strong>in</strong>g membrane potential, and v t is the <strong>in</strong>stantaneous thresholdpotential. Though it looks like the model has ten parameters, but it is equivalentto (8.3,8.4) and hence it has only four <strong>in</strong>dependent parameters. As we described <strong>in</strong>Sect. 5.2.4, the parameters k and b can be found know<strong>in</strong>g the neuron’s rheobase and<strong>in</strong>put resistance. The sum of all slow currents that modulate the spike-generationmechanism are comb<strong>in</strong>ed <strong>in</strong> the phenomenological variable u with outward currentstaken with the plus sign.The sign of b determ<strong>in</strong>es whether u is an amplify<strong>in</strong>g (b < 0) or a resonant (b > 0)variable. In the latter case, the neuron sags <strong>in</strong> response to hyperpolarized pulses ofcurrent, peaks <strong>in</strong> response to depolarized subthreshold pulses, and produces rebound(post-<strong>in</strong>hibitory) responses. The recovery time constant is a. The spike cut-off value isv peak , and the voltage reset value is c. The parameter d describes the total amount ofoutward m<strong>in</strong>us <strong>in</strong>ward currents activated dur<strong>in</strong>g the spike and affect<strong>in</strong>g the after-spikebehavior. All these parameters can be easily fit to any particular neuron type, as weshow <strong>in</strong> subsequent sections.Implementation and phase portraitThe follow<strong>in</strong>g MATLAB code simulates the model and produces Fig. 8.6a.C=100; vr=-60; vt=-40; k=0.7;a=0.03; b=-2; c=-50; d=100;vpeak=35;T=1000; tau=1;n=round(T/tau);% parameters used for RS% neocortical pyramidal neurons% spike cutoff% time span and step (ms)% number of simulation steps


Simple Models 287membrane potential, v (mV)40200-20-40(a) (b) v peak-600 200 400 600 800 1000time (ms)upstrokev resetAHP20 msdownstroke(reset)recovery variable, u500-50I=0 I=70AHPrest<strong>in</strong>gu-nullcl<strong>in</strong>ev-nullcl<strong>in</strong>eafter-spike resetspik<strong>in</strong>g limit cycle attractor-60-40-200 20 35v resetvpeakmembrane potential, v (mV)v r v t(c)Figure 8.6: (a) Output of the MATLAB code simulat<strong>in</strong>g the simple model (8.5, 8.6).(b) Comparison of the simulated (cont<strong>in</strong>uous curve) and experimental (dashed curve)voltage traces shows two major discrepancies marked by arrows. (c) Phase portrait ofthe model.v=vr*ones(1,n); u=0*v; % <strong>in</strong>itial valuesI=[zeros(1,0.1*n),70*ones(1,0.9*n)];% pulse of <strong>in</strong>put dc-currentfor i=1:n-1% forward Euler methodv(i+1)=v(i)+tau*(k*(v(i)-vr)*(v(i)-vt)-u(i)+I(i))/C;u(i+1)=u(i)+tau*a*(b*(v(i)-vr)-u(i));if v(i+1)>=vpeak% a spike is fired!v(i)=vpeak;% padd<strong>in</strong>g the spike amplitudev(i+1)=c;% membrane voltage resetu(i+1)=u(i+1)+d;% recovery variable updateend;end;plot(tau*(1:n), v);% plot the resultNotice that the spikes were padded to v peak to avoid amplitude jitter associatedwith the f<strong>in</strong>ite simulation time step tau=1 ms. In Fig. 8.6b we magnify the simulatedvoltage trace and compare it with record<strong>in</strong>g of a neocortical pyramidal neuron (dashedcurve). There are two discrepancies, marked by arrows: The pyramidal neuron has(1) sharper spike upstroke and (2) smoother spike downstroke. The first discrepancycan be removed by assum<strong>in</strong>g that the coefficient k of the square polynomial <strong>in</strong> (8.5)


288 Simple Models<strong>in</strong>tegrate-and-fire modelsimple modelspike cutoffspikesare drawnby handspikes aregeneratedthresholdrest<strong>in</strong>gresetthreshold ?rest<strong>in</strong>greset<strong>in</strong>put<strong>in</strong>putFigure 8.7: Voltage reset <strong>in</strong> the <strong>in</strong>tegrate-and-fire model and <strong>in</strong> the simple model.is voltage-dependent, e.g., k = 0.7 for v ≤ v t and k = 7 for v > v t , or by us<strong>in</strong>gthe modification of the simple model presented <strong>in</strong> Ex. 13 and Ex. 17. The seconddiscrepancy results from the <strong>in</strong>stantaneous after-spike resett<strong>in</strong>g, and it is of a lesserimportance because it does not affect the decision whether or when to fire. However,the slope of the downstroke may become important <strong>in</strong> studies of gap-junction-coupledspik<strong>in</strong>g neurons.The phase portrait of the simple model is depicted <strong>in</strong> Fig. 8.6c. Injection of thestep of dc-current I = 70 pA shifts the v-nullcl<strong>in</strong>e (square parabola) up and makesthe rest<strong>in</strong>g state, denoted by a black square, disappear. The trajectory approaches thespik<strong>in</strong>g limit cycle attractor, and when it crosses the cutoff vertical l<strong>in</strong>e v peak = 35 mV,it is reset to the white square, result<strong>in</strong>g <strong>in</strong> periodic spik<strong>in</strong>g behavior. Notice the slowafterhyperpolarization (AHP) follow<strong>in</strong>g the reset that is due to the dynamics of therecovery variable u. Depend<strong>in</strong>g on the parameters, the model can have other types ofphase portraits, spik<strong>in</strong>g, and burst<strong>in</strong>g behavior, as we demonstrate <strong>in</strong> the rest of thechapter.In Fig. 8.7 we illustrate the difference between the <strong>in</strong>tegrate-and-fire neuron and thesimple model. The <strong>in</strong>tegrate-and-fire model is said to fire spikes when the membranepotential reaches a preset threshold value. The potential is reset to a new value, andthe spikes are drawn by hand. In contrast, the simple model generates the up-strokeof the spike due to the <strong>in</strong>tr<strong>in</strong>sic (regenerative) properties of the voltage equation. Thevoltage reset occurs not at the threshold, but at the peak of the spike. In fact, thefir<strong>in</strong>g threshold <strong>in</strong> the simple model is not a parameter, but a property of the bifurcationmechanism of excitability. Depend<strong>in</strong>g on the bifurcation of equilibrium, the model maynot even have a well-def<strong>in</strong>ed threshold, similarly to conductance-based models.When implement<strong>in</strong>g numerically the voltage reset, whether at the threshold or thepeak of the spike, one needs to be aware of the numerical errors, which translate <strong>in</strong>tothe errors of spike tim<strong>in</strong>g. These errors are <strong>in</strong>versely proportional to the slope of thevoltage variable at the reset value. The slope is small <strong>in</strong> the <strong>in</strong>tegrate-and-fire model, so


Simple Models 289clever numerical methods are needed to catch the exact moment of threshold cross<strong>in</strong>g(Hansel et al. 1998). In contrast, the slope is nearly <strong>in</strong>f<strong>in</strong>ite <strong>in</strong> the simple model, sono special methods are needed to identify the peak of the spike.In Fig. 8.8 we used the model to reproduce the 20 of the most fundamental neurocomputationalproperties of biological neurons. Let us check that the model is thesimplest possible system that can exhibit the k<strong>in</strong>d of behavior <strong>in</strong> the figure. Indeed,it has only one non-l<strong>in</strong>ear term, i.e., v 2 . Remov<strong>in</strong>g the term makes the model l<strong>in</strong>earand equivalent to the resonate-and-fire neuron (though it becomes analytically solvable).Remov<strong>in</strong>g the recovery variable u makes the model equivalent to the quadratic<strong>in</strong>tegrate-and-fire neuron with all its limitations, such as <strong>in</strong>ability to burst or to bea resonator. In summary, we found the simplest possible model capable of spik<strong>in</strong>g,burst<strong>in</strong>g, be<strong>in</strong>g an <strong>in</strong>tegrator or a resonator, and it should be the model of choice <strong>in</strong>simulations of large-scale networks of spik<strong>in</strong>g neurons.8.1.5 Canonical modelsIt is quite rare, if ever possible, to know precisely the parameters describ<strong>in</strong>g dynamicsof a neuron (many erroneously th<strong>in</strong>k that the Hodgk<strong>in</strong>-Huxley model of squid axon isan exception). Indeed, even if all ionic channels expressed by the neuron are known,the parameters describ<strong>in</strong>g their k<strong>in</strong>etics are usually obta<strong>in</strong>ed via averag<strong>in</strong>g over manyneurons; there are measurement errors; the parameters change slowly, etc. Thus, we areforced to consider families of neuronal models with free parameters, e.g. the family ofI Na +I K -models. It is more productive from computational neuroscience po<strong>in</strong>t of viewto consider families of neuronal models hav<strong>in</strong>g a common property, e.g., the family ofall <strong>in</strong>tegrators, the family of all resonators, or the family of “fold/homocl<strong>in</strong>ic” burstersconsidered <strong>in</strong> the next chapter. How can we study the behavior of the entire family ofneuronal models if we have no <strong>in</strong>formation about most of its members?The canonical model approach addresses this issue. Briefly, a model is canonicalfor a family if there is a piece-wise cont<strong>in</strong>uous change of variables that transforms anymodel from the family <strong>in</strong>to this one, as we illustrate <strong>in</strong> Fig. 8.10. The change of variablesdoes not have to be <strong>in</strong>vertible, so the canonical model is usually lower-dimensional,simple, and tractable. Yet, it reta<strong>in</strong>s many important features of the family. Forexample, if the canonical model has multiple attractors, then each member of thefamily has multiple attractors. If the canonical model has a periodic solution, theneach member of the family has a periodic (quasi-periodic or chaotic) solution. If thecanonical model can burst, then each member of the family can burst. The advantageof this approach is that we can study universal neuro-computational properties thatare shared by all members of the family s<strong>in</strong>ce all such members can be put <strong>in</strong>to thecanonical form by a change of variables. Moreover, we need not actually present sucha change of variables explicitly, so derivation of canonical models is possible even whenthe family is so broad that most of its members are given implicitly, e.g., the family of“all resonators”.The process of deriv<strong>in</strong>g canonical models is more an art than a science, s<strong>in</strong>ce a


290 Simple Models(A) tonic spik<strong>in</strong>g(B) phasic spik<strong>in</strong>g (C) tonic burst<strong>in</strong>g (D) phasic burst<strong>in</strong>g<strong>in</strong>put dc-current20 ms(E) mixed mode (F) spike frequency (G) Class 1 excitable (H) Class 2 excitableadaptation(I) spike latency (J) subthreshold (K) resonator (L) <strong>in</strong>tegratoroscillations(M) rebound spike (N) rebound burst (O) threshold (P) bistabilityvariability(Q) depolariz<strong>in</strong>g (R) accommodation (S) <strong>in</strong>hibition-<strong>in</strong>duced (T) <strong>in</strong>hibition-<strong>in</strong>ducedafter-potential spik<strong>in</strong>g burst<strong>in</strong>gDAPFigure 8.8: Summary of neuro-computational properties exhibited by the simple model;see Ex. 11. The figure is reproduced with permission from www.izhikevich.com.(electronic version of the figure and reproduction permissions are freely available atwww.izhikevich.com)


Simple Models 291Figure 8.9:neuronfamily of modelsx=f 1 (x) x=f 2 (x) x=f 3 (x) x=f 4 (x)h 1h 2 h 3 h4y=g(y)canonicalmodelFigure 8.10: <strong>Dynamical</strong> system ẏ = g(y) is a canonical model for the family{f 1 , f 2 , f 3 , f 4 } of neural models ẋ = f(x) because each such model can be transformed<strong>in</strong>to the form ẏ = g(y) by the piece-wise cont<strong>in</strong>uous change of variables h i .


292 Simple Modelsgeneral algorithm for do<strong>in</strong>g this is not known. However, much success has been achieved<strong>in</strong> some important cases. The canonical model for a system near an equilibrium is thetopological normal form at the equilibrium (Kuznetsov 1995). Such canonical model islocal, but it can be extended to describe global dynamics. For example, the quadratic<strong>in</strong>tegrate-and-fire model with a fixed v reset < 0 is a global canonical model for all Class1 excitable systems, i.e., systems near saddle-node on <strong>in</strong>variant circle bifurcation. Thesame model with variable v reset is a global canonical model for all systems near saddlenodehomocl<strong>in</strong>ic orbit bifurcation considered <strong>in</strong> Sect. 6.3.6. The phase model ˙ϑ = 1derived <strong>in</strong> Chap. 10 is a global canonical model for the family of nonl<strong>in</strong>ear oscillatorshav<strong>in</strong>g exponentially stable limit cycle attractors. Other examples of canonical modelsfor spik<strong>in</strong>g and burst<strong>in</strong>g can be found <strong>in</strong> subsequent chapters of this book.The vector-field of excitable conductance-based models <strong>in</strong> the subthreshold regionand <strong>in</strong> the region correspond<strong>in</strong>g to the up-stroke of the spike can be converted <strong>in</strong>to thesimple form (8.3, 8.4), possibly with u be<strong>in</strong>g a vector. Therefore, the simple model(8.3, 8.4) is a local canonical model for the spike-generation mechanism and the spikeup-stroke of the Hodgk<strong>in</strong>-Huxley-type neuronal models. It is not a global canonicalmodel because it ignores the spike downstroke. Yet, it describes remarkably well thespik<strong>in</strong>g and burst<strong>in</strong>g dynamics of many biological neurons, as we demonstrate next.8.2 CortexIn this section we consider the six most fundamental classes of fir<strong>in</strong>g patterns observed<strong>in</strong> the mammalian neocortex and depicted <strong>in</strong> Fig. 8.11 (Connors and Gutnick 1990,Gray and McCormick 1996, Gibson et al. 1999). Though most biologists agree withthe classification <strong>in</strong> the figure, many would po<strong>in</strong>t out that it is greatly oversimplified(Markram et al. 2004), that the dist<strong>in</strong>ction between the classes is not sharp, that thereare subclasses with<strong>in</strong> each class (Nowak et al. 2003, Toledo-Rodriguez et al. 2004),and that neurons can change their fir<strong>in</strong>g classes depend<strong>in</strong>g on the state of the bra<strong>in</strong>(Steriade 2004).• (RS) Regular spik<strong>in</strong>g neurons fire tonic spikes with adapt<strong>in</strong>g (decreas<strong>in</strong>g) frequency<strong>in</strong> response to <strong>in</strong>jected pulses of dc-current. Most of them have Class 1excitability <strong>in</strong> the sense that the <strong>in</strong>terspike frequency vanishes when the amplitudeof the <strong>in</strong>jected current decreases. Morphologically, these neurons are sp<strong>in</strong>ystellate cells <strong>in</strong> layer 4 and pyramidal cells <strong>in</strong> layers 2,3,5, and 6.• (IB) Intr<strong>in</strong>sically burst<strong>in</strong>g neurons generate a burst of spikes at the beg<strong>in</strong>n<strong>in</strong>g ofa strong depolariz<strong>in</strong>g pulse of current, and then switch to tonic spik<strong>in</strong>g mode.Those are excitatory pyramidal neurons found <strong>in</strong> all cortical layers, but mostabundant <strong>in</strong> layer 5.• (CH) Chatter<strong>in</strong>g neurons fire high-frequency bursts of spikes with relatively short<strong>in</strong>terburst periods, hence another name — FRB or fast rhythmic burst<strong>in</strong>g. Outputof such a cell fed to the loudspeaker “sounds a lot like a helicopter — cha,


Simple Models 293regular spik<strong>in</strong>g (RS)excitatorychatter<strong>in</strong>g (CH)300 pA600 pA100 pA300 pA35 pA200 pA<strong>in</strong>tr<strong>in</strong>sically burst<strong>in</strong>g (IB)50 mV600 pA100 ms500 pA400 pA<strong>in</strong>hibitoryfast spik<strong>in</strong>g (FS) low threshold spik<strong>in</strong>g (LTS)500 pA 400 pA300 pAlate spik<strong>in</strong>g (LS)200 pA175 pA100 pA150 pAdelayFigure 8.11: Six most fundamental classes of fir<strong>in</strong>g patterns of neocortical neurons<strong>in</strong> response to pulses of depolariz<strong>in</strong>g dc-current. RS and IB are <strong>in</strong> vitro record<strong>in</strong>gsof pyramidal neurons of layer 5 of primary visual cortex of a rat, CH was recorded<strong>in</strong> vivo <strong>in</strong> cat’s visual cortex (area 17, data provided by D. McCormick). FS wasrecorded <strong>in</strong> vitro <strong>in</strong> rat’s primary visual cortex, LTS was recorded <strong>in</strong> vitro <strong>in</strong> layer 4or 6 of rat’s barrel cortex (data provided by B. Connors). LS was recorded <strong>in</strong> layer1 of rat’s visual cortex (data provided by S. Hestr<strong>in</strong>). All record<strong>in</strong>gs are plotted onthe same voltage and time scale, and the data are available on the author’s webpage(www.izhikevich.com).


294 Simple Modelscha, cha — real fast”, accord<strong>in</strong>g to Gray and McCormick (1996). CH neuronswere found <strong>in</strong> visual cortex of adult cats, and, morphologically, they are sp<strong>in</strong>ystellate or pyramidal neurons of layer 2-4, ma<strong>in</strong>ly layer 3.• (FS) Fast spik<strong>in</strong>g <strong>in</strong>terneurons fire high-frequency tonic spikes with relatively constantperiod. They exhibit Class 2 excitability (Tateno et al. 2004). When themagnitude of the <strong>in</strong>jected current decreases below a certa<strong>in</strong> critical value, theyfire irregular spikes, switch<strong>in</strong>g randomly between rest<strong>in</strong>g and spik<strong>in</strong>g states. Morphologically,FS neurons are sparsely sp<strong>in</strong>y or asp<strong>in</strong>y nonpyramidal cells (basketor chandelier; see Kawaguchi and Kubota 1997) provid<strong>in</strong>g local <strong>in</strong>hibition alongthe horizontal (<strong>in</strong>tra-lam<strong>in</strong>ar) direction of the neocortex (Bacci et al. 2003).• (LTS) Low-threshold spik<strong>in</strong>g neurons fire tonic spikes with pronounced spikefrequency adaptation and rebound (post-<strong>in</strong>hibitory) spikes, often called “lowthresholdspikes” by biologists (hence the name). They seem to be able to firelow-frequency spike tra<strong>in</strong>s, though their excitability class has not been determ<strong>in</strong>edyet. Morphologically, LTS neurons are nonpyramidal <strong>in</strong>terneurons provid<strong>in</strong>g local<strong>in</strong>hibition along the vertical (<strong>in</strong>ter-lam<strong>in</strong>ar) direction of the neocortex (Bacci etal. 2003).• (LS) Late spik<strong>in</strong>g neurons exhibit voltage ramp <strong>in</strong> response to <strong>in</strong>jected dc-currentnear the rheobase, result<strong>in</strong>g <strong>in</strong> delayed spik<strong>in</strong>g with latencies as long as 1 sec.There is a pronounced sub-threshold oscillation dur<strong>in</strong>g the ramp, but the dischargefrequency is far less than that of FS neurons. Morphologically, LS neuronsare nonpyramidal <strong>in</strong>terneurons (neurogliaform; see Kawaguchi and Kubota 1997)found <strong>in</strong> all layers of neocortex (Kawaguchi 1995), especially <strong>in</strong> layer 1 (Chu etal. 2003).Our goal is to use the simple model (8.5, 8.6) presented <strong>in</strong> the previous section toreproduce each of the fir<strong>in</strong>g types. We want to capture the dynamic mechanism of spikegeneration of each neuron, so that the model reproduces the correct responses to manytypes of the <strong>in</strong>puts, and not only to the pulses of dc-current. We strive to have notonly qualitative but also quantitative agreement with the published data on neuron’srest<strong>in</strong>g potential, <strong>in</strong>put resistance, rheobase, F-I behavior, the shape of the upstrokeof the action potential, etc., though this is impossible <strong>in</strong> many cases mostly becausethe data are contradictory. To f<strong>in</strong>e-tune the model, we use record<strong>in</strong>gs of real neurons.We consider the tun<strong>in</strong>g successful when quantitative difference between simulated andrecorded responses is smaller than the difference between responses of two “sister”neurons recorded <strong>in</strong> the same slice. We do not want to claim that the simple modelexpla<strong>in</strong>s the mechanism of generation of any of the fir<strong>in</strong>g patterns recorded <strong>in</strong> realneurons (simply because the mechanism is usually not known). Although <strong>in</strong> many<strong>in</strong>stances we must resist the temptation to use the Wolfram (2002) new-k<strong>in</strong>d-of-sciencecriterion: “if it looks the same — it must be the same”.


Simple Models 2958.2.1 Regular spik<strong>in</strong>g (RS) neuronsRegular spik<strong>in</strong>g neurons are the major class of excitatory neurons <strong>in</strong> the neocortex.Many are Class 1 excitable, as we show <strong>in</strong> Fig. 7.3 us<strong>in</strong>g <strong>in</strong> vitro record<strong>in</strong>gs of a layer 5pyramidal cell of rat’s primary visual cortex; see also Tateno et al. (2004). RS neuronshave a transient K + current I A , whose slow <strong>in</strong>activation delays the onset of the firstspike and <strong>in</strong>creases the <strong>in</strong>terspike period, and a persistent K + current I M , which isbelieved to be responsible for the spike frequency adaptation seen <strong>in</strong> Fig. 7.43. Let ususe the simple model (8.5, 8.6) to capture qualitative and some quantitative featuresof typical RS neurons.We assume that the rest<strong>in</strong>g membrane potential is v r = −60 mV and the <strong>in</strong>stantaneousthreshold potential is v t = −40 mV; that is, <strong>in</strong>stantaneous depolarizations above−40 mV cause the neuron to fire, as <strong>in</strong> Fig. 3.15. Assum<strong>in</strong>g that the rheobase is 50pA, and the <strong>in</strong>put resistance is 80 MΩ, we f<strong>in</strong>d k = 0.7 and b = −2. We take themembrane capacitance C = 100 pF, which yields a membrane time constant of 8 ms.S<strong>in</strong>ce b < 0, depolarizations of v decrease u as if the major slow current is the<strong>in</strong>activat<strong>in</strong>g K + current I A . The <strong>in</strong>activation time constant of I A is around 30 ms<strong>in</strong> the subthreshold voltage range, hence we take a = 0.03 ≈ 1/30. The membranepotential of a typical RS neuron reaches the peak value v peak = +35 mV dur<strong>in</strong>g aspike (the precise value has little effect on dynamics) and then repolarizes to c = −50mV or below, depend<strong>in</strong>g on the fir<strong>in</strong>g frequency. The parameter d describes the totalamount of outward m<strong>in</strong>us <strong>in</strong>ward currents activated dur<strong>in</strong>g the spike and affect<strong>in</strong>g theafter-spike behavior. Try<strong>in</strong>g different values, we f<strong>in</strong>d that d = 100 gives a reasonableF-I relationship, at least <strong>in</strong> the low-frequency range.As it follows from Ex. 10, we can also <strong>in</strong>terpret u as the membrane potential ofa passive dendritic compartment, taken with the m<strong>in</strong>us sign. Thus, when b < 0, thevariable u represents the comb<strong>in</strong>ed action of slow <strong>in</strong>activation of I A and slow charg<strong>in</strong>gof the dendritic tree. Both processes slow down the frequency of somatic spik<strong>in</strong>g.Notice that we round-up all the parameters, e.g., we use d = 100 and not 93.27.Nevertheless, the simulated voltage responses <strong>in</strong> Fig. 8.12 agree quantitatively withthe <strong>in</strong> vitro record<strong>in</strong>gs of the layer 5 pyramidal neuron used <strong>in</strong> Fig. 7.3. Tweak<strong>in</strong>gthe parameters, consider<strong>in</strong>g multidimensional u, or add<strong>in</strong>g multiple dendritic compartments,one can def<strong>in</strong>itely improve the quantitative correspondence between the modeland the <strong>in</strong> vitro data of that particular neuron, but this is not our goal here. Instead,we want to understand the qualitative dynamics of RS neurons us<strong>in</strong>g the geometry oftheir phase portraits.Phase plane analysisFigure 8.13 shows record<strong>in</strong>gs of two pyramidal RS neurons from the same slice whilean automated procedure <strong>in</strong>jects pulses of dc-current to determ<strong>in</strong>e their rheobase. Theneuron on the left exhibits monotonically <strong>in</strong>creas<strong>in</strong>g (ramp<strong>in</strong>g) or decreas<strong>in</strong>g responsesof membrane potential to weak <strong>in</strong>put pulses, long latencies of the first spike and norebound spikes, whereas the neuron on the right exhibits non-monotone overshoot<strong>in</strong>g


296 Simple Modelslayer 5 neuronsimple modelI=100 pAI=85 pAI=70 pA<strong>in</strong>put+35 mV200 ms-60 mV<strong>in</strong>putresetAHPrecovery, u100500-50restresetAHPI>0I=0I=60 pA-60 -40 -20 0 20membrane potential, v (mV)spike35Figure 8.12: Comparison of <strong>in</strong> vitro record<strong>in</strong>gs of a regular spik<strong>in</strong>g (RS) pyramidalneurons with simulations of the simple model 100 ˙v = 0.7(v + 60)(v + 40) − u + I,˙u = 0.03{−2(v + 60) − u}, if v ≥ +35, then v ← −50, u ← u + 100.responses to positive pulses, sags and rebound spikes to negative pulses (as <strong>in</strong> Fig. 7.48),relatively short latencies of the first spike, and other resonance phenomena. The evenmore extreme example <strong>in</strong> Fig. 7.42 shows a pyramidal neuron execut<strong>in</strong>g a subthresholdoscillation before switch<strong>in</strong>g to a tonic spik<strong>in</strong>g mode.The difference between the types <strong>in</strong> Fig. 8.13 can be expla<strong>in</strong>ed by the sign of theparameter b <strong>in</strong> the simple model (8.5, 8.6), which depends on the relative contributionsof amplify<strong>in</strong>g and resonant slow currents and gat<strong>in</strong>g variables. When b < 0 (or b ≈ 0,e.g., b = 0.5 <strong>in</strong> Fig. 8.14), the neuron is a pure <strong>in</strong>tegrator near saddle-node on <strong>in</strong>variantcircle bifurcation. Greater values of b > 0 put the model near the transition from an<strong>in</strong>tegrator to a resonator the via co-dimension-2 Bogdanov-Takens bifurcation studied<strong>in</strong> Sect. 6.3.3 and 7.2.11.The sequence of bifurcations when b > 0 is depicted <strong>in</strong> Fig. 8.15. Injection ofdepolariz<strong>in</strong>g current below the neuron’s rheobase transforms the rest<strong>in</strong>g state <strong>in</strong>toa stable focus and results <strong>in</strong> damped oscillations of the membrane potential. The


Simple Models 2975 mV-30 mVS-30 mV100 msPrestrestub0uI>0I>0I=0I=0restPu=bvvu=bvrestvuI>0uI>0I=0I=0Srestu=bvvu=bvrestvFigure 8.13: Two types of qualitative behavior of RS neurons. Some exhibit monotoneresponses to weak <strong>in</strong>jected currents (case b < 0), others exhibit non-monotoneovershoot<strong>in</strong>g responses (case b > 0). Shown are <strong>in</strong> vitro record<strong>in</strong>gs of two RS neuronsfrom the same slice of rat’s primary visual cortex while an automated procedure wastry<strong>in</strong>g to determ<strong>in</strong>e the neurons’ rheobase. Phase portraits are drawn by hand andthey illustrate a possible dynamic mechanism of the phenomenon.


298 Simple Modelsb=-240200saddle-node on<strong>in</strong>variant circle bifurcationI=40 I=52 I=60u-nullcl<strong>in</strong>ev-nullcl<strong>in</strong>e-20spikerecovery variable, u-40b=0.54030saddle-node on<strong>in</strong>variant circle bifurcationI=70 I=75 I=8020100spike-10-20-60 -55 -50 -45 -40-60 -55 -50 -45 -40 -60 -55 -50 -45 -40membrane potential, v (mV)Figure 8.14: Saddle-node on <strong>in</strong>variant circle bifurcations <strong>in</strong> the RS neuron model asthe magnitude of the <strong>in</strong>jected current I <strong>in</strong>creases.10090big saddle homocl<strong>in</strong>icorbit bifurcationI=120 I=124.5 I=1258070spikespik<strong>in</strong>gbig homocl<strong>in</strong>iclimit cycle60recovery variable, u504010090saddle homocl<strong>in</strong>icorbit bifurcationorbitsubcritical Andronov-HopfbifurcationI=126.5 I=127 I=127.58070605040-60 -55 -50 -45 -40-60 -55 -50 -45 -40 -60 -55 -50 -45 -40membrane potential, v (mV)Figure 8.15: The sequence of bifurcations of the RS model neuron (8.5, 8.6) <strong>in</strong> resonatorregime. Parameters as <strong>in</strong> Fig. 8.12 and b = 5; see also Fig. 6.40.


Simple Models 29920 mV100 ms49 pArecovery variablespik<strong>in</strong>g limitcycleBABA-60 mVmembrane potentialFigure 8.16: Stutter<strong>in</strong>g behavior of an RS neuron (data were provided by Dr. Klaus M.Stiefel, P28-36 adult mouse, coronal slices, 300µm, layer II/III pyramid, visual cortex).attraction doma<strong>in</strong> of the focus (shaded region <strong>in</strong> the figure) is bounded by the stablemanifold of the saddle. As I <strong>in</strong>creases, the stable manifold makes a loop and becomesa big homocl<strong>in</strong>ic orbit giv<strong>in</strong>g birth to a spik<strong>in</strong>g limit cycle attractor. When I = 125,stable rest<strong>in</strong>g and spik<strong>in</strong>g states co-exist, which plays an important role <strong>in</strong> expla<strong>in</strong><strong>in</strong>gthe paradoxical stutter<strong>in</strong>g behavior of some neocortical neurons discussed later. AsI <strong>in</strong>creases, the saddle quantity, i.e., the sum of eigenvalues of the saddle, becomespositive. When the stable manifold makes another, smaller loop, it gives birth toan unstable limit cycle, which then shr<strong>in</strong>ks to the rest<strong>in</strong>g equilibrium and results <strong>in</strong>subcritical Andnronov-Hopf bifurcation.What is the excitability class of the RS model neuron <strong>in</strong> Fig. 8.15? If a slow ramp ofcurrent is <strong>in</strong>jected, the rest<strong>in</strong>g state of the neuron becomes a stable focus and then losesstability via subcritical Andronov-Hopf bifurcation. Hence the neuron is a resonatorexhibit<strong>in</strong>g Class 2 excitability. Now suppose steps of dc-current of amplitude I = 125pA or less are <strong>in</strong>jected. The trajectory starts at the <strong>in</strong>itial po<strong>in</strong>t (v, u) = (−60, 0),which is the rest<strong>in</strong>g state when I = 0, and then approaches the spik<strong>in</strong>g limit cycle.Because the limit cycle was born via a homocl<strong>in</strong>ic bifurcation to the saddle, it has alarge period and hence the neuron is Class 1 excitable. Thus, depend<strong>in</strong>g on the natureof stimulation, i.e., ramps vs. pulses, we can observe small or large spik<strong>in</strong>g frequencies,at least <strong>in</strong> pr<strong>in</strong>ciple.In practice, it is quite difficult to catch homocl<strong>in</strong>ic orbits to saddles because they aresensitive to noise. Injection of a constant current just below the neuron’s rheobase <strong>in</strong>Fig. 8.15 would result <strong>in</strong> random transitions between the rest<strong>in</strong>g state and a periodicspik<strong>in</strong>g state. Indeed, the two attractors co-exist and are near each other, so weakmembrane noise can push the trajectory <strong>in</strong> and out of the shaded region result<strong>in</strong>g <strong>in</strong>a stutter<strong>in</strong>g spik<strong>in</strong>g, illustrated <strong>in</strong> Fig. 8.16, m<strong>in</strong>gled with subthreshold oscillations.Such a behavior is also exhibited by FS <strong>in</strong>terneurons studied later <strong>in</strong> this section.


300 Simple Models(a)simple modellayer 5 neuron (<strong>in</strong> vitro)20 mV20 ms(b)32 nS8 nSg GABAg AMPAresponse<strong>in</strong>put(c)1 seclayer 5 neuron(<strong>in</strong> vitro)simple modelFigure 8.17: (a) Comparison of responses of a rat motor cortex layer 5 pyramidal neuronof RS type and the simple model (8.5, 8.6) to <strong>in</strong> vivo-like stochastic <strong>in</strong>put (8.7) withthe random conductances <strong>in</strong> (b). Part (a) is a magnification of a small region <strong>in</strong> (c).Shown are simulations of 30 ˙v = 3(v+55)(v+42)−u+I(t), ˙u = 0.01{−0.25(v+55)−u},if v ≥ +10, then v ← −40, u ← u + 90. Data was k<strong>in</strong>dly provided by Niraj S. Desaiand Betsy C. Walcott.In vivo-like conditionsIn Fig. 8.17a (dashed curve) we show the response of <strong>in</strong> vitro recorded layer 5 pyramidalneuron of rat motor cortex to fluctuat<strong>in</strong>g <strong>in</strong> vivo-like <strong>in</strong>put. First, random excitatoryand <strong>in</strong>hibitory conductances, g AMPA (t) and g GABA (t) (Fig. 8.17b), were generated us<strong>in</strong>gthe Ornste<strong>in</strong>-Uhlenbeck stochastic process (Uhlenbeck and Ornste<strong>in</strong> 1930), which wasorig<strong>in</strong>ally developed to describe Brownian motion, but can equally well describe <strong>in</strong>vivo-like fluctuat<strong>in</strong>g synaptic conductances produced by random fir<strong>in</strong>gs (Destexhe etal. 2001). Let E AMPA = 0 mV and E GABA = −65 mV denote the reverse potentials ofexcitatory and <strong>in</strong>hibitory synapses, respectively. The correspond<strong>in</strong>g currentI(t) =excitatory <strong>in</strong>put{ }} {g AMPA (t)(E AMPA − V (t)) +<strong>in</strong>hibitory <strong>in</strong>put{ }} {g GABA (t)(E GABA − V (t)) , (8.7)was <strong>in</strong>jected <strong>in</strong>to the neuron us<strong>in</strong>g the dynamic clamp protocol (Sharp et al. 1993),where V (t) denotes the <strong>in</strong>stantaneous membrane potential of the neuron. The sameconductances were <strong>in</strong>jected <strong>in</strong>to the simple model (8.5, 8.6), whose parameters wereadjusted to fit this particular neuron. The superimposed voltage traces, depicted <strong>in</strong>


Simple Models 301Fig. 8.17a, show a reasonable fit. The simple model predicts more than 90% of spikesof the <strong>in</strong> vitro neuron, often with a submillisecond precision, see Fig. 8.17c. Of course,we should not expect to get a total fit, s<strong>in</strong>ce we do not explicitly model the sourcesof <strong>in</strong>tr<strong>in</strong>sic and synaptic noise present <strong>in</strong> the cortical slice. In fact, presentation of thesame <strong>in</strong>put to the same neuron a few m<strong>in</strong>utes later produces a response with spikejitter, miss<strong>in</strong>g spikes, and extra spikes (as <strong>in</strong> Fig. 7.24) comparable with those <strong>in</strong> thesimulated response.8.2.2 Intr<strong>in</strong>sically burst<strong>in</strong>g (IB) neuronsThe class of <strong>in</strong>tr<strong>in</strong>sically burst<strong>in</strong>g (IB) neurons forms a cont<strong>in</strong>uum of cells that differ<strong>in</strong> their degree of “burst<strong>in</strong>ess”, and it probably should consists of subclasses. On oneextreme, responses of IB neurons to <strong>in</strong>jected pulses of dc-current have <strong>in</strong>itial stereotypicalbursts (Fig. 8.18a) of high-frequency spikes followed by low-frequency tonicspik<strong>in</strong>g. Many IB neurons bursts even when the current is barely superthreshold andnot strong enough to elicit a susta<strong>in</strong>ed response (as <strong>in</strong> Fig. 8.21, bottom). On theother extreme, bursts could be seen only <strong>in</strong> response to sufficiently strong current,as <strong>in</strong> Fig. 8.11 or Fig. 9.1b. Weaker stimulation elicits regular spik<strong>in</strong>g responses. Incomparison with typical RS neurons, the regular spik<strong>in</strong>g response of IB neurons haslower fir<strong>in</strong>g frequency, higher rheobase (threshold) current, exhibit shorter latency tothe first spike and noticeable afterdepolarizations (ADPs), compare RS and IB cell <strong>in</strong>Fig. 8.11.Magnifications of the responses of two IB neurons <strong>in</strong> Fig. 8.18b and c show that the<strong>in</strong>terspike <strong>in</strong>tervals with<strong>in</strong> the burst may be <strong>in</strong>creas<strong>in</strong>g or decreas<strong>in</strong>g, reflect<strong>in</strong>g possiblydifferent ionic mechanisms of burst generation and term<strong>in</strong>ation. In any case, the <strong>in</strong>itialhigh-frequency spik<strong>in</strong>g is caused by the excess of the <strong>in</strong>ward current or the deficit ofthe outward current needed to repolarize the membrane potential below the threshold.As a result, many spikes are needed to build-up outward current to term<strong>in</strong>ate thehigh-frequency burst. After the neuron recovers, it fires low frequency tonic spikesbecause there is a residual outward current (or residual <strong>in</strong>activation of <strong>in</strong>ward current)that prevents the occurrence of another burst. Many IB neurons can actually fire twoor more bursts before they switch <strong>in</strong>to tonic spik<strong>in</strong>g mode, as <strong>in</strong> Fig. 8.18a. Belowwe present two models of IB neurons, one rely<strong>in</strong>g on the <strong>in</strong>terplay of voltage-gatedcurrents, another rely<strong>in</strong>g on the <strong>in</strong>terplay of fast somatic and slow dendritic spikes.Let us use the available data on the IB neuron <strong>in</strong> Fig. 8.11 to build a simple onecompartmentmodel (8.5, 8.6) exhibit<strong>in</strong>g IB fir<strong>in</strong>g patterns. The neuron has rest<strong>in</strong>gstate at v r = −75 mV and <strong>in</strong>stantaneous threshold at v t = −45 mV. Its rheobase is350 pA, and the <strong>in</strong>put resistance is around 30 MΩ, result<strong>in</strong>g <strong>in</strong> k = 1.2 and b = 5.The peak of the spike is at +50 mV, and the after-spike resett<strong>in</strong>g po<strong>in</strong>t is aroundc = −56 mV. The parameters a = 0.01 and d = 130 give a reasonable fit of theneuron’s current-frequency relationship.The phase portraits <strong>in</strong> Fig. 8.19 expla<strong>in</strong> the mechanism of fir<strong>in</strong>g of IB patterns <strong>in</strong>the simple model. When I = 0, the model has an equilibrium at −75 mV, which is


302 Simple Models(a)burst<strong>in</strong>gspik<strong>in</strong>g20 mVADP100 ms(b)<strong>in</strong>creas<strong>in</strong>g ISIs(c)decreas<strong>in</strong>g ISIs20 mV 20 mV100 ms 100 ms500 pA 600 pAFigure 8.18: (a) burst<strong>in</strong>g and spik<strong>in</strong>g <strong>in</strong> an IB neuron (layer 5 of somatosensory cortexof a 4 week old rat at 35C; data was k<strong>in</strong>dly provided by Greg Stuart and MaartenKole). Notice the afterdepolarization (ADP). (b) IB neuron of a cat (modified fromFig. 2 of Timofeev et al. (2000)). (c) pyramidal neuron of rat’s visual cortex. Noticethat IB neurons may exhibit bursts with <strong>in</strong>creas<strong>in</strong>g or decreas<strong>in</strong>g <strong>in</strong>ter-spike <strong>in</strong>tervals(ISIs).the <strong>in</strong>tersection of the v-nullcl<strong>in</strong>e (dashed parabola) and the u-nullcl<strong>in</strong>e (straight l<strong>in</strong>e).Injection of a depolariz<strong>in</strong>g current moves the v-nullcl<strong>in</strong>e up. The pulse of current ofmagnitude I = 300 pA is below the neuron’s rheobase, so the trajectory moves fromthe old rest<strong>in</strong>g state (black square) to the new one (black circle). S<strong>in</strong>ce b > 0, thetrajectory overshoots the new equilibrium. The pulse of magnitude I = 370 pA isbarely above the rheobase, so that the model exhibits low-frequency tonic fir<strong>in</strong>g withsome spike frequency adaptation. Elevat<strong>in</strong>g the fast nullcl<strong>in</strong>e by <strong>in</strong>ject<strong>in</strong>g I = 500pA transforms the first spike <strong>in</strong>to a doublet. Indeed, the after-the-first-spike resett<strong>in</strong>gpo<strong>in</strong>t (white square marked “1”) is below the parabola, so the second spike is firedimmediately. Similarly, <strong>in</strong>jection of an even stronger current of magnitude I = 550 pAtransforms the doublet <strong>in</strong>to a burst of three spikes, each rais<strong>in</strong>g the after-spike resett<strong>in</strong>gpo<strong>in</strong>t. Once the resett<strong>in</strong>g po<strong>in</strong>t is <strong>in</strong>side the parabola, the neuron is <strong>in</strong> tonic spik<strong>in</strong>gmode.Figure 8.20a shows simultaneous record<strong>in</strong>g of somatic and dendritic membrane potentialsof a layer 5 pyramidal neuron. Somatic spike backpropagates <strong>in</strong>to the dendrite,activates voltage-gated dendritic Na + and Ca 2+ currents (Stuart et al. 1999, Hausseret al. 2000), and results <strong>in</strong> a slower dendritic spike clearly seen <strong>in</strong> the figure. The slowdendritic spike depolarizes the soma result<strong>in</strong>g <strong>in</strong> an ADP, which is typical <strong>in</strong> many


Simple Models 303layer 5 neuronsimple modelAHP3421550 pAAHP500 pA21-50 mV370 pA200 msreset21-75 mV12spike-60 mV-75 mVpeakrecovery, u4002000restpeakv-nullcl<strong>in</strong>ev-nullcl<strong>in</strong>e, I=0300 pA<strong>in</strong>put<strong>in</strong>put-80 -60 -40 -20 0membrane potential, v (mV)Figure 8.19: Comparison of <strong>in</strong> vitro record<strong>in</strong>gs of an <strong>in</strong>tr<strong>in</strong>sically burst<strong>in</strong>g (IB) neuronwith the simple model 150 ˙v = 1.2(v + 75)(v + 45) − u + I, ˙u = 0.01{5(v + 75) − u},if v ≥ +50, then v ← −56, u ← u + 130. White squares denote the reset po<strong>in</strong>tsnumbered accord<strong>in</strong>g to the spike number.IB cells. Depend<strong>in</strong>g on the strength of the <strong>in</strong>jected dc current and the state of theneuron, the ADP can be large enough to cause another somatic spike, as we illustrate<strong>in</strong> Fig. 7.52. The somatic spike may <strong>in</strong>itiate another dendritic spike, etc., result<strong>in</strong>g <strong>in</strong>a burst <strong>in</strong> Fig. 8.20b. This mechanism is known as the dendritic-somatic p<strong>in</strong>g-pong(Wang 1999) and it occurs <strong>in</strong> the P<strong>in</strong>sky-R<strong>in</strong>zel (1994) model of hippocampal CA3 neuron,the sensory neuron of weakly electric fish (Doiron et al. 2002), and <strong>in</strong> chatter<strong>in</strong>gneurons considered below.Let us build a two-compartment simple model that simulates the somatic and dendriticspike generation of IB neurons. S<strong>in</strong>ce we do not know the rheobase, <strong>in</strong>putresistance, rest<strong>in</strong>g and <strong>in</strong>stantaneous threshold potentials of dendritic tree of IB neurons,we cannot determ<strong>in</strong>e parameters of the dendritic compartment. Instead, we feedthe somatic record<strong>in</strong>g V (t) <strong>in</strong> Fig. 8.20a <strong>in</strong>to the model dendritic compartment and


304 Simple Models(a)dendriticspike20 mV25 ms(b)somadendriterecorded (<strong>in</strong> vitro)(c)ADP(d)12345simulated6dendritic recovery variable, udvd-nullcl<strong>in</strong>e6rest<strong>in</strong>g1reset452evoked bysomatic spikepeak of spike3u d -nullcl<strong>in</strong>edendritic membrane potential, v dFigure 8.20: Somatic and dendritic spike (a) and burst (b) <strong>in</strong> an IB neuron. Thedendritic spike <strong>in</strong> (a) is simulated <strong>in</strong> (c) us<strong>in</strong>g the simple model described <strong>in</strong> Fig. 8.20.Phase portrait (d) describes the geometry of dendritic spike-generation mechanism.(Record<strong>in</strong>gs are from layer 5 of somatosensory cortex of a 4 week old rat at 35C;dendritic electrode is 0.43mm from the soma; data was k<strong>in</strong>dly provided by Greg Stuartand Maarten Kole).f<strong>in</strong>e-tune the parameters so that simulated dendritic spike <strong>in</strong> Fig. 8.20c “looks like”the recorded one <strong>in</strong> Fig. 8.20a.The phase portrait <strong>in</strong> Fig. 8.20d expla<strong>in</strong>s the peculiarities of the shape of simulateddendritic spike. Recorded somatic spike quickly depolarizes the dendritic membranepotential from po<strong>in</strong>t 1 to po<strong>in</strong>t 2, and starts the regenerative process — the up-strokeof a spike. Upon reach<strong>in</strong>g the peak of the spike (3), the dendritic membrane potentialand the recovery variable are reset by the action of fast voltage-gated K + currents,which are not modeled here explicitly. The reset po<strong>in</strong>t (4) is near the stable manifoldof the saddle, so the membrane potential slowly repolarizes (5) and returns to therest<strong>in</strong>g state (6).In Fig. 8.21 we put the somatic and dendritic compartment together, adjust someof the parameters, and simulate the response of the IB neuron to pulses of current ofvarious amplitudes. Notice that the model correctly reproduces the transient burstof 2 closely spaced spikes when stimulation is weak, and the rhythmic burst<strong>in</strong>g withdecreas<strong>in</strong>g number of spikes per burst when stimulation is strong. Us<strong>in</strong>g this approach,one can build models of pyramidal neurons hav<strong>in</strong>g multiple dendritic compartments,


Simple Models 305layer 5 neuronsimple modelI=700 pAI=560 pA+50 mV20 mV100 msI=400 pAFigure 8.21: Comparison of <strong>in</strong> vitro record<strong>in</strong>gs of an <strong>in</strong>tr<strong>in</strong>sically burst<strong>in</strong>g (IB) neuron(layer 5 of somatosensory cortex of a 4 week old rat at 35C; data was k<strong>in</strong>dly providedby Greg Stuart and Maarten Kole) with the two-compartment simple model. Soma:150 ˙v = 3(v + 70)(v + 45) + 50(v d − v) − u + I, ˙u = 0.01{5(v + 70) − u}, if v ≥ +50,then v ← −52, u ← u + 240. Active dendrite: 30 ˙v d = (v d + 50) 2 + 20(v − v d ) − u d ,˙u d = 3{15(v d + 50) − u d }, if v d ≥ +20, then v d ← −20, u d ← u d + 500.as we do next.8.2.3 Multi-compartment dendritic treeIn Fig. 8.22 we simulate an IB pyramidal neuron hav<strong>in</strong>g 47 compartments (Fig. 8.22a,b),each described by a simple model with parameters provided <strong>in</strong> the caption of the figure.Our goal is to illustrate a number of <strong>in</strong>terest<strong>in</strong>g phenomena that occur <strong>in</strong> neuronalmodels hav<strong>in</strong>g active dendrites, i.e., dendrites capable of generat<strong>in</strong>g action potentials.In Fig. 8.22c we <strong>in</strong>ject a current <strong>in</strong>to compartment 4 on the apical dendrite toevoke an excitatory post-synaptic potential (EPSP) of 4 mV, which is subthresholdfor the spike-generation mechanism. This depolarization produces a current that passivelyspreads to neighbor<strong>in</strong>g compartments, and eventually <strong>in</strong>to the somatic compartment.However, the somatic EPSP is much weaker, only 1 mV, reflect<strong>in</strong>g thedistance-dependent attenuation of dendritic synaptic <strong>in</strong>puts. Notice also that somaticEPSP is delayed and it has a wider time course, which is the result of dendritic low-passfilter<strong>in</strong>g, or smooth<strong>in</strong>g, of subthreshold neuronal signals. The further the stimulationsite is from the soma, the weaker, more delayed, and longer last<strong>in</strong>g is the somaticEPSP. For many years, dendrites were thought to be passive conductors whose sole


306 Simple Models(a)74 5 6 3210(b)C 7C 6C 5C 4C 3C 2C 1C 0(c)-60 mV10 msEPSP = 4 mV(compartment C 2 )EPSP = 1 mV(soma, C 0 )synaptic <strong>in</strong>putto compartment C 2(d) C 6 +C 7 (e) C 6 +C 6 (f) C 6 +C 6 +Iall(g) soma(h) soma+IallC 7EPSP=12 mVEPSP=12 mVC 6C 5C 4C 3C 2C 120 ms50 mVsomaC 0failure to propagateforward-propagat<strong>in</strong>g action potentialfailure to propagateback-propagat<strong>in</strong>g action potentialFigure 8.22: (a) Hand draw<strong>in</strong>g and (b) a 47-compartment representation of a layer5 pyramidal neuron. (c) Injection of excitatory synaptic <strong>in</strong>put <strong>in</strong>to the compartment2 evokes a large excitatory postsynaptic potential (EPSP) <strong>in</strong> that compartment, butmuch smaller EPSP <strong>in</strong> the somatic compartment. (d) Synaptic <strong>in</strong>puts to compartments6 and 7 result <strong>in</strong> large EPSPs there, but no dendritic spike. (e) The same synaptic<strong>in</strong>puts <strong>in</strong>to compartment 6 result <strong>in</strong> dendritic spike, which fails to propagate forwardto the soma. (f) The same <strong>in</strong>put comb<strong>in</strong>ed with background excitation I all = 60 pA toall compartments results <strong>in</strong> forward propagat<strong>in</strong>g dendritic spikes. (g) Strong synaptic<strong>in</strong>put to the soma results <strong>in</strong> a spike that fails to propagate <strong>in</strong>to the dendritic tree.(h) The same <strong>in</strong>put comb<strong>in</strong>ed with <strong>in</strong>jection of I all = 70 pA to all compartments (tosimulate <strong>in</strong> vivo tonic background <strong>in</strong>put) promotes back-propagation of spike <strong>in</strong>to thedendritic tree. Each compartment is simulated by the simple model 100 ˙v = 3(v +60)(v + 50) − u + I, ˙u = 0.01{5(v + 60) − u}. Soma: if v ≥ +50, then v ← −55,u ← u + 500. Dendrites: if v ≥ +10, then v ← −35, u ← u + 1000. The conductancebetween any two adjacent compartments is 70 nS.


Simple Models 307purpose is to collect and low-pass filter the synaptic <strong>in</strong>put.Now, we explore the active properties of dendrites and their dependence on the location,tim<strong>in</strong>g, and strength of synaptic <strong>in</strong>put. First, let us stimulate two synapses that<strong>in</strong>nervate two sister dendritic compartments, e.g., compartments 6 and 7 <strong>in</strong> Fig. 8.22dthat could <strong>in</strong>teract via their mother compartment 5. Each synaptic <strong>in</strong>put evokes astrong EPSP of 12 mV, but due to their separation, the EPSPs do not add up andno dendritic spike is fired. The result<strong>in</strong>g somatic EPSP is only 0.15 mV due to thepassive attenuation. In Fig. 8.22e we provide exactly the same synaptic <strong>in</strong>put, but<strong>in</strong>to the same compartment, i.e., compartment 6. The EPSPs add up and result <strong>in</strong> adendritic spike, which propagates <strong>in</strong>to the mother compartment 5 and then <strong>in</strong>to thesister compartment 7 (which was not stimulated), but it fails to propagate along theapical dendrite <strong>in</strong>to the soma. Nevertheless, the somatic compartment exhibits anEPSP of 1.5 mV, hardly seen <strong>in</strong> the figure. Thus, the location of synaptic stimulation,with all other conditions be<strong>in</strong>g equal, made a difference. In Fig. 8.22f we comb<strong>in</strong>e thesynaptic stimulation to compartment 6 with <strong>in</strong>jection of a weak current, I all , to allcompartments of the neuron. This current represents a tonic background excitation tothe neuron that is always present <strong>in</strong> vivo. It depolarizes the membrane potential by 2.5mV and facilitates the propagation of the dendritic spike along the apical dendrite allthe way <strong>in</strong>to the soma. The same effect could be achieved by an appropriately timedexcitatory synaptic <strong>in</strong>put arriv<strong>in</strong>g to an <strong>in</strong>termediate compartment, e.g., compartment3 or 2. Not surpris<strong>in</strong>gly, an appropriately timed <strong>in</strong>hibitory <strong>in</strong>put to an <strong>in</strong>termediatecompartment on the apical dendrite could stop the forward-propagat<strong>in</strong>g dendritic spike<strong>in</strong> Fig. 8.22f.In Fig. 8.22g and h we illustrate the opposite phenomenon — back-propagat<strong>in</strong>gspikes from soma to dendrites. A superthreshold stimulation of the somatic compartmentevokes a burst of three spikes, which fails to propagate along the apical dendritesalone, but can propagate if comb<strong>in</strong>ed with a tonic depolarization of the dendritic tree.We see that dendritic trees can do more than just averag<strong>in</strong>g and low-pass filter<strong>in</strong>g ofdistributed synaptic <strong>in</strong>puts. Separate parts of the tree can perform <strong>in</strong>dependent localsignal process<strong>in</strong>g and even fire dendritic spikes. Depend<strong>in</strong>g on the synaptic <strong>in</strong>putsto other parts of the tree, the spikes can be localized or they can forward-propagate<strong>in</strong>to the soma, caus<strong>in</strong>g the cell to fire. Spikes at the soma can backpropagate <strong>in</strong>to thedendrites, trigger<strong>in</strong>g spike-time-dependent processes, such as synaptic plasticity.8.2.4 Chatter<strong>in</strong>g (CH) neuronsChatter<strong>in</strong>g neurons, also known as fast rhythmic burst<strong>in</strong>g (FRB) neurons, generatehigh-frequency repetitive bursts <strong>in</strong> response to <strong>in</strong>jected depolariz<strong>in</strong>g currents. Themagnitude of the dc-current determ<strong>in</strong>es the <strong>in</strong>terburst period, which could be as longas 100 ms or as short as 15 ms, and the number of spikes with<strong>in</strong> each burst, typically2 to 5, as we illustrate <strong>in</strong> Fig. 8.23 us<strong>in</strong>g <strong>in</strong> vivo record<strong>in</strong>gs of pyramidal neuron of catvisual cortex.An RS model neuron as shown <strong>in</strong> Fig. 8.12 can be easily transformed <strong>in</strong>to a CH


308 Simple Modelschatter<strong>in</strong>g neuron (<strong>in</strong> vivo)simple modelI=600 pAI=400 pAI=300 pA+25 mVI=200 pA50 ms-40 mVFigure 8.23: Comparison of <strong>in</strong> vivo record<strong>in</strong>gs from cat’s primary visual cortex withsimulations of the simple model 50 ˙v = 1.5(v+60)(v+40)−u+I, ˙u = 0.03{(v+60)−u},if v ≥ +25, then v ← −40, u ← u+150. Data were k<strong>in</strong>dly provided by D. McCormick,and are available on the author’s webpage.neuron by <strong>in</strong>creas<strong>in</strong>g the after-spike reset voltage to c = −40 mV, mimick<strong>in</strong>g decreasedK + and <strong>in</strong>creased Na + currents activated dur<strong>in</strong>g each spike. The phase portrait <strong>in</strong>Fig. 8.24 expla<strong>in</strong>s the mechanism of chatter<strong>in</strong>g of the simple model (8.5, 8.6). Astep of depolariz<strong>in</strong>g current shifts the fast quadratic nullcl<strong>in</strong>e up and the trajectoryquickly moves rightward to fire a spike. The after-spike reset po<strong>in</strong>t (white squaremarked “1” <strong>in</strong> the figure) is outside the parabola nullcl<strong>in</strong>e, so another spike is firedimmediately, etc., until the total amount of outward current is large enough to stopthe burst, that is, until the variable u moves the reset po<strong>in</strong>t (white square marked“5”) <strong>in</strong>side the quadratic parabola. The trajectory makes a brief excursion to theleft knee (afterhyperpolarization) and then moves rightward aga<strong>in</strong>, <strong>in</strong>itiat<strong>in</strong>g anotherburst. S<strong>in</strong>ce the second burst starts with an elevated value of u, it has fewer spikes —a phenomenon exhibited by many CH neurons.8.2.5 Low-threshold spik<strong>in</strong>g (LTS) <strong>in</strong>terneuronsLow-threshold spik<strong>in</strong>g <strong>in</strong>terneurons behave similarly to RS excitatory neurons (b > 0)<strong>in</strong> the sense that they exhibit regular spik<strong>in</strong>g patterns <strong>in</strong> response to <strong>in</strong>jected pulses of


Simple Models 309100080010 msrecovery variable, u600400200v-nullcl<strong>in</strong>e, I=550 pAv-nullcl<strong>in</strong>e, I=0 pAAHP5432112 3 4 5AHP0u-nullcl<strong>in</strong>espike-200-70 -60 -50 -40 -30 -20 -10 0 10 20membrane potential, v (mV)Figure 8.24: Phase portrait of the simple model <strong>in</strong> Fig. 8.23 exhibit<strong>in</strong>g CH fir<strong>in</strong>gpattern.current. There are some subtle differences: The response of an LTS cell to a weak depolariz<strong>in</strong>gcurrent consists of a phasic spike or a doublet with a relatively short latencyfollowed by low-frequency (less than 10 Hz) subthreshold oscillation of membrane potential.Stronger pulses elicit tonic spikes with slow frequency adaptation, decreas<strong>in</strong>gamplitudes and decreas<strong>in</strong>g after-hyperpolarizations, as one can see <strong>in</strong> Fig. 8.11.LTS neurons have more depolarized rest<strong>in</strong>g potentials, lower threshold potentials,and lower <strong>in</strong>put resistances than those of RS neurons. To match the <strong>in</strong> vitro fir<strong>in</strong>gpatterns of LTS <strong>in</strong>terneuron of rat’s barrel cortex <strong>in</strong> Fig. 8.25, we take the simple modelof RS neuron and adjust the rest<strong>in</strong>g and <strong>in</strong>stantaneous threshold potentials v r = −56mV and v t = −42 mV, and the values p = 1 and b = 8 result<strong>in</strong>g <strong>in</strong> the rheobasecurrent of 120 pA and the <strong>in</strong>put resistance of 50 MΩ. To model the decreas<strong>in</strong>g natureof the spike and AHP amplitudes, we assume that the peak of the spike and the afterspikeresett<strong>in</strong>g po<strong>in</strong>t depend on the value of the recovery variable u. This completelyunnecessary cosmetic adjustment has a mild effect on the quantitative behavior of themodel but renders a more “realistic” look to the simulated voltage traces <strong>in</strong> Fig. 8.25.The class of excitability of LTS neurons has not been studied systematically, thoughthe neurons seem to be able to fire periodic spike tra<strong>in</strong>s with a frequency as low as thatof RS neurons (Beierle<strong>in</strong> et al. 2003, Tateno and Rob<strong>in</strong>son, personal communication).The conjecture that they are near saddle-node on <strong>in</strong>variant circle bifurcation, andhence are Class 1 excitable <strong>in</strong>tegrators, seems to be at odds with the observation thattheir membrane potential exhibits slow damped oscillation and that they can fire post<strong>in</strong>hibitoryrebound spikes (Bacci et al. 2003), called low-threshold spikes (hence thename). They are better characterized as be<strong>in</strong>g at the transition from <strong>in</strong>tegrators to


310 Simple ModelsLTS neuron (<strong>in</strong> vitro)simple modelI=300 pAspike peakresetdecreas<strong>in</strong>g AHPsdecreas<strong>in</strong>g amplitudesI=200 pAspike peak20 mV100 msresetv-nullcl<strong>in</strong>ev-nullcl<strong>in</strong>e, I=0u-nullcl<strong>in</strong>eI=125 pAspike peak-56 mVrestdampedoscillationrecovery, u3002001000attractiondoma<strong>in</strong>focusspikeI=100 pAspike peak-60 -40 -20 0 20 40membrane potential, v (mV)Figure 8.25: Comparison of <strong>in</strong> vitro record<strong>in</strong>gs of a low threshold spik<strong>in</strong>g (LTS) <strong>in</strong>terneuron(rat’s barrel cortex, data provided by B. Connors) with simulations of thesimple model 100 ˙v = (v + 56)(v + 42) − u + I, ˙u = 0.03{8(v + 56) − u}, if v ≥ 40 − 0.1u,then v ← −53 + 0.04u, u ← u + 20.resonators, with phase portraits as <strong>in</strong> Fig. 8.15.A possible explanation for the subthreshold oscillations <strong>in</strong> LTS (and some RS)neurons is given <strong>in</strong> Fig. 8.13, case b > 0. The rest<strong>in</strong>g state is a stable node whenI = 0, but it becomes a stable focus when the magnitude of the <strong>in</strong>jected current isnear the neuron’s rheobase. After fir<strong>in</strong>g a phasic spike, the trajectory spirals <strong>in</strong> tothe focus exhibit<strong>in</strong>g damped oscillation. Its frequency is the imag<strong>in</strong>ary part of thecomplex-conjugate eigenvalues of the equilibrium, and it is small because the systemis near Bogdanov-Takens bifurcation.A possible explanation for the rebound spike <strong>in</strong> LTS (or some RS) neurons is given<strong>in</strong> Fig. 8.26. The shaded region is the attraction doma<strong>in</strong> of the rest<strong>in</strong>g state (blackcircle), which is bounded by the stable manifold of the saddle (white circle). A sufficientlystrong hyperpolarized pulse moves the trajectory to the new, hyperpolarized


Simple Models 311membrane potential, v (mV)40200-20-40rest<strong>in</strong>g-60-80I=0I=0I=-1 nA<strong>in</strong>jected currentAsagB0 100 200 300time (ms)CDhyperpolarized staterest<strong>in</strong>grecovery variable, u2001000-100-200v-nullcl<strong>in</strong>e, I=-1 nAsagBv-nullcl<strong>in</strong>e, I=0ADresetrest<strong>in</strong>g-80 -60 -40 -20membrane potential, v (mV)Chyperpolarized statesaddlereboundspikeFigure 8.26: The mechanism of sag and rebound spike of the model <strong>in</strong> Fig. 8.25.equilibrium (black square), which is outside the attraction doma<strong>in</strong>. Upon release fromthe hyperpolarization, the trajectory fires a phasic spike and then returns to the rest<strong>in</strong>gstate. Some LTS <strong>in</strong>terneurons fire bursts of spikes, and for that reason are calledburst-spik<strong>in</strong>g non-pyramidal (BSNP) neurons.8.2.6 Fast spik<strong>in</strong>g (FS) <strong>in</strong>terneuronsFast spik<strong>in</strong>g neurons fire “fast” tonic spike tra<strong>in</strong>s of relatively constant amplitudeand frequency <strong>in</strong> response to depolarized pulses of current. In a systematic study,Tateno et al. (2004) have shown that FS neurons have Class 2 excitability <strong>in</strong> the sensethat their frequency-current (F-I) relation has a discont<strong>in</strong>uity around 20 Hz. Whenstimulated with barely superthreshold current, such neurons exhibit irregular fir<strong>in</strong>grandomly switch<strong>in</strong>g between spik<strong>in</strong>g and fast subthreshold oscillatory mode (Kubotaand Kawaguchi 1999, Tateno et al. 2004).The absence of spike frequency adaptation <strong>in</strong> FS neurons is mostly due to the fastK + current that activates dur<strong>in</strong>g the spike, produces deep AHP, completely de<strong>in</strong>activatesNa + current, and thereby facilitates the generation of the next spike. Block<strong>in</strong>gthe K + current by TEA (Erisir et al. 1999) removes AHP, leaves residual <strong>in</strong>activationof the Na + current and slows down the spik<strong>in</strong>g, essentially transform<strong>in</strong>g the FS fir<strong>in</strong>gpattern <strong>in</strong>to LTS.The existence of fast subthreshold oscillations of membrane potential suggest thatthe rest<strong>in</strong>g state of the FS neurons is near Andronov-Hopf bifurcation. Stutter<strong>in</strong>gbehavior at the threshold currents po<strong>in</strong>ts to the co-existence of rest<strong>in</strong>g and spik<strong>in</strong>gstates, as <strong>in</strong> Fig. 8.16, and suggests that the bifurcation is of the subcritical type.However, FS neurons do not fire post-<strong>in</strong>hibitory (rebound) spikes — the feature usedto dist<strong>in</strong>guish them experimentally from LTS types. Thus, we cannot use the simplemodel (8.5, 8.6) <strong>in</strong> its present form to simulate FS neurons because the model withl<strong>in</strong>ear slow nullcl<strong>in</strong>e would fire rebound spikes accord<strong>in</strong>g to the mechanism depicted <strong>in</strong>Fig. 8.26. In addition, the simple model has a non-monotone I-V relation, whereas FS


312 Simple ModelsFS neuron (<strong>in</strong> vitro)simple model2 31I=400 pA2 3121I=200 pAAHP1 2AHPrecovery variable, u4003002001000I=100 pA-60 -50 -40 -30 -20 -10membrane potenrial, v (mV)20 mV40 ms-55 mVresetrecovery variable, u4020spik<strong>in</strong>g limitv-nullcl<strong>in</strong>ecycleu-nullcl<strong>in</strong>eI=73.2 pAspike0rest-55 -50 -45 -40membrane potenrial, v (mV)Figure 8.27: Comparison of <strong>in</strong> vitro record<strong>in</strong>gs of a fast spik<strong>in</strong>g (FS) <strong>in</strong>terneuron of layer5 rat’s visual cortex with simulations of the simple model 20 ˙v = (v +55)(v +40)−u+I,˙u = 0.2{U(v) − u}, if v ≥ 25, then v ← −45 mV. Slow nonl<strong>in</strong>ear nullcl<strong>in</strong>e U(v) = 0when v < v b and U(v) = 0.025(v − v b ) 3 when v ≥ v b with v b = −55 mV. Shaded areadenotes the attraction doma<strong>in</strong> of the rest<strong>in</strong>g state.neurons have monotone relation.The absence of rebound responses <strong>in</strong> FS neurons means that the phenomenologicalrecovery variable (activation of fast K + current) does not decrease significantly belowthe rest<strong>in</strong>g value when the membrane potential is hyperpolarized. That is, the slowu-nullcl<strong>in</strong>e becomes horizontal <strong>in</strong> the hyperpolarized voltage range. Accord<strong>in</strong>gly, wesimulate FS neuron <strong>in</strong> Fig. 8.27 us<strong>in</strong>g the simple model (8.5) with nonl<strong>in</strong>ear u-nullcl<strong>in</strong>e.The phase portraits and bifurcation diagram of the FS neuron model are qualitativelysimilar to the fast subsystem of a “subHopf/fold cycle” burster: Injection ofdc-current I creates a stable and unstable limit cycles via fold limit cycle bifurcation.The frequency of the newborn stable cycle is around 20 Hz, hence the discont<strong>in</strong>uityof the F-I curve and Class 2 excitability. There is a bistability of rest<strong>in</strong>g and spik<strong>in</strong>gstates, as <strong>in</strong> Fig. 8.27, bottom, so that noise can switch the state of the neuron back


Simple Models 313LS neuron (<strong>in</strong> vitro)simple model100 ms20 mVI=200 pAI=150 pAI=125 pAFigure 8.28: Comparison of <strong>in</strong> vitro record<strong>in</strong>gs of a late spik<strong>in</strong>g (LS) <strong>in</strong>terneuron oflayer 1 rat’s neocortex with simulations of the simple two-compartment model. Soma:20 ˙v = 0.3(v + 66)(v + 40) + 1.2(v d − v) − u + I, ˙u = 0.17{5(v + 66) − u}, if v ≥ 30, thenv ← −45, u ← u+100. Passive dendrite (dotted curve): ˙v d = 0.01(v−v d ). Weak noisewas added to simulations to unmask the subthreshold oscillations. (Record<strong>in</strong>gs werek<strong>in</strong>dly provided by Zhiguo Chu, Mario Galarreta and Shaul Hestr<strong>in</strong>. Traces I = 125and I = 150 are from one cell, trace I = 200 is from another cell.)and forth and result <strong>in</strong> irregular stutter<strong>in</strong>g spik<strong>in</strong>g with subthreshold oscillations <strong>in</strong>the 10-40 Hz range between the spike tra<strong>in</strong>s. Further <strong>in</strong>crease of I shr<strong>in</strong>ks the amplitudeof the unstable limit cycle, results <strong>in</strong> the subcritical Andronov-Hopf bifurcation ofthe rest<strong>in</strong>g state, removes the co-existence of attractors, and leaves only tonic spik<strong>in</strong>gmode.8.2.7 Late spik<strong>in</strong>g (LS) <strong>in</strong>terneuronsWhen stimulated with long pulses of dc-current, late spik<strong>in</strong>g neurons exhibit a longvoltage ramp, barely seen <strong>in</strong> Fig. 8.28, bottom, and then switch <strong>in</strong>to a tonic fir<strong>in</strong>g mode.A stronger stimulation may evoke an immediate (transient) spike followed by a longramp and a long latency to the second spike. There are pronounced fast subthresholdoscillations dur<strong>in</strong>g the voltage ramp <strong>in</strong>dicat<strong>in</strong>g the existence of at least two time scales:


314 Simple Models(1) fast oscillations result<strong>in</strong>g from the <strong>in</strong>terplay of amplify<strong>in</strong>g and resonant currents,and (2) slow ramp result<strong>in</strong>g from the slow k<strong>in</strong>etic of an amplify<strong>in</strong>g variable, such asslow <strong>in</strong>activation of an outward current (e.g., K + A-current) or slow activation of an<strong>in</strong>ward current, or both. In addition, the ramp could result from the slow charg<strong>in</strong>g ofthe dendritic compartment of the neuron.The exact mechanism responsible for the slow ramp <strong>in</strong> LS neurons is not known atpresent. Fortunately, we do not need to know the mechanism to simulate LS neuronsus<strong>in</strong>g the simple model approach. Indeed, simple models with passive dendrites areequivalent to simple models with l<strong>in</strong>ear amplify<strong>in</strong>g currents. For example, the model <strong>in</strong>Fig. 8.28 consists of a 2-dimensional system (v, u) responsible for the spike-generationmechanism at the soma and a l<strong>in</strong>ear equation for the passive dendritic compartmentv d .When stimulated with the threshold current, i.e., just above the neuronal rheobase,LS neurons often exhibit stutter<strong>in</strong>g behavior seen <strong>in</strong> Fig. 8.28, middle. Subthresholdoscillations, voltage ramps, and stutter<strong>in</strong>g are consistent with the follow<strong>in</strong>g geometricalpicture: Abrupt onset of stimulation evokes a transient spike followed by brief hyperpolarizationand then susta<strong>in</strong>ed depolarization. While depolarized, the fast subsystemaffects the slow subsystem, e.g., slowly charges the dendritic tree or slowly <strong>in</strong>activatesthe K + current. In any case, there is a slow decrease of the outward current, or equivalently,slow <strong>in</strong>crease of the <strong>in</strong>ward current that drives the fast subsystem throughthe subcritical Andronov-Hopf bifurcation. Because of the co-existence of rest<strong>in</strong>g andspik<strong>in</strong>g states near the bifurcation, the neuron can be switched from one state to theother by the membrane noise. Once the bifurcation is passed, the neuron is <strong>in</strong> thetonic spik<strong>in</strong>g mode. Overall, LS neurons can be thought of as be<strong>in</strong>g FS neurons with aslow subsystem that damps any abrupt changes, delays the onset of spik<strong>in</strong>g, and slowsdown its frequency.8.2.8 Diversity of <strong>in</strong>hibitory <strong>in</strong>terneuronsIn contrast to excitatory neocortical pyramidal neurons, which have stereotypical morphologicaland electrophysiological classes (RS, IB, CH), <strong>in</strong>hibitory neocortical <strong>in</strong>terneuronshave wildly diverse classes with various fir<strong>in</strong>g patters that cannot be classifiedas FS, LTS, or LS. Markram et al. (2004) reviewed recent results on the relationshipbetween electrophysiology, pharmacology, immunohistochemistry and gene-expressionpatterns of <strong>in</strong>hibitory <strong>in</strong>terneurons. An extreme <strong>in</strong>terpretation of their f<strong>in</strong>d<strong>in</strong>gs is thatthere is a cont<strong>in</strong>uum of different classes of <strong>in</strong>terneurons rather than a set of 3 classes.Figure 8.29 summarizes five of the most ubiquitous groups <strong>in</strong> the cont<strong>in</strong>uum:• (NAC) non-accommodat<strong>in</strong>g <strong>in</strong>terneurons fire repetitively without frequency adaptation<strong>in</strong> response to a wide range of susta<strong>in</strong>ed somatic current <strong>in</strong>jections. ManyFS and LS neurons are of this type.• (AC) accommodat<strong>in</strong>g <strong>in</strong>terneurons fire repetitively with frequency adaptationand therefore do not reach high fir<strong>in</strong>g rates of NAC neurons. Some FS and LS


Simple Models 315cells, but mostly LTS cells are of this type.• (STUT) stutter<strong>in</strong>g <strong>in</strong>terneurons fire high-frequency clusters of regular spikes<strong>in</strong>term<strong>in</strong>gled with unpredictable periods of quiescence. Some FS and LS cellsexhibit this fir<strong>in</strong>g type.• (BST) burst<strong>in</strong>g <strong>in</strong>terneurons fire a cluster of 3 to 5 spikes rid<strong>in</strong>g on a slowdepolariz<strong>in</strong>g wave followed by a strong slow AHP.• (IS) irregular spik<strong>in</strong>g <strong>in</strong>terneurons fire s<strong>in</strong>gle spikes randomly with pronouncedfrequency accommodation.NAC and AC are the most common response types found <strong>in</strong> neocortex. Each groupcan be divided <strong>in</strong>to three subgroups depend<strong>in</strong>g on the type of the onset of the responseto a step depolarization:• (c) classical response is when the first spike has the same shape as any otherspike <strong>in</strong> the response.• (b) burst response is when the first three or more spikes are clustered <strong>in</strong>to a burst.• (d) delayed response is when there is noticeable delay before the onset of spik<strong>in</strong>g.The BST type has a different subdivision: repetitive (r), <strong>in</strong>itial (i), or transient (t)burst<strong>in</strong>g.In Fig. 8.30 we use the simple model (8.5, 8.6) to reproduce all fir<strong>in</strong>g patterns of the<strong>in</strong>terneurons, <strong>in</strong>clud<strong>in</strong>g the delayed irregular spik<strong>in</strong>g (d-IS) pattern that was omitted<strong>in</strong> Fig. 8.29. We use one-fit-all set of parameters C = 100, k = 1, v r = −60 mV andv t = −40 mV, and we vary the parameters a, b, c, and d. We do not strive to reproducethe patterns quantitatively, but only qualitatively.The parameters for the NAC and AC cells were similar to those for RS neurons, withan additional passive dendritic compartment for the delayed response. The parametersfor the STUT and IS cells were similar to those of LS <strong>in</strong>terneuron with some m<strong>in</strong>ormodifications that affect the <strong>in</strong>itial burst<strong>in</strong>ess and delays. Irregular stutter<strong>in</strong>g <strong>in</strong> thesetypes results from the co-existence of stable rest<strong>in</strong>g equilibrium and spik<strong>in</strong>g limit cycleattractor, as <strong>in</strong> the case of FS and LS neurons considered above. The level of <strong>in</strong>tr<strong>in</strong>sicnoise controls the probabilities of transitions between the attractors. The parametersfor the BST cells were similar to those of IB and CH pyramidal cells. Vary<strong>in</strong>g theparameters a, b, c, and d, we <strong>in</strong>deed can get all the fir<strong>in</strong>g patterns <strong>in</strong> Fig. 8.29 plus many<strong>in</strong>termediate patterns, thereby creat<strong>in</strong>g a cont<strong>in</strong>uum of types of <strong>in</strong>hibitory <strong>in</strong>terneurons.8.3 ThalamusThalamus is the major gateway to the neocortex <strong>in</strong> the sense that no sensory signal,such as vision, hear<strong>in</strong>g, touch, taste, etc., can reach the neocortex without first pass<strong>in</strong>g


316 Simple Modelsthrough an appropriate thalamic nucleus. Anatomically, the thalamic system consistsof three major types of neurons: thalamocortical (TC) neurons, which relay signals <strong>in</strong>tothe neocortex, reticular thalamic nucleus (RTN) neurons, and thalamic <strong>in</strong>terneurons,which provide local reciprocal <strong>in</strong>hibition (Shepherd 2004). The three types have dist<strong>in</strong>ctelectrophysiological properties and fir<strong>in</strong>g patterns.There are undoubtedly subtypes with each type of thalamic neurons, however, theclassification is not as elaborate as the one <strong>in</strong> neocortex. This, and the differencebetween species, age and various thalamic nuclei, expla<strong>in</strong>s the contradictory reportsof different fir<strong>in</strong>g patterns <strong>in</strong> presumably the same types of thalamic neurons. Belowwe use the simple model (8.5, 8.6) to simulate a “typical” TC, TRN, and <strong>in</strong>terneuron.The reader should realize, though, that our attempt is as <strong>in</strong>complete as the attemptto simulate a “typical” neocortical neuron ignor<strong>in</strong>g the fact that there are RS, IB, CH,FS, etc., cells.8.3.1 Thalamo-cortical (TC) relay neuronsThalamocortical (TC) relay neurons, the type of thalamic neurons that project sensory<strong>in</strong>put to the cortex, have two prom<strong>in</strong>ent models of fir<strong>in</strong>g, illustrated <strong>in</strong> Fig. 8.31: tonicand burst mode. Both modes are ubiquitous <strong>in</strong> vitro and <strong>in</strong> vivo, <strong>in</strong>clud<strong>in</strong>g awake andbehav<strong>in</strong>g animals, and both represent different patterns of relay of sensory <strong>in</strong>formation<strong>in</strong>to the cortex (Sherman 2001). The transition between the fir<strong>in</strong>g modes depend onthe degree of <strong>in</strong>activation of low-threshold Ca 2+ T-current (Jahnsen and Ll<strong>in</strong>as 1984,McCormick and Huguenard 1992), which <strong>in</strong> turn depends on the hold<strong>in</strong>g membranepotential of the TC neuron.In tonic mode, the rest<strong>in</strong>g membrane potential of a TC neuron is around −60 mV,which is above the <strong>in</strong>activation threshold of the T-current. The slow Ca 2+ current is<strong>in</strong>activated and is not available to contribute to spik<strong>in</strong>g behavior. The neuron fires Na + -K + tonic spikes with a relatively constant frequency that depends on the amplitude ofthe <strong>in</strong>jected current and could be as low as a few Hertz (Zhan et al. 1999). Such acell, illustrated <strong>in</strong> Fig. 8.31, is a typical Class 1 excitable system near a saddle-node on<strong>in</strong>variant circle bifurcation. It exhibits regular spik<strong>in</strong>g behavior similar to that of RSneocortical neurons. It relays transient <strong>in</strong>puts <strong>in</strong>to outputs, and for this reason, manyrefer to the tonic mode as relay mode of fir<strong>in</strong>g.To switch a TC neuron <strong>in</strong>to the burst mode, an <strong>in</strong>jected dc-current or <strong>in</strong>hibitorysynaptic <strong>in</strong>put must hyperpolarize the membrane potential to around −80 mV for atleast 50-100 ms. While hyperpolarized, the Ca 2+ T-current de<strong>in</strong>activates and becomesavailable. As soon as the membrane potential is returned to the rest<strong>in</strong>g or depolarizedstate, there is an excess of the <strong>in</strong>ward current that drives the neuron over thresholdand results <strong>in</strong> a rebound burst of high-frequency spikes, as <strong>in</strong> Fig. 8.31, called a lowthreshold(LT) spike or a Ca 2+ spike.In Fig. 8.31, right, we simulate TC neuron us<strong>in</strong>g simple model (8.5, 8.6), treat<strong>in</strong>g uas the low-threshold Ca 2+ current. S<strong>in</strong>ce the current is <strong>in</strong>activated <strong>in</strong> the tonic mode,i.e., u ≈ 0, we take b = 0. The rest<strong>in</strong>g and threshold voltages of the neuron <strong>in</strong> the


Simple Models 317figure are v r = −60 mV and v t = −50 mV. The value p = 1.6 results <strong>in</strong> 40 pA rheobasecurrent and 60 MΩ <strong>in</strong>put resistance, and the membrane capacitance C = 200 pF givesthe right current-frequency (F-I) relationship. Thus, <strong>in</strong> the tonic mode, our model isessentially the quadratic <strong>in</strong>tegrate-and-fire neuron 200 ˙v = 1.6(v + 60)(v + 50) + I withthe after-spike reset from +35 mV to −60 mV.To model slow Ca 2+ dynamics <strong>in</strong> the burst mode, we assume that hyperpolarizationsbelow the Ca 2+ <strong>in</strong>activation threshold of −65 mV decrease u, thereby creat<strong>in</strong>g <strong>in</strong>wardcurrent. In the l<strong>in</strong>ear case, we take ˙u = 0.01{b(v + 65) − u} with b = 0 when v ≥ −65and b = 15 when v < −65, result<strong>in</strong>g <strong>in</strong> the piecewise l<strong>in</strong>ear u-nullcl<strong>in</strong>e depicted <strong>in</strong>Fig. 8.31, bottom. Prolonged hyperpolarization below −65 mV decreases u and movesthe trajectory outside the attraction doma<strong>in</strong> of the rest<strong>in</strong>g state (shaded region <strong>in</strong> thefigure). Upon release from the hyperpolarization, the model fires a rebound burst ofspikes, variable u → 0 reflect<strong>in</strong>g <strong>in</strong>activation of Ca 2+ , and the trajectory reenters theattraction doma<strong>in</strong> of the rest<strong>in</strong>g state. Steps of depolarized current produce reboundbursts followed by tonic spik<strong>in</strong>g with adapt<strong>in</strong>g frequency. A better quantitative agreementwith TC record<strong>in</strong>gs can be achieved when two slow variables, u 1 and u 2 , areused.8.3.2 Reticular thalamic nucleus (RTN) neuronsReticular thalamic nucleus (RTN) neurons provide reciprocal <strong>in</strong>hibition to TC relayneurons. RTN and TC cells are similar <strong>in</strong> the sense that they have two fir<strong>in</strong>g modes,illustrated <strong>in</strong> Fig. 8.32: They fire tra<strong>in</strong>s of s<strong>in</strong>gle spikes follow<strong>in</strong>g stimulation fromrest<strong>in</strong>g or depolarized potentials <strong>in</strong> the tonic mode, and rebound bursts upon releasefrom hyperpolarized potentials <strong>in</strong> the burst mode.The parameters of the simple model <strong>in</strong> Fig. 8.32 are adjusted to match the <strong>in</strong> vitrorecord<strong>in</strong>g of the RTN cell <strong>in</strong> the figure, and they differ from the parameters of the TCmodel cell. Nevertheless, the mechanism of rebound burst<strong>in</strong>g of RTN neuron is thesame as that of the TC neuron <strong>in</strong> Fig. 8.31, bottom. In contrast, the tonic mode of fir<strong>in</strong>gis different. S<strong>in</strong>ce b > 0, the model neuron is near the transition from an <strong>in</strong>tegratorto a resonator; It can fire transient spikes followed by slow subthreshold oscillationsof membrane potential; It has co-existence of stable rest<strong>in</strong>g and spik<strong>in</strong>g states, withthe bifurcation diagram similar to the one <strong>in</strong> Fig. 8.15, and it can stutter and produceclustered spikes when stimulated with barely threshold current. Interest<strong>in</strong>gly, similarbehavior of TC neurons was reported by Pirchio et al. (1997), Pedroarena and Ll<strong>in</strong>as(1997) and Li et al. (2003). We will return to the issue of subthreshold oscillations andstutter<strong>in</strong>g spik<strong>in</strong>g when we consider stellate cells of entorh<strong>in</strong>al cortex <strong>in</strong> Sect. 8.4.4.8.3.3 Thalamic <strong>in</strong>terneuronsIn contrast to TC and RTN neurons, thalamic <strong>in</strong>terneurons do not have a prom<strong>in</strong>entburst mode, though they can fire rebound spikes upon release from hyperpolarization(Pape and McCormick 1995). They have action potentials with short duration, and


318 Simple Modelsthey are able to generate high-frequency tra<strong>in</strong>s of spikes reach<strong>in</strong>g 800 Hz, like corticalFS <strong>in</strong>terneurons. The simple model <strong>in</strong> Fig. 8.33 reproduces all these features. Its phaseportrait and bifurcation diagram is similar to the one <strong>in</strong> Fig. 8.15, but its dynamicshas a much faster time scale.8.4 Other <strong>in</strong>terest<strong>in</strong>g casesThe neocortical and thalamic neurons span an impressive range of dynamic behavior.Many neuronal types found <strong>in</strong> other bra<strong>in</strong> regions have dynamics quite similar to someof the types discussed above, while many do not.8.4.1 Hippocampal CA1 pyramidal neuronsHippocampal pyramidal neurons and <strong>in</strong>terneurons are similar to those of the neocortex,and hence could be simulated us<strong>in</strong>g the simple model presented <strong>in</strong> Sect. 8.2. Let uselaborate us<strong>in</strong>g the pyramidal neurons of CA1 region of hippocampus as an example.Jensen et al. (1994) suggested to classify all CA1 pyramidal neurons accord<strong>in</strong>g totheir propensity to fire bursts of spikes, often called complex spikes. The majority (morethan 80%) of CA1 pyramidal neurons are non-burst<strong>in</strong>g cells, whereas the rema<strong>in</strong><strong>in</strong>gexhibit some form of bursts, which are def<strong>in</strong>ed <strong>in</strong> this context as sets of three or moreclosely spaced spikes. There are five different classes:• (NB) Non-burst<strong>in</strong>g cells generate accommodat<strong>in</strong>g tra<strong>in</strong>s of tonic spikes <strong>in</strong> responseto depolariz<strong>in</strong>g pulses of dc-current and a s<strong>in</strong>gle spike <strong>in</strong> response to abrief superthreshold pulse of current, as <strong>in</strong> Fig. 8.34A.• (HTB) High-threshold bursters fire bursts only <strong>in</strong> response to strong long pulsesof current, but fire s<strong>in</strong>gle spikes <strong>in</strong> response to weak or brief pulses of current, as<strong>in</strong> Fig. 8.34B.• (LTB I) Grade I low-threshold bursters fire bursts <strong>in</strong> response to long pulses, buts<strong>in</strong>gle spikes <strong>in</strong> response to brief pulses of current, as <strong>in</strong> Fig. 8.34C.• (LTB II) Grade II low-threshold bursters fire stereotypical bursts also <strong>in</strong> responseto brief pulses, as <strong>in</strong> Fig. 8.34D.• (LTB III) F<strong>in</strong>ally, grade III low-threshold bursters fire rhythmic bursts spontaneously,which are depicted <strong>in</strong> Fig. 8.34E us<strong>in</strong>g two time scales.NB neurons are equivalent to neocortical pyramidal neurons of RS type, whereas HTBand LTB I neurons are equivalent to neocortical pyramidal neurons of IB type. Theauthor is not aware of any systematic studies of the ability of IB neurons to fire stereotypicalbursts <strong>in</strong> response to brief pulses, as <strong>in</strong> Fig. 8.34Db, or to have <strong>in</strong>tr<strong>in</strong>sic rhythmicactivity, as <strong>in</strong> Fig. 8.34E. Therefore, it is not clear whether there are any analogues ofLTB grade II and III neurons <strong>in</strong> the neocortex.


Simple Models 319The classification of hippocampal CA1 pyramidal neurons <strong>in</strong>to five different classesdoes not imply a fundamental difference <strong>in</strong> the ionic mechanism of spike-generation, butonly a quantitative difference. This follows from the observation that pharmacologicalmanipulations can gradually and reversibly transform an NB neuron <strong>in</strong>to LTB IIIneuron and visa-versa by elevat<strong>in</strong>g the extracellular concentration of K + (Jensen et al.1994), reduc<strong>in</strong>g extracellular Ca 2+ (Su et al. 2001), or block<strong>in</strong>g K + M-current (Yueand Yaari 2004), or manipulat<strong>in</strong>g Ca 2+ current dynamics <strong>in</strong> apical dendrites (Mageeand Carruth 1999).In Fig. 8.35 we modify the simple model for neocortical RS neuron to reproducefir<strong>in</strong>g patterns of hippocampal pyramidal cells. To get the cont<strong>in</strong>uum of responses, fromNB to LTB II, we fix all the parameters and vary only the after-spike reset parameterc <strong>in</strong> an <strong>in</strong>crement of 5 mV, and the parameter d. These phenomenological parametersdescribe the effect of high-threshold <strong>in</strong>ward and outward currents activated dur<strong>in</strong>g eachspike and affect<strong>in</strong>g the after-spike behavior. Increas<strong>in</strong>g c corresponds to up-regulat<strong>in</strong>gslow I Na,p or down-regulat<strong>in</strong>g slow K + currents, which leads to transition from NBto LTB III <strong>in</strong> the CA1 slice (Su et al. 2001) and <strong>in</strong> the simple model <strong>in</strong> Fig. 8.35.Interest<strong>in</strong>gly, the same procedure results <strong>in</strong> transitions from RS to IB and possibly toCH classes <strong>in</strong> neocortical pyramidal neurons (Izhikevich 2003). This is consistent withthe observation by Steriade (2004) that many neocortical neurons can change theirfir<strong>in</strong>g classes <strong>in</strong> vivo depend<strong>in</strong>g on the state of the bra<strong>in</strong>.8.4.2 Sp<strong>in</strong>y projection neurons of neostriatum and basal gangliaSp<strong>in</strong>y projection neurons, the major class of neurons <strong>in</strong> neostriatum and basal ganglia,display a prom<strong>in</strong>ent bistable behavior <strong>in</strong> vivo shown <strong>in</strong> Fig. 8.36 (Wilson and Groves1981, Wilson 1993): They shift the membrane potential from hyperpolarized to depolarizedstates <strong>in</strong> response to synchronous excitatory synaptic <strong>in</strong>put from cortex and/orthalamus. In vitro studies of such neurons reveal a slowly <strong>in</strong>activat<strong>in</strong>g K + A-current,which is believed to be responsible for the ma<strong>in</strong>tenance of the up- and down-states, <strong>in</strong>addition to the synaptic <strong>in</strong>put. Indeed, the K + current is completely de<strong>in</strong>activated atthe hyperpolarized potentials (down-states), and reduces the response of the neuronto any synaptic <strong>in</strong>put. In contrast, prolonged depolarization (up-state) <strong>in</strong>activates thecurrent and makes the neuron more excitable and ready to fire spikes.The most remarkable feature of neostriatal sp<strong>in</strong>y neurons is depicted <strong>in</strong> Fig. 8.37.In response to depolariz<strong>in</strong>g current pulses, the neurons display a prom<strong>in</strong>ent slowlydepolariz<strong>in</strong>g (ramp) potential, and hence long latency to spike discharge (Nisenbaumet al. 1994). The ramp is mostly due to the slow <strong>in</strong>activation of K + A-current andslow charg<strong>in</strong>g of the dendritic tree. The delay to spike could be as long as 1 sec, butthe subsequent spike tra<strong>in</strong> has a shorter relatively constant period that depends on themagnitude of the <strong>in</strong>jected current.Let us use the simple model (8.5, 8.6) to simulate the responses of sp<strong>in</strong>y neuronsto current pulses. The rest<strong>in</strong>g membrane potential of the neuron <strong>in</strong> Fig. 8.37 is around


320 Simple Modelsv r = −80 mV, and we set v t = −25 mV, p = 1, and b = −20 to get 30 MΩ <strong>in</strong>putresistance and 300 pA rheobase current. We take a = 0.01 to reflect the slow <strong>in</strong>activationof the K + A-current <strong>in</strong> the subthreshold voltage range. The membrane potential<strong>in</strong> the figure reaches the peak of +40 mV dur<strong>in</strong>g the spike and then resets to −55 mVor lower, depend<strong>in</strong>g on the fir<strong>in</strong>g frequency. The value d = 150 provides a reasonablematch of the <strong>in</strong>terspike frequencies for all magnitudes of <strong>in</strong>jected current. Notice thatb < 0, so that u represents either slow <strong>in</strong>activation of I A or slow charg<strong>in</strong>g of the passivedendritic compartment, or both. In any case, it is a slow amplify<strong>in</strong>g variable, which isconsistent with the observation that sp<strong>in</strong>y neurons do not “sag” <strong>in</strong> response to hyperpolariz<strong>in</strong>gcurrent pulses, do not “peak” <strong>in</strong> response to depolariz<strong>in</strong>g pulses (Nisenbaumet al. 1994), and do not generate rebound (post-<strong>in</strong>hibitory) spikes.Injection of a depolariz<strong>in</strong>g current shifts the v-nullcl<strong>in</strong>e of the simple model up, andthe rest<strong>in</strong>g state disappears via saddle-node bifurcation. The trajectory slowly movesthrough the ghost of the bifurcation po<strong>in</strong>t (shaded rectangle <strong>in</strong> the figure), result<strong>in</strong>g<strong>in</strong> the long latency to the first spike. The spike resets the trajectory to a po<strong>in</strong>t (whitesquare) below the ghost, result<strong>in</strong>g <strong>in</strong> significantly smaller delays to subsequent spikes.Because the resett<strong>in</strong>g po<strong>in</strong>t is so close to the saddle-node bifurcation po<strong>in</strong>t, the simplemodel, and probably the sp<strong>in</strong>y projection neuron <strong>in</strong> the figure, is near the co-dimension-2 saddle-node homocl<strong>in</strong>ic orbit bifurcation discussed <strong>in</strong> Sect. 6.3.6.8.4.3 Mesencephalic V neurons of bra<strong>in</strong>stemThe best examples of resonators, with fast subthreshold oscillations, Class 2 excitability,rebound spikes, etc., are mesencephalic V (mesV) neurons of bra<strong>in</strong>stem (Wu et al.2001) and primary sensory neurons of dorsal root ganglion (Amir et al. 2002, Jian etal. 2004). Mes V neurons of the bra<strong>in</strong>stem have monotone I-V curves, whereas thesimple model with l<strong>in</strong>ear equation for u does not. In Fig. 8.38 we use a modificationof the simple model to simulate the responses of mesV neuron (data from Fig. 7.3) topulses of depolariz<strong>in</strong>g current.The model’s phase portrait is qualitatively similar to that of the FS <strong>in</strong>terneurons<strong>in</strong> Fig. 8.27. The rest<strong>in</strong>g state is a stable focus, result<strong>in</strong>g <strong>in</strong> damped or noise-<strong>in</strong>ducedsusta<strong>in</strong>ed oscillations of the membrane potential. Their amplitude and frequency dependon I and could be larger than 5 mV and 100 Hz, respectively. The focus losesstability via subcritical Andronov-Hopf bifurcation. Because of the co-existence of therest<strong>in</strong>g and spik<strong>in</strong>g states, the mesV neuron can burst, and so can the simple model ifnoise or a slow resonant variable is added.8.4.4 Stellate cells of entorh<strong>in</strong>al cortexThe entorh<strong>in</strong>al cortex occupies a privileged anatomical position that allows it to gatethe ma<strong>in</strong> flow of <strong>in</strong>formation to and out of the hippocampus. In vitro studies show thatstellate cells, a major class of neurons <strong>in</strong> entorh<strong>in</strong>al cortex, exhibit <strong>in</strong>tr<strong>in</strong>sic subthresholdoscillations with a slow dynamics of the k<strong>in</strong>d shown <strong>in</strong> Fig. 8.39b (Alonso and Ll<strong>in</strong>as


Simple Models 3211989, Alonso and Kl<strong>in</strong>k 1993, Kl<strong>in</strong>k and Alonso 1993, Dickson et al. 2000). The oscillationsare generated by the <strong>in</strong>terplay between persistent Na + current and h-current, andthey are believed to set the theta rhythmicity <strong>in</strong> the entorh<strong>in</strong>al-hippocampal network.The caption of Fig. 8.39 provides parameters of the simple model (8.5, 8.6) thatcaptures the slow oscillatory dynamics of an entorh<strong>in</strong>al stellate cell recorded <strong>in</strong> vitro<strong>in</strong> adult rat. The cell sags to <strong>in</strong>jected hyperpolariz<strong>in</strong>g current <strong>in</strong> Fig. 8.39a and thenfires a rebound spike upon release from hyperpolarization. From a neurophysiologicalpo<strong>in</strong>t of view, the sag and rebound response are due to the open<strong>in</strong>g of the h-current;From a mathematical po<strong>in</strong>t of view, they are caused by the resonant slow variable u,which could also describe de<strong>in</strong>activation of a transient Na + current and deactivationof low-threshold K + current. The geometrical explanation of these responses is similarto the one provided for LTS <strong>in</strong>terneurons <strong>in</strong> Fig. 8.26. Positive steps of current evoke atransient or susta<strong>in</strong>ed spik<strong>in</strong>g activity. Notice that the first spike is actually a doublet<strong>in</strong> the record<strong>in</strong>g and <strong>in</strong> the simulation <strong>in</strong> Fig. 8.39a (I = 200 pA).Stellate cells <strong>in</strong> the entorh<strong>in</strong>al cortex of adult animals can exhibit damped or susta<strong>in</strong>edsubthreshold oscillations <strong>in</strong> a frequency range from 5 to 15 Hz. The oscillationscan be clearly seen when the cell is depolarized by <strong>in</strong>jected dc-current, as <strong>in</strong> Fig. 8.39b.The stronger the current, the higher the amplitude and frequency of oscillations, whichoccasionally result <strong>in</strong> spikes or even bursts of spikes (Alonso and Kl<strong>in</strong>k 1993). The simplemodel also exhibits slow damped oscillations because its rest<strong>in</strong>g state is a stablefocus. The focus loses stability via subcritical Andronov-Hopf bifurcation, and hence itcoexists with a spik<strong>in</strong>g limit cycle. To enable susta<strong>in</strong>ed oscillations and random spikes,we add channel noise to the v-equation (White et al. 2000).In Fig. 8.39c we expla<strong>in</strong> the mechanism of random transitions between subthresholdoscillations and spikes, which is similar to the mechanism of stutter<strong>in</strong>g <strong>in</strong> RS andFS neurons. When weak dc-current is <strong>in</strong>jected (left), the attraction doma<strong>in</strong> of therest<strong>in</strong>g state (shaded region) is separated from the rest of the phase space by thestable manifold to the saddle equilibrium (denoted separatrix). Noisy perturbationsevoke small susta<strong>in</strong>ed noisy oscillations around the rest<strong>in</strong>g state with an occasionalspike when the separatrix is crossed. Increas<strong>in</strong>g the level of <strong>in</strong>jected dc-current results<strong>in</strong> the series of bifurcations similar to those <strong>in</strong> Fig. 8.15. As a result, there is a coexistenceof a large amplitude (spik<strong>in</strong>g) limit cycle attractor and a small unstable limitcycle, which encompasses the attraction doma<strong>in</strong> of the rest<strong>in</strong>g state (right). Noisyperturbations can randomly switch the activity between these attractors, result<strong>in</strong>g <strong>in</strong>the random burst<strong>in</strong>g activity <strong>in</strong> Fig. 8.39b.8.4.5 Mitral neurons of olfactory bulbMitral cells recorded <strong>in</strong> slices of rat ma<strong>in</strong> olfactory bulb exhibit <strong>in</strong>tr<strong>in</strong>sic bistability ofmembrane potentials (Heyward et al. 2001). They spontaneously alternate betweentwo membrane potentials separated by 10 mV: a relatively depolarized (up-state) andhyperpolarized (down-state) potentials. The membrane potential could be switchedbetween the states by a brief depolariz<strong>in</strong>g or hyperpolariz<strong>in</strong>g pulse of current, as we


322 Simple Modelsshow <strong>in</strong> Fig. 7.36. In response to stimulation, the cells are more likely to fire <strong>in</strong> theup-state than <strong>in</strong> the down-state.Current-voltage (I-V) relations of such mitral cells have three zeros <strong>in</strong> the subthresholdvoltage range confirm<strong>in</strong>g that there are three equilibria, two stable correspond<strong>in</strong>gto the up- and down-state, and one unstable — saddle. There are no subthresholdoscillations <strong>in</strong> the down-state, hence it is a node, and the cell is an <strong>in</strong>tegrator. Thereare small-amplitude 40 Hz oscillations <strong>in</strong> the up-state, hence it is a focus and the cellis a resonator.To model the bistability, we use the simple model with a piece-wise l<strong>in</strong>ear slownullcl<strong>in</strong>e that approximates non-l<strong>in</strong>ear activation functions n ∞ (v) near the “threshold”of the current and a passive dendritic compartment. In many respects, the model issimilar to the one for late spik<strong>in</strong>g (LS) cortical <strong>in</strong>terneurons. In Fig. 8.40 we f<strong>in</strong>e-tunethe model to simulate responses of a rat mitral cell to pulses of current of variousamplitude. To prevent noise-<strong>in</strong>duced spontaneous transitions between the up- anddown-states, the cell <strong>in</strong> the figure was hold at −75 mV by <strong>in</strong>jection of a large negativecurrent. Its responses to weak positive pulses of current show a fast ris<strong>in</strong>g phasefollowed by an abrupt step (arrow <strong>in</strong> the figure) to a constant value correspond<strong>in</strong>gto the up-state. Increas<strong>in</strong>g the magnitude of stimulation elicits tra<strong>in</strong>s of spikes with aconsiderable latency, whose cause has yet to be determ<strong>in</strong>ed experimentally. The latencycould be the result of slow activation of an <strong>in</strong>ward current, slow <strong>in</strong>activation of anoutward current, e.g. K + A-current, or just slow charg<strong>in</strong>g of the dendritic compartment.All three cases correspond to an additional slow variable <strong>in</strong> the simple model, whichwe <strong>in</strong>terpret as a membrane potential of a passive dendritic compartment.To understand the dynamics of the simple model, and hopefully of the mitral cell,we simulate its responses <strong>in</strong> Fig. 8.41 to the activation of the olfactory nerve (ON).In the top of Fig. 8.41, the cell is held at I = 0 pA. Its phase portrait clearly showsthe co-existence of stable node and focus equilibria separated by a saddle. The shadedregion corresponds to the attraction doma<strong>in</strong> of the focus equilibrium. To fire a spikefrom the up-state, noise or external stimulation must push the state of the systemfrom the shaded region over the threshold to the right. The cell returns to the downstateright after the spike. Much stronger stimulation is needed to fire the cell fromthe down-state. Typically, the cell is switched to the up-state first, spend some timeoscillat<strong>in</strong>g at 40 Hz, and then fire a spike (Heyward et al. 2001).In the bottom of Fig. 8.41, the cell is held at a slightly depolariz<strong>in</strong>g current I =7 pA. The node equilibrium disappeared via saddle-node bifurcation, so there is nodown-state, but only its ghost. Stimulation at the up-state results <strong>in</strong> a spike, afterhyperpolarization,and slow transition through the ghost of the down-state back tothe up-state. Further <strong>in</strong>creas<strong>in</strong>g the hold<strong>in</strong>g current results <strong>in</strong> the stable manifoldto the upper saddle (marked “threshold” <strong>in</strong> the figure) to make a loop, and thenbecome a homocl<strong>in</strong>ic trajectory to the saddle giv<strong>in</strong>g birth to an unstable limit cycle,which shr<strong>in</strong>ks to the focus and makes it lose stability via subcritical Andronov-Hopfbifurcation. Notice that this phase portrait and the bifurcation scenario is differentfrom the one <strong>in</strong> Fig. 7.36. However, <strong>in</strong> both cases, the neuron is an <strong>in</strong>tegrator <strong>in</strong>


Simple Models 323the down-state and a resonator <strong>in</strong> the up-state! The same property is exhibited bycerebella Purk<strong>in</strong>je cells (see Fig. 7.37), and possibly by other neurons kept <strong>in</strong> theup-state (<strong>in</strong>tr<strong>in</strong>sically or extr<strong>in</strong>sically).Review of Important Concepts• Integrate-and-fire neuron is a l<strong>in</strong>ear model hav<strong>in</strong>g a stable node equilibrium,an artificial threshold, and a reset.• Resonate-and-fire neuron is a l<strong>in</strong>ear model hav<strong>in</strong>g a stable focusequilibrium, an artificial threshold, and a rest.• Though technically not spik<strong>in</strong>g neurons, these models are useful foranalytical studies, i.e., to prove theorems.• The quadratic <strong>in</strong>tegrate-and-fire model captures the non-l<strong>in</strong>earityof the spike-generation mechanism of real neurons hav<strong>in</strong>g Class 1excitability (saddle-node on <strong>in</strong>variant circle bifurcation).• Its simple extension, model (8.5, 8.6), reproduces quantitatively subthreshold,spik<strong>in</strong>g, and burst<strong>in</strong>g activity of all known types of corticaland thalamic neurons <strong>in</strong> response to pulses of dc-current.• The simple model makes testable hypotheses on the dynamic mechanismsof excitability <strong>in</strong> these neurons.• The model is especially suitable for simulations of large-scale modelsof the bra<strong>in</strong>.


324 Simple ModelsBibliographical NotesMany people have used the <strong>in</strong>tegrate-and-fire neuron, treat<strong>in</strong>g it as a folklore model.It was Tuckwell’s (1988) “Introduction to Theoretical Neurobiology” that gave appropriatecredit to its <strong>in</strong>ventor — Lapicque (1907). Although better models, such as thequadratic <strong>in</strong>tegrate-and-fire model, are available now, many scientists cont<strong>in</strong>ue to favorthe leaky <strong>in</strong>tegrate-and-fire neuron mostly because of its simplicity. Such an attitude isunderstandable when one wants to derive analytical results. However, purely computa-


Simple Models 325tional papers can actually suffer from us<strong>in</strong>g the model because of its weird properties,such as the logarithmic F-I curve and fixed threshold.The resonate-and-fire model was <strong>in</strong>troduced by Izhikevich (2001), and then byRichardson, Brunel, and Hakim (2003) and Brunel, Hakim, and Richardson (2003).These authors <strong>in</strong>itially called the model “resonate-and-fire”, but then changed its nameto “generalized <strong>in</strong>tegrate-and-fire” (GIF), possibly to avoid confusion.A better choice is the quadratic <strong>in</strong>tegrate-and-fire neuron <strong>in</strong> the normal form (8.2)or <strong>in</strong> the ϑ-form (8.8); see Ex. 7. The ϑ-form was first suggested <strong>in</strong> the context ofcircle/circle (parabolic) burst<strong>in</strong>g by Ermentrout and Kopell (1986a,b). Later, Ermentrout(1996) used this model to generalize numerical results by Hansel et al. (1995)on synchronization of Class 1 excitable systems, discussed <strong>in</strong> Chap. 10. Hoppensteadtand Izhikevich (1997) <strong>in</strong>troduced the canonical model approach, provided many examplesof canonical models, and proved that the quadratic <strong>in</strong>tegrate-and-fire model wascanonical <strong>in</strong> the sense that all Class 1 excitable systems could be transformed <strong>in</strong>to thismodel by a piece-wise cont<strong>in</strong>uous change of variables. They also suggested to call themodel “Ermentrout-Kopell canonical model”, but most scientists follow Ermentroutand call it “theta-neuron”.The model presented <strong>in</strong> Sect. 8.1.4 was first suggested by Izhikevich (2000; Eq. (4)and (5) with voltage reset discussed <strong>in</strong> Sect. 2.3.1) <strong>in</strong> the ϑ-form. The form presentedhere first appeared <strong>in</strong> Izhikevich (2003). The representation of the function I + v 2 <strong>in</strong>the form (v − v r )(v − v t ) was suggested by Latham et al. (2000).We stress that the simple model is useful only when one wants to simulate largescalenetworks of spik<strong>in</strong>g neurons. He or she still needs to use the Hodgk<strong>in</strong>-Huxley-typeconductance based models to study the behavior of one neuron or a small network ofneurons. The parameter values that match fir<strong>in</strong>g patterns of biological neurons presented<strong>in</strong> this chapter are only educated guesses (the same is true for conductance-basedmodels). More experiments are needed to reveal the true spike-generation mechanismof any particular neuron. An additional <strong>in</strong>sights <strong>in</strong>to the question “which model ismore realistic” is <strong>in</strong> Fig. 1.8.Look<strong>in</strong>g at the simple model, one gets an impression that the spike generationmechanism of RS neurons is the simplest <strong>in</strong> the neocortex. This is probably true,however, the complexity of the RS neurons, most of which are pyramidal cells, is hidden<strong>in</strong> their extensive dendritic trees hav<strong>in</strong>g voltage- and Ca 2+ -gated currents. Study<strong>in</strong>gdendritic dynamics is a subject of a 500-page book by itself, and we purposefullyomitted this subject. We recommend read<strong>in</strong>g Dendrites by Stuart et al. (1999), recentreviews by Hausser and Mel (2003) and Williams and Stuart (2003), and the sem<strong>in</strong>alpaper by Arshavsky et al. (1971; Russian language edition - 1969).


326 Simple ModelsExercises1. (Integrate-and-fire network) The simplest implementation of a pulse-coupled <strong>in</strong>tegrate-and-fireneural network has the form˙v i = b i − v i + ∑ j≠ic ij δ(t − t j ) ,where t j is the moment of fir<strong>in</strong>g of the jth neuron, i.e., the moment v j (t j ) =1. Thus, whenever the jth neuron fires, the membrane potentials of the otherneurons are adjusted <strong>in</strong>stantaneously by c ij , i ≠ j. Show that the same <strong>in</strong>itialconditions may result <strong>in</strong> different solutions, depend<strong>in</strong>g on the implementationdetails.2. (Latham et al. 2000) Determ<strong>in</strong>e the relationship between the normal form forsaddle-node bifurcation (6.2) and the equation˙V = a(V − V rest )(V − V thresh ) .3. Show that the period of oscillations <strong>in</strong> the quadratic <strong>in</strong>tegrate-and-fire model(8.2) isT = √ 1 (atan v peak√ − atan v )reset√b b bwhen b > 0.4. Show that the period of oscillations <strong>in</strong> the quadratic <strong>in</strong>tegrate-and-fire model(8.2) with v peak = 1 is(T = 12 √ ln 1 − √ |b||b| 1 + √ |b| − ln v reset − √ )|b|v reset + √ |b|when b < 0 and v reset > √ |b|.5. Justify the bifurcation diagram <strong>in</strong> Fig. 8.3.6. Brizzi et al. (2004) have shown that shunt<strong>in</strong>g <strong>in</strong>hibition of cat motoneurons raisesthe fir<strong>in</strong>g threshold, rheobase current, and shifts the F-I curve to the right withoutchang<strong>in</strong>g the shape of the curve. Use the quadratic <strong>in</strong>tegrate-and-fire model toexpla<strong>in</strong> the effect. (H<strong>in</strong>t: Consider ˙v = b − gv + v 2 with g ≥ 0, v reset = −∞, andv peak = +∞.)7. (Theta neuron) Determ<strong>in</strong>e when the quadratic <strong>in</strong>tegrate-and-fire neuron (8.2) isequivalent to the theta neuron˙ϑ = (1 − cos ϑ) + (1 + cos ϑ)r , (8.8)where r is the bifurcation parameter and ϑ ∈ [−π, π] is a phase variable on theunit circle.


Simple Models 3278. (Another theta neuron) Show that the quadratic <strong>in</strong>tegrate-and-fire neuron (8.2)is equivalent to˙ϑ = ϑ 2 + (1 − |ϑ|) 2 r .where ϑ ∈ [−1, 1] and r have the same mean<strong>in</strong>g as <strong>in</strong> the previous exercise. Arethere any other “theta neurons”?9. When is the l<strong>in</strong>ear version of (8.3, 8.4),˙v = I − v − u if v = 1, then˙u = a(bv − u) v ← 0, u ← u + d,equivalent to the <strong>in</strong>tegrate-and-fire or resonate-and-fire model?10. Show that the simple model (8.3, 8.4) with b < 0 is equivalent to the quadratic<strong>in</strong>tegrate-and-fire neuron with a passive dendritic compartment.11. All membrane potential responses <strong>in</strong> Fig. 8.8 were obta<strong>in</strong>ed us<strong>in</strong>g model (8.3,8.4) with appropriate values of the parameters. Use MATLAB to experimentwith the model and reproduce the figure.12. Simulate the FS spik<strong>in</strong>g neuron <strong>in</strong> Fig. 8.27 us<strong>in</strong>g simple model (8.5, 8.6) withl<strong>in</strong>ear equation for u. What can you say about its possible bifurcation structure?13. Fit the record<strong>in</strong>gs of RS neuron <strong>in</strong> Fig. 8.12 us<strong>in</strong>g the follow<strong>in</strong>g modelC ˙v = I − g(v − v r ) + p(v − v t ) 2 + − u if v = v peak , then˙u = a(b(v − v r ) − u) v ← c, u ← u + dwhere x + = x when x > 0 and x + = 0 when otherwise. This model better fitsthe upstroke of the action potential.14. Explore numerically the model (8.3, 8.4) with a nonl<strong>in</strong>ear after-spike reset v ←f(u), u ← g(u), where f and g are some functions.15. [M.S.] Analyze the generalization of the system (8.3, 8.4)where e is another parameter.˙v = I + v 2 + evu − u if v = 1, then˙u = a(bv − u) v ← c, u ← u + d16. [M.S.] Analyze the generalization of the follow<strong>in</strong>g system, related to the exponential<strong>in</strong>tegrate-and-fire modelwhere k is another parameter.˙v = I − v + ke v − u if v = 1, then˙u = a(bv − u) v ← c, u ← u + d


328 Simple Models17. [M.S.] Analyze the follow<strong>in</strong>g system˙v = I − v + kv 2 + − u if v = 1, then˙u = a(bv − u) v ← c, u ← u + dwhere v + = v when v > 0 and v + = 0 otherwise.18. [M.S.] F<strong>in</strong>d an analytical solution to the system (8.3, 8.4) with time-dependent<strong>in</strong>put, I = I(t).19. [M.S.] Determ<strong>in</strong>e the complete bifurcation diagram of the system (8.3, 8.4).


Simple Models 329Figure 8.29: An alternative classification of neocortical <strong>in</strong>hibitory <strong>in</strong>terneurons (modifiedfrom Markram et al. 2004). Five major classes: non-accommodat<strong>in</strong>g (NAC),accommodat<strong>in</strong>g (AC), stutter<strong>in</strong>g (STUT), burst<strong>in</strong>g (BST), and irregular spik<strong>in</strong>g (IS).Most classes conta<strong>in</strong> subclasses: delay (d), classic (c), and burst (b). For burst<strong>in</strong>g<strong>in</strong>terneurons, the three types are repetitive (r), <strong>in</strong>itial (i), and transient (t). Subclassd-IS is not provided <strong>in</strong> the orig<strong>in</strong>al picture by Markram et al. (2004).


330 Simple Modelsa35 mV350 msc-NAC b-NAC d-NACbc-AC b-AC d-ACc30 mV1 secc-STUT b-STUT d-STUTdi-BST r-BST t-BSTec-IS b-IS d-ISFigure 8.30: Simulations of the simple model with various parameters can reproduceall fir<strong>in</strong>g patterns of neocortical <strong>in</strong>hibitory <strong>in</strong>terneurons <strong>in</strong> Fig. 8.29.


Simple Models 331cat TC neuronsimple modeltonic mode burst mode tonic mode burst mode200 ms-60 mV-80 mVrecovery variable, u1000-100-200attractiondoma<strong>in</strong> ofrest<strong>in</strong>g stateu-nullcl<strong>in</strong>ev-nullcl<strong>in</strong>enodehyperpolarizedstatesaddletonic moderebound burst modespike cutoffreset-80 -60 -40 -20 0 20 40membrane potential, v (mV)Figure 8.31: Comparison of <strong>in</strong> vitro record<strong>in</strong>gs of a thalamocortical (TC) cell of catdorsal lateral geniculate nucleus with simulations of the simple model 200 ˙v = 1.6(v +60)(v + 50) − u + I, ˙u = 0.01{b(v + 65) − u}, b = 15 if v ≤ −65 and b = 0 otherwise.When v ≥ 35 + 0.1u, then v ← −60 − 0.1u, u ← u + 10. Injected current pulsesare <strong>in</strong> 50 pA <strong>in</strong>crements. In burst mode, the cell was hyperpolarized to −80 mV priorto <strong>in</strong>ject<strong>in</strong>g a depolariz<strong>in</strong>g pulse of current (data provided by C. L. Cox and S. M.Sherman)


332 Simple Modelsrat RTN neuronsimple modeltonic mode burst mode tonic mode burst mode200 ms-65 mV-85 mVFigure 8.32: Comparison of <strong>in</strong> vitro record<strong>in</strong>gs of a reticular thalamic nucleus (RTN)neuron of a rat with simulations of the simple model 40 ˙v = 0.25(v +65)(v +45)−u+I,˙u = 0.015{b(v + 65) − u}, b = 10 if v ≤ −65 and b = 2 otherwise. When v ≥ 0 (spikecutoff), then v ← −55, u ← u + 50. Injected current pulses are 50, 70, 110 pA. Inburst mode, the cell was hyperpolarized to −80 mV prior to <strong>in</strong>ject<strong>in</strong>g a depolariz<strong>in</strong>gpulse of current (data provided S.H. Lee and C.L. Cox).


Simple Models 333cat thalamic <strong>in</strong>terneuronsimple model20 mV50 ms-60 mVFigure 8.33: Comparison of <strong>in</strong> vitro record<strong>in</strong>gs of dorsal lateral geniculate nucleus<strong>in</strong>terneuron of a cat with simulations of the simple model 20 ˙v = 0.5(v + 60)(v +50) − u + I, ˙u = 0.05{7(v + 60) − u}. When v ≥ 20 − 0.08u (spike cutoff), thenv ← −65 + 0.08u, u ← u + 50. Injected current pulses are 50, 100, 200, 250 pA (dataprovided by C. L. Cox and S. M. Sherman)


334 Simple ModelsFigure 8.34: Classification of hippocampal CA1 pyramidal neurons. A-E, <strong>in</strong> vitrorecord<strong>in</strong>gs from five different pyramidal neurons arranged accord<strong>in</strong>g to a gradient of<strong>in</strong>creas<strong>in</strong>g propensity to burst. The neurons were stimulated with current pulses of 200ms duration and amplitude 50 pA and 100 pA (a), or brief (3-5 ms) superthresholdpulses (b). Non-burster (NB) neuron fires tonic spikes <strong>in</strong> response to long pulses anda s<strong>in</strong>gle spike <strong>in</strong> response to brief pulses. High-threshold burster (HTB) fires burstsonly <strong>in</strong> response to strong long pulses, and s<strong>in</strong>gle spikes <strong>in</strong> response to weak or briefpulses. Grade I low-threshold burster (LTB I) generates bursts <strong>in</strong> response to longpulses of current, but s<strong>in</strong>gle spikes <strong>in</strong> response to brief pulses. Grade II LTB (LTB II)fires bursts <strong>in</strong> response to both long (a) and brief (b) current pulses. Grade III LTB(LTB III), <strong>in</strong> addition to fir<strong>in</strong>g bursts <strong>in</strong> response to long and brief pulses of current(not shown) also fires spontaneous rhythmic bursts, shown <strong>in</strong> contracted and expandedtime scales (reproduced from Su et al. (2001) with permission).


Simple Models 335A B C D ENB HTB LTB I LTB II LTB IIIa +40 mV a a aI=100 pA-60mV500 msI=50 pAb b b b50 ms100 ms10 mV10 msrecovery variable, u2001000restv-nullcl<strong>in</strong>eresetu-nullcl<strong>in</strong>e-60 -40 -20 0 20 40I=50 pA I=50 pA I=50 pA I=50 pA0 0 0-60 -40 -20 0 20 40 -60 -40 -20 0 20 40 -60 -40 -20 0 20 40membrane potential, v (mV)Figure 8.35: Simulations of hippocampal CA1 pyramidal neurons (compare withFig. 8.34) us<strong>in</strong>g simple model 50 ˙v = 0.5(v+60)(v+45)−u+I, ˙u = 0.02{0.5(v+60)−u}.When v ≥ 40 (spike cutoff), then v ← c and u ← u + d. Here c = −50, −45, −40, −35mV and d = 50, 50, 55, 60 for A-D, respectively. Parameters <strong>in</strong> E are the same as <strong>in</strong> D,but I = 33 pA.500 ms10 mVup-state-49 mVdown-state-63 mVFigure 8.36: Neostriatal sp<strong>in</strong>y neurons have two-state behavior <strong>in</strong> vivo (data providedby Charles Wilson).


336 Simple Modelssp<strong>in</strong>y neuronlatencysimple modelI=640 pAI=520 pAI=430 pAI=420 pA+40 mVI=400 pArecovery variable, u recovery variable, u500-50050 mV100 ms500-5000rest<strong>in</strong>g0v-nullcl<strong>in</strong>eI=640 pAI=400 pAu-nullcl<strong>in</strong>eresetghost ofsaddle-noderesetI>0I=0spikespike-100 -80 -60 -40 -20 0membrane potential, v (mV)Figure 8.37: Comparison of <strong>in</strong> vitro record<strong>in</strong>gs of a neostriatal sp<strong>in</strong>y projection neuronof a rat with simulations of the simple model 50 ˙v = (v + 80)(v + 25) − u + I, ˙u =0.01{−20(v + 80) − u}, if v ≥ 40, then v ← −55, u ← u + 150 (<strong>in</strong> vitro data werek<strong>in</strong>dly provided by C. Wilson).


Simple Models 337800resetI=600 pA4000I=500 pAI=500 pA20 mV25 msreset+10 mV8004000spik<strong>in</strong>g limit cycleI=470 pAI=400 pAI=200 pAoscillationsrest-50 mVrecovery variable, u8004000restu-nullcl<strong>in</strong>ev-nullcl<strong>in</strong>espikeI=400 pA-60 -40 -20 0membrane potential, v (mV)Figure 8.38: Comparison of <strong>in</strong> vitro record<strong>in</strong>gs of rat’s bra<strong>in</strong>stem mes V neuron (fromFig. 7.3) with simulations of the simple model 25 ˙v = (v + 50)(v + 30) − u + I, ˙u =0.5{U(v + 50) − u}, with cubic slow nullcl<strong>in</strong>e U(x) = 25x + 0.009x 3 . If v ≥ 10, thenv ← −40.


338 Simple Models(a)stellate cellof entorh<strong>in</strong>alcortexsimple model-60 mVsag15 mV1 sec-500 pA 100 pA 165 pA 200 pA(b)stellate cell of entorh<strong>in</strong>al cortexsimple modelI=173 pAI=170 pAI=167 pAI=165 pA(c)recovery variable, u20016012080separatrixspikespikesI=165 pA limit cycleI=173 pAattractor-60 -55 -50 -45 -40 -35-60 -55 -50 -45 -40 -35membrane potential, V (mV)membrane potential, V (mV)Figure 8.39: Comparison of <strong>in</strong> vitro record<strong>in</strong>gs of stellate neurons of rat’s entorh<strong>in</strong>alcortex with simulations of the simple model 200 ˙v = 0.75(v + 60)(v + 45) − u + I,˙u = 0.01{15(v + 60) − u}, if v ≥ 30, then v ← −50. (a) Responses to steps of dccurrent.(b) Subthreshold oscillations and occasional spikes at various levels of <strong>in</strong>jecteddc-current. (c) Phase portraits correspond<strong>in</strong>g to two levels of <strong>in</strong>jected dc-current. Weaknoise was added to simulations to unmask subthreshold oscillations. (Data were k<strong>in</strong>dlyprovided by Brian Burton and John A. White. All record<strong>in</strong>gs are from the same neuronexcept steps of −500 pA and 200 pA were recorded from a different neuron. Spikes arecut at 0 mV.)


Simple Models 339rat's mitral cell (<strong>in</strong> vitro)40 mV100 mssimple modelI=35pAI=20 pAup-stateI=15pAI= 10pAsomadendriteFigure 8.40: Comparison of <strong>in</strong> vitro record<strong>in</strong>gs of mitral neurons of rat’s olfactory bulbwith simulations of the simple two-compartment model. Soma: 40 ˙v = (v+55)(v+50)+0.5(v d −v)−u+I, ˙u = 0.4{(U(v)−u} with U(v) = 0 when v < v b and U(v) = 20(v−v b )when v ≥ v b = −48 mV. If v ≥ 35, then v ← −50. Passive dendrite (dotted curve):˙v d = 0.0125(v − v d ). Weak noise was added to simulations to unmask subthresholdoscillations <strong>in</strong> the up-state. The membrane potential of the neuron is held at −75mV by <strong>in</strong>ject<strong>in</strong>g a strong negative current, and then stimulated with steps of positivecurrent. (Data were k<strong>in</strong>dly provided by Philip Heyward.)


340 Simple Modelsrat's mitral cell (<strong>in</strong> vitro)simple modelrat's mitral cell (<strong>in</strong> vitro)simple model20 mVup-state (-46 mV)down-state (-55 mV)spikes cut at -20 mV100 msrecovery variable, u recovery variable, u20015010050020015010050I=0 pA0u-nullcl<strong>in</strong>edown-state(node)I = 7 pAv-nullcl<strong>in</strong>esaddleattraction doma<strong>in</strong>of the up-stateup-stateup-state(focus)thresholdsaddlethresholdstimulationsaddlespikespike-60 -55 -50 -45 -40 -35membrane potential, v (mV)Figure 8.41: Voltage responses of a rat’s mitral cell and a simple model from Fig. 8.40at two different values of the hold<strong>in</strong>g current. Right: Phase portraits of somatic compartmentsshow coexistence of stable node (down-state) and stable focus (up-state)equilibria. Spikes are emitted only from the up-state.Figure 8.42: Louis Lapicque — the <strong>in</strong>ventor of the<strong>in</strong>tegrate-and-fire neuron.


Chapter 9Burst<strong>in</strong>gA neuron can fire a s<strong>in</strong>gle spike or a stereotypical burst of spikes, depend<strong>in</strong>g on thenature of stimulation and the <strong>in</strong>tr<strong>in</strong>sic neuronal properties. Typically, burst<strong>in</strong>g occursdue to the <strong>in</strong>terplay of fast currents responsible for spik<strong>in</strong>g activity, and slow currentsthat modulate the activity. In this chapter we study this <strong>in</strong>terplay <strong>in</strong> detail.To understand the geometry of burst<strong>in</strong>g, it is customary to assume that the fast andslow currents have drastically different time scales. In this case we can dissect a burster,i.e., freeze its slow currents and use them as parameters that control the fast spik<strong>in</strong>gsubsystem. Dur<strong>in</strong>g burst<strong>in</strong>g, the slow parameters drive the fast subsystem throughbifurcations of equilibria and limit cycles. We provide a topological classification ofbursters based on these bifurcations, and show that different topological types havedifferent neuro-computational properties.9.1 ElectrophysiologyMany spik<strong>in</strong>g neurons can exhibit burst<strong>in</strong>g activity if manipulated, e.g., pharmacologically.In Fig. 9.1 we depict a few well-known examples of neurons that burst undernatural conditions without any manipulation. Some require an <strong>in</strong>jected dc-current tobias the membrane potential, others do not. One can only be amazed by the diversityof burst<strong>in</strong>g patterns and time scales. In this chapter we consider electrophysiologicaland bifurcation mechanisms responsible for the generation of these patterns.Is zebra a black animal with white stripes or a white animal with black stripes? Thisseem<strong>in</strong>gly silly question is pert<strong>in</strong>ent to every burst<strong>in</strong>g pattern: Does burst<strong>in</strong>g activitycorrespond to an <strong>in</strong>f<strong>in</strong>ite period of quiescence <strong>in</strong>terrupted by groups of spikes or doesit correspond to an <strong>in</strong>f<strong>in</strong>ite spike tra<strong>in</strong> <strong>in</strong>terrupted by short periods of quiescence?Biologists are mostly concerned with the question of what makes the neuron fire thefirst spike <strong>in</strong> a burst and what keeps it <strong>in</strong> the spik<strong>in</strong>g regime afterwards. The questionof why the spik<strong>in</strong>g stops is often forgotten. It turns out that to fully understand theionic mechanism of burst<strong>in</strong>g, we need to concentrate on the second question, i.e., weneed to treat burst<strong>in</strong>g as an <strong>in</strong>f<strong>in</strong>ite spike tra<strong>in</strong> that is chopped <strong>in</strong>to short bursts by aslow (resonant) current that builds up dur<strong>in</strong>g the spik<strong>in</strong>g phase and recovers dur<strong>in</strong>g the341


342 Burst<strong>in</strong>g(a) cortical chatter<strong>in</strong>g neuron(b) cortical <strong>in</strong>tr<strong>in</strong>sically burst<strong>in</strong>g neuron(c) cortical <strong>in</strong>tr<strong>in</strong>sically burst<strong>in</strong>g neuron(d) thalamic reticular neuron500 ms50 ms(e) thalamocortical relay neuron100 ms(f) hyppocampal pyramical neuron20 mV-66 mV(g) respiratory neuron <strong>in</strong> pre-Botz<strong>in</strong>ger complex(h) rodent trigem<strong>in</strong>al neuronFigure 9.1: Examples of <strong>in</strong>tr<strong>in</strong>sic bursters. (a) and (b) cat primary visual corticalneurons (modified from Nowak et al. 2003). (c) cortical neuron <strong>in</strong> anesthetized cat(modified from Timofeev et al. 2000). (d) thalamic reticular (RE) neuron (modifiedfrom Steriade 2003). (e) Cat thalamocortical relay neuron (modified from McCormickand Pape 1990). (f) CA1 pyramidal neuron exhibit<strong>in</strong>g grade II low-threshold burst<strong>in</strong>gpattern (modified from Su et al. 2001). (g) respiratory neuron <strong>in</strong> the pre-Botz<strong>in</strong>gercomplex (modified from Butera et al. 1999). (h) Trigem<strong>in</strong>al <strong>in</strong>terneuron from ratbra<strong>in</strong>stem (modified from Del Negro et al. 1998).4 sec


Burst<strong>in</strong>g 343Figure 9.2: Is burst<strong>in</strong>g a spik<strong>in</strong>g state <strong>in</strong>terrupted by periods of quiescence or is it aquiescent state <strong>in</strong>terrupted by groups of spikes?V(t)20 mV50 msI(t)Figure 9.3: Forced burst<strong>in</strong>g <strong>in</strong> the I Na,p +I K -model with parameters as <strong>in</strong> Fig. 4.1a andtime-dependent <strong>in</strong>jected current I(t).quiescent phase. Before proceed<strong>in</strong>g to a general case, let us consider a simple example.9.1.1 Example: The I Na,p +I K +I K(M) -modelAny model neuron capable of spik<strong>in</strong>g can also burst, as e.g., the I Na,p +I K -model <strong>in</strong>Fig. 9.3. However, this example is not <strong>in</strong>terest<strong>in</strong>g because the neuron is forced to burstby the time-dependent <strong>in</strong>put I(t).In contrast, a modification of the I Na,p +I K -model <strong>in</strong> Fig. 9.4 fires a burst of spikes<strong>in</strong> response to a brief pulse of current. The first spike <strong>in</strong> the burst is caused by thestimulation, but the subsequent spikes are generated autonomously due to the <strong>in</strong>tr<strong>in</strong>sicproperties of the neuron, and they outlast the stimulation. Such a burst is stereotypicaland fairly <strong>in</strong>dependent of the amplitude or the duration of the pulse that triggered it.To make the I Na,p +I K -model burst, we took parameters as <strong>in</strong> Fig. 6.7a, so that


344 Burst<strong>in</strong>gmembranepotential (mV)-20-40-60I(t)I(t)slow K + activationgate, n slow0.060.040.0200 20 40time (ms)a00 50 100 150 200time (ms)bFigure 9.4: Intr<strong>in</strong>sic burst<strong>in</strong>g <strong>in</strong> the I Na,p +I K +I K(M) -model (7.1) consist<strong>in</strong>g of theI Na,p +I K -model with parameters as <strong>in</strong> Fig. 4.1a and fast K + current (g K = 9, τ(V ) =0.152) and a slow K + current with g slow = 5, V 1/2 = −20 mV, k = 5 mV andτ slow (V ) = 20 ms. (a) Burst excitability when I = 0. (b) Periodic burst<strong>in</strong>g whenI = 5.there is a coexistence of the rest<strong>in</strong>g and spik<strong>in</strong>g states. The brief pulse of currentexcites the neuron, i.e., moves its state <strong>in</strong>to the attraction doma<strong>in</strong> of the spik<strong>in</strong>g limitcycle and <strong>in</strong>itiates periodic activity. Without any other modification, the model wouldproduce an <strong>in</strong>f<strong>in</strong>ite spike tra<strong>in</strong>. To stop the tra<strong>in</strong>, we added a slower high-thresholdpersistent K + current similar to I K(M) that provides a negative feedback. This M-current is deactivated at rest. However, dur<strong>in</strong>g the active (spik<strong>in</strong>g) phase, the currentslowly activates, as <strong>in</strong>dicated by the slow build-up of its gat<strong>in</strong>g variable n slow <strong>in</strong> thefigure. The neuron becomes less and less excitable, and eventually cannot susta<strong>in</strong>spik<strong>in</strong>g activity. If, <strong>in</strong>stead of a pulse of current, a constant current is applied, theneuron can burst periodically, as <strong>in</strong> Fig. 9.4b.This model presents only one of many possible examples of bursters, which we study<strong>in</strong> this chapter. However, it illustrates a number of important issues common to allbursters. For <strong>in</strong>stance, <strong>in</strong> contrast to the forced burst<strong>in</strong>g <strong>in</strong> Fig. 9.3, this burst<strong>in</strong>g is<strong>in</strong>tr<strong>in</strong>sic or autonomous. This stereotypical burst<strong>in</strong>g pattern results from the <strong>in</strong>tr<strong>in</strong>sicvoltage-sensitive currents, and not from a time-dependent <strong>in</strong>put. The behavior <strong>in</strong>Fig. 9.4a is called burst excitability to emphasize that the model is an excitable systemwith the exception that superthreshold stimulation elicits a burst of spikes <strong>in</strong>stead of as<strong>in</strong>gle spike. Hippocampal pyramidal neurons of the type “grade III bursters” depicted<strong>in</strong> Fig. 8.34Eb exhibit burst excitability.Biologists sometimes refer to the burst<strong>in</strong>g <strong>in</strong> Fig. 9.4b as be<strong>in</strong>g conditional, becauserepetitive burst<strong>in</strong>g occurs when a certa<strong>in</strong> condition is satisfied, e.g., positive I is <strong>in</strong>-


Burst<strong>in</strong>g 345<strong>in</strong>terburst periodquiescentperiodactivephase<strong>in</strong>terspike(<strong>in</strong>traburst)periodduty cycle =active phase<strong>in</strong>terburst periodV(t)Figure 9.5: Basic characteristics of burst<strong>in</strong>g dynamics.jected. From a mathematical po<strong>in</strong>t of view, every burster is conditional, s<strong>in</strong>ce it existsfor some values of the parameters but not others.9.1.2 Fast-Slow DynamicsIn general, every burst<strong>in</strong>g pattern consists of oscillations with two time scales: fastspik<strong>in</strong>g oscillation with<strong>in</strong> a s<strong>in</strong>gle burst (<strong>in</strong>traburst oscillation, or spik<strong>in</strong>g), modulatedby a slow oscillation between the bursts (<strong>in</strong>terburst oscillation); see Fig. 9.5. Typically,though not necessarily (see exercises at the end of this chapter), two time scales resultfrom two <strong>in</strong>teract<strong>in</strong>g processes <strong>in</strong>volv<strong>in</strong>g fast and slow currents. For example, thespik<strong>in</strong>g <strong>in</strong> Fig. 9.4 is generated by the fast I Na,p +I K -subsystem and modulated by theslow I K(M) -subsystem.There are two questions associated with each burst<strong>in</strong>g pattern:• What <strong>in</strong>itiates susta<strong>in</strong>ed spik<strong>in</strong>g dur<strong>in</strong>g the burst?• What term<strong>in</strong>ates susta<strong>in</strong>ed spik<strong>in</strong>g temporarily and ends the burst?The answer to the first question is relatively simple: Repetitive spik<strong>in</strong>g is <strong>in</strong>itiated andsusta<strong>in</strong>ed by the positive <strong>in</strong>jected current I, or some other source of persistent <strong>in</strong>wardcurrent that causes the neuron to fire (most biologists are <strong>in</strong>terested <strong>in</strong> identify<strong>in</strong>gthis source). Surpris<strong>in</strong>gly, the second question is the most important for build<strong>in</strong>g amodel of burst<strong>in</strong>g. While the neuron fires, relatively slow processes somehow make itnon-excitable and eventually term<strong>in</strong>ate the fir<strong>in</strong>g. Such slow processes result <strong>in</strong> a slowbuildup of an outward current or <strong>in</strong> a slow decrease of an <strong>in</strong>ward current needed tosusta<strong>in</strong> the spik<strong>in</strong>g. Dur<strong>in</strong>g the quiescent phase, the neuron slowly recovers and rega<strong>in</strong>sthe ability to generate action potentials aga<strong>in</strong>.Let us discuss possible ionic mechanisms responsible for the term<strong>in</strong>ation of spik<strong>in</strong>gwith<strong>in</strong> a burst. Suppose we are given a neuronal model that is capable of susta<strong>in</strong>edspik<strong>in</strong>g activity, at least when a positive I is <strong>in</strong>jected. To transform an <strong>in</strong>f<strong>in</strong>ite spiketra<strong>in</strong> <strong>in</strong>to a f<strong>in</strong>ite burst of spikes, it suffices to add a slow resonant current or gat<strong>in</strong>g


346 Burst<strong>in</strong>gvoltage-gatedCa2+-gated<strong>in</strong>activation of<strong>in</strong>ward currentrest<strong>in</strong>gde<strong>in</strong>activationof <strong>in</strong>wardcurrentrepolarizationdepolarization<strong>in</strong>activationof <strong>in</strong>wardcurrentspik<strong>in</strong>grest<strong>in</strong>gCa 2+buffer<strong>in</strong>gde<strong>in</strong>activationof <strong>in</strong>wardcurrentrepolarizationdepolarization<strong>in</strong>activationof <strong>in</strong>wardcurrentCa 2+ <strong>in</strong>fluxand buildupspik<strong>in</strong>gactivation ofoutward currentrest<strong>in</strong>gdeactivationof outwardcurrentrepolarizationdepolarizationactivationof outwardcurrentspik<strong>in</strong>grest<strong>in</strong>gCa 2+buffer<strong>in</strong>gdeactivationof outwardcurrentrepolarizationdepolarizationactivationof outwardcurrentCa 2+ <strong>in</strong>fluxand buildupspik<strong>in</strong>gspik<strong>in</strong>grest<strong>in</strong>gbuildup of resonant gate or [Ca2+]<strong>in</strong>(activation of outward current)(<strong>in</strong>activation of <strong>in</strong>ward current)recovery of resonant gate or [Ca2+]<strong>in</strong>(deactivation of outward current)(de<strong>in</strong>activation of <strong>in</strong>ward current)Figure 9.6: Four major classes of burst<strong>in</strong>g models are def<strong>in</strong>ed by the slow resonantgat<strong>in</strong>g variable that modulates spik<strong>in</strong>g activity.variable (see Sect. 5.1.1) that modulates the spik<strong>in</strong>g via a slow negative feedback. Theresonant gat<strong>in</strong>g variable can describe <strong>in</strong>activation of an <strong>in</strong>ward current or activation ofan outward current, both voltage- or Ca 2+ -dependent (see Fig. 5.17). Hence, there arefour major classes of burst<strong>in</strong>g models, summarized <strong>in</strong> Fig. 9.6:• Voltage-gated <strong>in</strong>activation of an <strong>in</strong>ward current, e.g., slow <strong>in</strong>activation of persistentNa + current or <strong>in</strong>activation of Ca 2+ transient T-current, or <strong>in</strong>activation ofthe h-current (most biologists refer to this as activation of the h-current by hyperpolarization).Repetitive spik<strong>in</strong>g slowly <strong>in</strong>activates (turns off) the <strong>in</strong>ward currentand makes the neuron less excitable and unable to susta<strong>in</strong> spik<strong>in</strong>g activity. Aftera while, the spik<strong>in</strong>g stops and the membrane potential repolarizes. The <strong>in</strong>wardcurrent slowly de-<strong>in</strong>activates (turns on) and depolarizes the membrane potential,


Burst<strong>in</strong>g 347possibly result<strong>in</strong>g <strong>in</strong> a new burst.• Voltage-gated activation of an outward current, e.g., slow activation of persistentK + current, such as M-current. Repetitive spik<strong>in</strong>g slowly activates the outwardcurrent, which eventually term<strong>in</strong>ates the spik<strong>in</strong>g activity. While at rest, theoutward current slowly deactivates (turns off) and unmasks <strong>in</strong>ward currents thatcan depolarize the membrane potential, possibly <strong>in</strong>itiat<strong>in</strong>g another burst.• Ca 2+ -gated <strong>in</strong>activation of an <strong>in</strong>ward current, e.g., slow <strong>in</strong>activation of highthresholdCa 2+ -currents I Ca(L) or I Ca(N) . Calcium entry dur<strong>in</strong>g repetitive spik<strong>in</strong>gleads to its <strong>in</strong>tracellular accumulation and slow <strong>in</strong>activation of Ca 2+ -channelsthat provide an <strong>in</strong>ward current needed for repetitive spik<strong>in</strong>g. As a result, theneuron cannot susta<strong>in</strong> spik<strong>in</strong>g activity and becomes quiescent. Dur<strong>in</strong>g this period,<strong>in</strong>tracellular Ca 2+ ions are removed, Ca 2+ channels are de-<strong>in</strong>activated, andthe neuron is primed to start a new burst.• Ca 2+ -gated activation of an outward current, e.g., slow activation of Ca 2+ -dependentK + -current I AHP . Calcium entry and buildup dur<strong>in</strong>g repetitive spik<strong>in</strong>gslowly activates the outward current and makes the neuron less and less excitable.When the spik<strong>in</strong>g stops, <strong>in</strong>tracellular Ca 2+ ions are removed, Ca 2+ -gatedoutward current deactivates (turns off), the neuron is no longer hyperpolarizedand ready to fire a new burst of spikes.In addition, the slow process may <strong>in</strong>clude Na + -, K + -, or Cl − -gated currents, suchas the “slack and slick” family of Na + -gated K + currents, or slow change of ionicconcentrations <strong>in</strong> the vic<strong>in</strong>ity of the cell membrane (so called Hodgk<strong>in</strong>-Frankenhaeuserlayer), which leads to slow change of the Nernst potential for ionic species. We do notelaborate these cases <strong>in</strong> the book.Notice that <strong>in</strong> some cases, the slow process modulates fast currents responsible forspik<strong>in</strong>g, while <strong>in</strong> other cases it produces an <strong>in</strong>dependent slow current that impedesspik<strong>in</strong>g. In any case, the slow process is directly responsible for the term<strong>in</strong>ation ofcont<strong>in</strong>uous spik<strong>in</strong>g, and <strong>in</strong>directly for its <strong>in</strong>itiation and ma<strong>in</strong>tenance.The four mechanisms <strong>in</strong> Fig. 9.6 and their comb<strong>in</strong>ations are ubiquitous <strong>in</strong> neurons,as we summarize <strong>in</strong> Fig. 9.7. However, there could be other, less obvious burst<strong>in</strong>gmechanisms. In Ex. 8–10 we provide examples of bursters hav<strong>in</strong>g slowly activat<strong>in</strong>gpersistent <strong>in</strong>ward current, such as I Na,p . These surpris<strong>in</strong>g examples show that buildupof the <strong>in</strong>ward current (or any other amplify<strong>in</strong>g gate) can also be responsible for theterm<strong>in</strong>ation of the active phase and for the repolarization of the membrane potential.To understand these mechanisms, one needs to study the geometry of burst<strong>in</strong>g.9.1.3 M<strong>in</strong>imal modelsLet us follow the ideas presented <strong>in</strong> Sect. 5.1 and determ<strong>in</strong>e m<strong>in</strong>imal models for burst<strong>in</strong>g.That is, we are <strong>in</strong>terested <strong>in</strong> classification of all fast-slow electrophysiologicalmodels that can exhibit susta<strong>in</strong>ed burst<strong>in</strong>g activity, as <strong>in</strong> Fig. 9.4b, at least for some


348 Burst<strong>in</strong>gslow dynamicsneuronvoltage-gatedactivationof outward<strong>in</strong>activationof <strong>in</strong>wardactivationof outwardCa2+-gated<strong>in</strong>activationof <strong>in</strong>wardreferencesneocortical chatter<strong>in</strong>gneuronsI K(M) I KslowWang (1999)pre-Botz<strong>in</strong>ger complex(respiratory rhythm)I KslowI NaslowButera et al. (1999)thalamic relay neurons I Ca(T) I hHuguenard andMcCormick (1992)thalamic reticularneuronshippocampal CA3neuronsI Ca(T) Destexhe et al. (1994)I AHP Traub et al. (1991)subiculum burst<strong>in</strong>gneuronsI AHP Stanford et al. (1998)midbra<strong>in</strong> dopam<strong>in</strong>ergicneuronsI K(Ca) ICa(L) Am<strong>in</strong>i et al. (1999)anterior burst<strong>in</strong>g (AB)neuron <strong>in</strong> lobsterstomatogastric ganglionIK(Ca)I Ca(L)Harris-Warrick andFlamm (1987)Aplysia abdom<strong>in</strong>alganglion R15 neuronI Ca(L)Canavier et al. (1991)Figure 9.7: Slow dynamics <strong>in</strong> burst<strong>in</strong>g neurons.values of parameters. A burst<strong>in</strong>g model is m<strong>in</strong>imal if removal of any current or gat<strong>in</strong>gvariable elim<strong>in</strong>ates the ability to burst.One way to build a fast-slow m<strong>in</strong>imal model for burst<strong>in</strong>g is to take a m<strong>in</strong>imal modelfor spik<strong>in</strong>g, which consists of an amplify<strong>in</strong>g and a resonant gate, see Fig. 5.17, and addanother slow resonant gate. S<strong>in</strong>ce there are many m<strong>in</strong>imal spik<strong>in</strong>g models <strong>in</strong> Fig. 5.17and four choices of slow resonant gates <strong>in</strong> Fig. 9.6, there are quite a few comb<strong>in</strong>ations,which fill out the squares <strong>in</strong> Fig. 9.8. We present only a few reasonable models <strong>in</strong> thefigure and ask the reader to fill <strong>in</strong> the blanks. Complet<strong>in</strong>g the table is an excellent testof one’s knowledge and understand<strong>in</strong>g of how different currents <strong>in</strong>teract to producenon-trivial fir<strong>in</strong>g patterns.Some of the m<strong>in</strong>imal models for burst<strong>in</strong>g might seem too bizarre at first glance.Yet the table <strong>in</strong> Fig. 9.8, upon completion, might prove to be a valuable tool thatcould allow experimenters to formulate various ionic hypotheses. For example, if oneuses pharmacological agents, e.g., TEA or Ba 2+ , to block Ca 2+ -gated K + channelsand show that burst<strong>in</strong>g persists, then the possible electrophysiological mechanisms of


Burst<strong>in</strong>g 349voltage-gatedCa2+-gated<strong>in</strong>activation of <strong>in</strong>ward currentamplify<strong>in</strong>g gateCa 2+ -gatedvoltage-gatedactivationof <strong>in</strong>wardcurrent<strong>in</strong>activationof outwardcurrentactivationof <strong>in</strong>wardcurrent<strong>in</strong>activationof outwardcurrentresonant gatevoltage-gatedCa 2+ -gatedactivation <strong>in</strong>activationof outward of <strong>in</strong>wardcurrent current<strong>in</strong>activationof <strong>in</strong>wardcurrentINa,t+IhINa,t(fast and slow)INa,t slow +IKINa,p+IK+Ihactivationof outwardcurrentamplify<strong>in</strong>g gateCa 2+ -gatedvoltage-gatedactivationof <strong>in</strong>wardcurrent<strong>in</strong>activationof outwardcurrentactivationof <strong>in</strong>wardcurrent<strong>in</strong>activationof outwardcurrentresonant gatevoltage-gatedCa 2+ -gatedactivation <strong>in</strong>activationof outward of <strong>in</strong>wardcurrent current<strong>in</strong>activationof <strong>in</strong>wardcurrentICa(N)ICa(L)+IKactivationof outwardcurrentactivation of outward currentamplify<strong>in</strong>g gateCa 2+ -gatedvoltage-gatedactivationof <strong>in</strong>wardcurrent<strong>in</strong>activationof outwardcurrentactivationof <strong>in</strong>wardcurrent<strong>in</strong>activationof outwardcurrentresonant gatevoltage-gatedCa 2+ -gatedactivation <strong>in</strong>activationof outward of <strong>in</strong>wardcurrent current<strong>in</strong>activationof <strong>in</strong>wardcurrentINa,t+IK(M)INa,p+IK+IK(M)activationof outwardcurrentamplify<strong>in</strong>g gateCa 2+ -gatedvoltage-gatedactivationof <strong>in</strong>wardcurrent<strong>in</strong>activationof outwardcurrentactivationof <strong>in</strong>wardcurrent<strong>in</strong>activationof outwardcurrentresonant gatevoltage-gatedCa 2+ -gatedactivation <strong>in</strong>activationof outward of <strong>in</strong>wardcurrent current<strong>in</strong>activationof <strong>in</strong>wardcurrentICa(T)+IAHPICa+IK+IAHPactivationof outwardcurrentFigure 9.8: Some m<strong>in</strong>imal models for burst<strong>in</strong>g.burst<strong>in</strong>g are conf<strong>in</strong>ed to the left column <strong>in</strong> Fig. 9.8. M<strong>in</strong>imal models <strong>in</strong> this columnwould provide testable hypotheses on the ionic basis of burst<strong>in</strong>g, and they could guidenovel experiments. If block abolishes burst<strong>in</strong>g, we cannot conclude that the blockedcurrent drives the burst<strong>in</strong>g — it may merely be necessary for provid<strong>in</strong>g backgroundstimulation.Notice that the I Na,tslow + I K -model and the I Na,t + I K(M) -model <strong>in</strong> the figure (seethe shaded rectangles) consist of the same gat<strong>in</strong>g variables: Na + activation gate m,<strong>in</strong>activation gate h, and K + activation gate n. Both models are equivalent to theHodgk<strong>in</strong>-Huxley model, the only difference be<strong>in</strong>g the choice of the slow gate. Thus, <strong>in</strong>contrast to the common biophysical folklore, the Hodgk<strong>in</strong>-Huxley model is a m<strong>in</strong>imalmodel for burst<strong>in</strong>g, and there are two fundamentally different ways <strong>in</strong> which one canmake it burst without any additional currents, as we show <strong>in</strong> Fig. 9.9. Of course, one


350 Burst<strong>in</strong>gslow <strong>in</strong>activation of <strong>in</strong>ward current120slow activation of outward current100V(t)100V(t)806040806020400200 500 1000 1500 2000 0 500 1000 1500 2000 2500 3000 3500abFigure 9.9: Hodgk<strong>in</strong>-Huxley (1952) model with three gat<strong>in</strong>g variables is m<strong>in</strong>imal forburst<strong>in</strong>g (modified from Fig. 1.10 <strong>in</strong> Izhikevich 2001).0uncoupledhalf-centeroscillator-mutual <strong>in</strong>hibition-Figure 9.10: Central pattern generation by mutually <strong>in</strong>hibitory oscillators.may argue that the model <strong>in</strong> the figure is not Hodgk<strong>in</strong>-Huxley at all, s<strong>in</strong>ce we changedthe k<strong>in</strong>etics of some currents by an order of magnitude.Th<strong>in</strong>k<strong>in</strong>g <strong>in</strong> terms of m<strong>in</strong>imal models, we can understand what is essential forspik<strong>in</strong>g and burst<strong>in</strong>g and what is not. In addition, we can clearly see that some wellknownconductance-based models form partially-ordered set. For example, the cha<strong>in</strong>of neuronal models Morris-Lecar (I Ca +I K ) ≺ Hodgk<strong>in</strong>-Huxley (I Na,t +I K ) ≺ Butera-R<strong>in</strong>zel-Smith (I Na,t +I K +I K,slow ) is obta<strong>in</strong>ed by add<strong>in</strong>g a conductance or gat<strong>in</strong>g variableto one model to get the next one. Here, A ≺ B means A is a subsystem of B.Understand<strong>in</strong>g the ionic bases of burst<strong>in</strong>g is an important step <strong>in</strong> analysis of burst<strong>in</strong>gdynamics. However, such an understand<strong>in</strong>g may not provide sufficient <strong>in</strong>formationon why the burst<strong>in</strong>g pattern looks as it does, what the neuro-computational propertiesof the neuron are, and how they depend on the parameters of the system. Indeed, weshowed <strong>in</strong> Chap. 5 that spik<strong>in</strong>g models based on quite different ionic mechanisms canhave identical dynamics and vice versa. This is true for burst<strong>in</strong>g models as well.


Burst<strong>in</strong>g 351Spik<strong>in</strong>gBistabilityu(t)Rest<strong>in</strong>guFigure 9.11: Parameter u can control spik<strong>in</strong>gbehavior of the fast subsystem <strong>in</strong> (9.1).When u changes slowly, the model exhibitsburst<strong>in</strong>g behavior.9.1.4 Central pattern generators and half-center oscillatorsBurst<strong>in</strong>g can also appear <strong>in</strong> small circuits of coupled spik<strong>in</strong>g neurons, such as the twomutually <strong>in</strong>hibitory oscillators <strong>in</strong> Fig. 9.10, called half-center oscillators. While onecell fires, the other is <strong>in</strong>hibited, then they switch roles, and so on. Such small circuits,suggested by Brown (1911), are the build<strong>in</strong>g blocks of central pattern generators <strong>in</strong>pyloric network of the lobster stomatogastric ganglion, medic<strong>in</strong>e leech heartbeat, fictivemotor patterns and swimm<strong>in</strong>g patterns of many vertebrates and <strong>in</strong>vertebrates (Marderand Bucher 2001).What makes the oscillators <strong>in</strong> Fig. 9.10 alternate? Wang and R<strong>in</strong>zel (1992) suggestedtwo mechanisms, release and escape, which were later ref<strong>in</strong>ed to <strong>in</strong>tr<strong>in</strong>sic orsynaptic by Sk<strong>in</strong>ner et al. (1994):• <strong>in</strong>tr<strong>in</strong>sic release: The active cell stops spik<strong>in</strong>g, term<strong>in</strong>ates <strong>in</strong>hibition andallows <strong>in</strong>hibited cell to fire.• <strong>in</strong>tr<strong>in</strong>sic escape: Inhibited cell recovers, starts to fire and shuts off the activecell.• synaptic release: The <strong>in</strong>hibition weakens, e.g., due to spike frequency adaptationor short-term synaptic depression, and allows the <strong>in</strong>hibited cell to fire.• synaptic escape: Inhibited cell depolarizes above certa<strong>in</strong> threshold and startsto <strong>in</strong>hibit the active cell.All four mechanisms assume that <strong>in</strong> addition to fast variables responsible for spik<strong>in</strong>g,there are also slow adaptation variables responsible for slow<strong>in</strong>g down or term<strong>in</strong>ation ofspik<strong>in</strong>g, recovery, or synaptic depression. Thus, similarly to the m<strong>in</strong>imal models above,the circuit has at least two time scales, i.e., it is a fast-slow system.9.2 GeometryTo understand the neuro-computational properties of bursters, we need to study thegeometry of their phase portraits. In general, it is quite a difficult task. However, itcan be accomplished <strong>in</strong> the special case of fast-slow dynamics.


352 Burst<strong>in</strong>g9.2.1 Fast-slow burstersWe say that a neuron is a fast-slow burster if its behavior can be described by afast-slow system of the formẋ = f(x, u) (fast spik<strong>in</strong>g),˙u = µg(x, u) (slow modulation).(9.1)The vector x ∈ R m describes fast variables responsible for spik<strong>in</strong>g. It <strong>in</strong>cludes themembrane potential V , activation and <strong>in</strong>activation gat<strong>in</strong>g variables for fast currents,etc. The vector u ∈ R k describes relatively slow variables that modulate fast spik<strong>in</strong>g,e.g., gat<strong>in</strong>g variable of a slow K + current, <strong>in</strong>tracellular concentration of Ca 2+ ions,etc. The small parameter µ represents the ratio of time scales between spik<strong>in</strong>g andmodulation. When we analyze models, we assume that µ ≪ 1; that is, it could be assmall as we wish. The results obta<strong>in</strong>ed via such an analysis may not have any sensewhen µ is of the order 0.1 or greater.To analyze bursters, we first assume that µ = 0, so that we can consider thefast and slow systems separately. This constitutes the method of dissection of neuralburst<strong>in</strong>g pioneered by R<strong>in</strong>zel (1985). In fact, we have done this many times <strong>in</strong> theprevious chapters when we substituted m = m ∞ (V ) <strong>in</strong>to the voltage equation. Thefast subsystem can be rest<strong>in</strong>g (but excitable), bistable, or spik<strong>in</strong>g depend<strong>in</strong>g on thevalue of u; see Fig. 9.11. Burst<strong>in</strong>g occurs when u visits the spik<strong>in</strong>g and quiescentareas periodically. Many important aspects of burst<strong>in</strong>g behavior can be understoodvia phase portrait analysis of the fast subsystemẋ = f(x, u) , x ∈ R m ,treat<strong>in</strong>g u ∈ R k as a vector of slowly chang<strong>in</strong>g bifurcation parameters.We say that the burster is of the “m+k” type when the fast subsystem is m-dimensional and the slow subsystem is k-dimensional. There are some “1+1” and“2+0” bursters, see Ex. 1 — Ex. 4, though they do not correspond to any knownneuron. Most of the burst<strong>in</strong>g models <strong>in</strong> this chapter are of the “2+1” and “2+2” type.9.2.2 Phase portraitsS<strong>in</strong>ce most burst<strong>in</strong>g models are at least of the “2+1” type, their phase space is atleast three-dimensional. Analyz<strong>in</strong>g and depict<strong>in</strong>g multi-dimensional phase portraits ischalleng<strong>in</strong>g. Even understand<strong>in</strong>g the geometry of a s<strong>in</strong>gle burst<strong>in</strong>g trajectory depicted<strong>in</strong> Fig. 9.12 is difficult unless one uses a stereoscope.In Fig. 9.13 we <strong>in</strong>vestigate geometrically the I Na,p +I K +I K(M) -model, which is afast-slow burster of the “2+1” type. The naked burst<strong>in</strong>g trajectory is shown <strong>in</strong> thelower left corner. We set µ = 0 (i.e., τ slow (V ) = +∞) and slice the three-dimensionalspace by planes n slow =const, shown <strong>in</strong> the top right corner. Phase portraits of thetwo-dimensional fast subsystem with fixed n slow are shown <strong>in</strong> the middle of the figure.Notice how the limit cycle attractors and the equilibria of the fast subsystem depend on


Burst<strong>in</strong>g 35300−10−10membrane potential, V (mV)−20−30−40−50−60membrane potential, V (mV)−20−30−40−50−60−70−70−80−0.0210 0.02 fast K + 0.5activation, n0.04 0.06 0.08 0−80−0.0210 0.50.02 fast K + activation, n0.04 0.06 0.08 0slow K + activation, n slowslow K + activation, n slowFigure 9.12: Stereoscopic image of a burst<strong>in</strong>g trajectory of the I Na,p +I K +I K(M) -model<strong>in</strong> the three-dimensional phase space (V, n, n slow ) (for cross-eye view<strong>in</strong>g).the value of n slow . Glu<strong>in</strong>g the phase portraits together, we see that there is a manifoldof limit cycle attractors (shaded cyl<strong>in</strong>der) that starts when n slow < 0 and ends <strong>in</strong> asaddle homocl<strong>in</strong>ic orbit bifurcation when n slow = 0.066. There is also a locus of stableand unstable equilibria that appears via a saddle-node bifurcation when n slow = 0.0033.Once we understand the transitions from one phase portrait to another as the slowvariable changes, we can understand the geometry of the burster. Suppose µ > 0 (i.e.,τ slow (V ) = 20 ms) so that n slow can evolve accord<strong>in</strong>g to its gat<strong>in</strong>g equation.Let us start with the membrane potential at the stable equilibrium correspond<strong>in</strong>gto rest<strong>in</strong>g state. The parameters of the I Na,p +I K +I K(M) -model (see caption to Fig. 9.4)are such that slow K + M-current deactivates at rest, i.e., n slow slowly decreases, andthe burst<strong>in</strong>g trajectory slides along the bold half-parabola correspond<strong>in</strong>g to the locusof stable equilibria. After a while, the K + current becomes so small, that it cannot holdthe membrane potential at rest. This happens when n slow passes the value 0.0033, thestable equilibrium coalesces with an unstable equilibrium (saddle), and they annihilateeach other via saddle-node bifurcation. S<strong>in</strong>ce the rest<strong>in</strong>g state no longer exists (see thephase portrait at the top left of Fig. 9.13) the trajectory jumps up to the stable limitcycle correspond<strong>in</strong>g to repetitive spik<strong>in</strong>g. This jump<strong>in</strong>g corresponds to the transitionfrom rest<strong>in</strong>g to spik<strong>in</strong>g behavior.While the fast subsystem fires spikes, the K + M-current slowly activates, i.e., n slowslowly <strong>in</strong>creases. The burst<strong>in</strong>g trajectory w<strong>in</strong>ds up around the cyl<strong>in</strong>der correspond<strong>in</strong>gto the manifold of limit cycles. Each rotation corresponds to fir<strong>in</strong>g a spike. After the9th spike <strong>in</strong> the figure, the K + current becomes so large that repetitive spik<strong>in</strong>g cannotbe susta<strong>in</strong>ed. This happens when n slow passes the value 0.066, the limit cycle becomes ahomocl<strong>in</strong>ic orbit to a saddle, and then disappears. The burst<strong>in</strong>g trajectory jumps down


354 Burst<strong>in</strong>gn-nullcl<strong>in</strong>en slow =-0.03V-nullcl<strong>in</strong>en slow =0.0033membrane potential, V (mV)0-20-40-60-80-0.030slow K + gat<strong>in</strong>g variable, n slow0.030.060.0900.5fast K + gat<strong>in</strong>g variable, n1saddlenodebifurcationn slow =0.03n slow =0.06n slow =0.066saddlehomocl<strong>in</strong>ic orbitbifurcationspik<strong>in</strong>gn slow =0.0910.8thresholdrest<strong>in</strong>gVfastnn slow0.60.40.20fast K+ activation gate, n-80 -60 -40 -20 0membrane potential, V (mV)Figure 9.13: Burst<strong>in</strong>g trajectory of the I Na,p +I K +I K(M) -model <strong>in</strong> three-dimensionalphase space and its slices n slow = const.


Burst<strong>in</strong>g 355to the stable equilibrium correspond<strong>in</strong>g to the rest<strong>in</strong>g state. This jump<strong>in</strong>g correspondsto the term<strong>in</strong>ation of the active phase of burst<strong>in</strong>g and transition to rest<strong>in</strong>g. While atrest, K + current deactivates, n slow decreases, and so on.Figure 9.13 presents the <strong>in</strong>ner structure of the geometrical mechanism of burst<strong>in</strong>gof the I Na,p +I K +I K(M) -model with parameters as <strong>in</strong> Fig. 9.4. Other values of theparameters can result <strong>in</strong> different geometrical mechanisms, summarized <strong>in</strong> Sect. 9.3.In all cases, our approach is the same: freeze the slow subsystem by sett<strong>in</strong>g µ = 0;analyze phase portraits of the fast subsystem treat<strong>in</strong>g the slow variable as a bifurcationparameter; then glue the phase portraits, let µ ≠ 0 but small, and see how the evolutionof the slow subsystem switches the fast subsystem between spik<strong>in</strong>g and rest<strong>in</strong>g states.The method usually breaks down if µ is not small enough, because evolution of the“slow” variable starts to <strong>in</strong>terfere with that of the fast variable. How small is smalldepends on the particulars of the equations describ<strong>in</strong>g burst<strong>in</strong>g activity. One shouldworry when µ is greater than 0.1.9.2.3 Averag<strong>in</strong>gWhat governs the evolution of the slow variable u? To study this question, we describea well-known and widely used method that reduces the fast-slow system (9.1) to its slowcomponent. In fact, we have already used this method <strong>in</strong> Chapters 3 and 4 to reducethe dimension of neuronal models via substitution m = m ∞ (V ). Us<strong>in</strong>g essentially thesame ideas, we take advantage of the two time scales <strong>in</strong> (9.1) and get rid of the fastsubsystem by means of a substitution x = x(u).When the neuron is rest<strong>in</strong>g, its membrane potential is at an equilibrium and allfast gat<strong>in</strong>g variables are at their steady-state values, so that x = x rest (u). Us<strong>in</strong>g thisfunction <strong>in</strong> the slow equation <strong>in</strong> (9.1) we obta<strong>in</strong>˙u = µg(x rest (u), u) (reduced slow subsystem) , (9.2)which can easily be studied us<strong>in</strong>g the geometrical methods presented <strong>in</strong> Chapters 3 or4.Let us illustrate all the steps <strong>in</strong>volved us<strong>in</strong>g the I Na,p +I K +I K(M) -model with n slowbe<strong>in</strong>g the gat<strong>in</strong>g variable of the slow K + M-current. First, we freeze the slow subsystem,i.e., set τ slow (V ) = ∞ so that µ = 1/τ slow = 0, and determ<strong>in</strong>e numerically the rest<strong>in</strong>gpotential V rest as a function of the slow variable n slow . The function V = V rest (n slow )is depicted <strong>in</strong> Fig. 9.14, top, and it co<strong>in</strong>cides with the solid half-parabola <strong>in</strong> Fig. 9.13.Then, we use this function <strong>in</strong> the gat<strong>in</strong>g equation for the M-current to obta<strong>in</strong> (9.2)ṅ slow = (n ∞,slow (V rest (n slow )) − n slow )/τ slow (V rest (n slow )) = ḡ(n slow )depicted <strong>in</strong> Fig. 9.14, bottom. Notice that ḡ < 0, mean<strong>in</strong>g that n slow decreases whilethe fast subsystem rests. The rate of decrease is fairly small when n slow ≈ 0.A similar method of reduction, with an extra step, can be used when the fastsubsystem fires spikes. Let x(t) = x spike (t, u) be a periodic function correspond<strong>in</strong>g to


356 Burst<strong>in</strong>g0maxmembrane potential, V (mV)-30-60-65m<strong>in</strong>spik<strong>in</strong>gV spike (t,n slow )V rest (n slow )-70averaged function g0.0150.0100.0050-0.005-0.02spik<strong>in</strong>gg(n slow )rest<strong>in</strong>g0 0.02 0.04 0.06 0.08 0.1slow K + gat<strong>in</strong>g variable, n slowFigure 9.14: Spik<strong>in</strong>g solutions V (t) = V spike (t, u slow ), rest<strong>in</strong>g membrane potential V =V rest (n slow ), and the reduced slow subsystem ṅ slow = ḡ(n slow ) of the I Na,p +I K +I K(M) -model. The reduction is not valid <strong>in</strong> the shaded regions.an <strong>in</strong>f<strong>in</strong>ite spike tra<strong>in</strong> of the fast subsystem when u is frozen. Slices of this function areshown <strong>in</strong> Fig. 9.14, top. Let T (u) be the period of spik<strong>in</strong>g oscillation. The periodicallyforced slow subsystem˙u = µg(x spike (t, u), u) (slow subsystem) (9.3)can be averaged and reduced to a simpler modelẇ = µḡ(w) (averaged slow subsystem) (9.4)by a near-identity change of variables w = u + o(µ), where o(µ) denotes small terms oforder µ or less. Hereḡ(w) = 1T (w)∫ T (w)0g(x spike (t, w), w) dtis the average of g, shown <strong>in</strong> Fig. 9.14, bottom, for the I Na,p +I K +I K(M) -model. Checkthat ḡ(w) = g(x rest (w), w) when the fast subsystem is rest<strong>in</strong>g. Limit cycles of the


Burst<strong>in</strong>g 357membranepotential, V (mV)slow K + activationgate, n slow0-20-40-600.080.04w (averaged)u (orig<strong>in</strong>al) Figure 9.15: The I Na,p +I K +I K(M) -model burster with orig<strong>in</strong>al andaveraged slow variable.00 10 20 30 40 50timeaveraged slow subsystem corresponds to burst<strong>in</strong>g dynamics, whereas equilibria correspondto either rest<strong>in</strong>g or periodic spik<strong>in</strong>g states of the full system (9.1) — the resultknown as Pontryag<strong>in</strong>–Rodyg<strong>in</strong> (1960) theorem. Interest<strong>in</strong>g regimes correspond to theco-existence of limit cycles and equilibria of the slow averaged system.The ma<strong>in</strong> purpose of averag<strong>in</strong>g consists <strong>in</strong> substitut<strong>in</strong>g the wiggle trajectory ofu(t) by a smooth trajectory of w(t), as we illustrate <strong>in</strong> Fig. 9.15. We purposely useda different letter, w, for the new slow variable to stress that (9.4) is not equivalentto (9.3). Their solutions are o(µ)-close to each other only when certa<strong>in</strong> conditions aresatisfied, see Guckenheimer and Holmes (1983) or Hoppensteadt and Izhikevich (1997).In particular, this straightforward averag<strong>in</strong>g breaks down when u passes slowly thebifurcation values. For example, the period, T (u), of x spike (t, u) may go to <strong>in</strong>f<strong>in</strong>ity, ashappens near saddle-node on <strong>in</strong>variant circle and saddle homocl<strong>in</strong>ic orbit bifurcations,or transients may take as long as 1/µ, or the averaged system (9.4) is not smooth. Allthese cases are encountered <strong>in</strong> burst<strong>in</strong>g models. Thus, one can use the reduced slowsubsystem only when the fast subsystem is sufficiently far away from a bifurcation,e.g., away from the shaded regions <strong>in</strong> Fig. 9.14.9.2.4 Equivalent voltageLet us consider a “2+1” burster with a slow subsystem depend<strong>in</strong>g only on the slow variableand the membrane potential V , as <strong>in</strong> the I Na,p +I K +I K(M) -model. The nonl<strong>in</strong>earequationg(V, u) = ḡ(u) (9.5)can be solved for V . The solution, V = V equiv (u), is referred to as be<strong>in</strong>g the equivalentvoltage (Kepler et al. 1992, Bertram et al. 1995), because it replaces the periodicfunction x spike (t, u) <strong>in</strong> (9.3) by an “equivalent” value of the membrane potential, sothat the reduced slow subsystem (9.3) has the same form,˙u = µg(V equiv (u), u) (slow subsystem), (9.6)as <strong>in</strong> (9.1). Check that V equiv (u) = V rest (u) when the fast subsystem is rest<strong>in</strong>g. An<strong>in</strong>terest<strong>in</strong>g mathematical possibility is when V equiv dur<strong>in</strong>g spik<strong>in</strong>g is below V rest , lead<strong>in</strong>g


358 Burst<strong>in</strong>g0membrane potential, V (mV)-10-20-30-40-50-60-70V equiv (n slow )n slow (V)V equiv (n slow )0 0.05 0.1K + activation gate, nslowlimit cycle (max)unstable equilblimit cycle (m<strong>in</strong>)n slow (V)saddlenode0 0.05 0.1K + activation gate, nslowFigure 9.16: Projection of burst<strong>in</strong>g trajectory of the I Na,p +I K +I K(M) -model onto the(n slow , V ) plane.to bizarre bursters hav<strong>in</strong>g amplify<strong>in</strong>g slow currents, such as the one <strong>in</strong> Ex. 10.We depict the equivalent voltage of the I Na,p +I K +I K(M) -model <strong>in</strong> Fig. 9.16, left(variable u corresponds to n slow ). In the same figure, we depict the steady-state activationfunction n = n ∞,slow (V ) (notice the flipped coord<strong>in</strong>ate system). We <strong>in</strong>terpretthe two curves as fast and slow nullcl<strong>in</strong>es of the reduced (V, n slow )-system. Dur<strong>in</strong>g theactive (spik<strong>in</strong>g) phase of burst<strong>in</strong>g, the reduced system slides along the upper branch ofV equiv (n slow ) to the right. When it reaches the end of the branch, it falls down to thelower branch correspond<strong>in</strong>g to rest<strong>in</strong>g, and slides along this branch to the left. When itreaches the left end of the lower branch, it jumps up to the upper branch, and therebycloses the hysteresis loop. Fig. 9.16, right, summarizes all the <strong>in</strong>formation needed tounderstand the transitions between rest<strong>in</strong>g and spik<strong>in</strong>g states <strong>in</strong> this model. It depictsthe burst<strong>in</strong>g trajectory, loci of equilibria of the fast subsystem, and the voltage range ofspik<strong>in</strong>g limit cycle as a function of the slow gate n slow . With some experience, one canread this complicated figure and visualize the three-dimensional geometry underly<strong>in</strong>gburst<strong>in</strong>g dynamics.9.2.5 Hysteresis loops and slow wavesSusta<strong>in</strong>ed burst<strong>in</strong>g activity of the fast-slow system (9.1) corresponds to periodic (orchaotic) activity of the reduced slow subsystem (9.6). Depend<strong>in</strong>g on the dimensionof u, i.e., on the number of slow variables, there could be two fundamentally differentways the slow subsystem oscillates.If the slow variable u is one-dimensional, then there must be a bistability of rest<strong>in</strong>gand spik<strong>in</strong>g states of the fast subsystem so that u oscillates via a hysteresis loop. Thatis, the reduced equation (9.6) consists of two parts, one for V equiv (u) correspond<strong>in</strong>gto spik<strong>in</strong>g, and one for V equiv (u) correspond<strong>in</strong>g to rest<strong>in</strong>g of the fast subsystem, as


Burst<strong>in</strong>g 359Spik<strong>in</strong>gu(t)activation of outward currents<strong>in</strong>activation of <strong>in</strong>ward currentsudeactivation of outward currentsde<strong>in</strong>activation of <strong>in</strong>ward currentsRest<strong>in</strong>gFigure 9.17: Hysteresis-loop periodic burst<strong>in</strong>g.Spik<strong>in</strong>gPerturbationRest<strong>in</strong>gFigure 9.18: Burst excitability: A perturbation causes a burst of spikes.<strong>in</strong> Fig. 9.16, left. Such a hysteresis loop burst<strong>in</strong>g can also occur when u is multidimensional,as we illustrate <strong>in</strong> Fig. 9.17. The vector-field on the top (spik<strong>in</strong>g) leafpushes u outside the spik<strong>in</strong>g area, whereas the vector-field on the bottom (rest<strong>in</strong>g) leafpushes u outside the rest<strong>in</strong>g area. As a result, u visits the spik<strong>in</strong>g and rest<strong>in</strong>g areasperiodically, and the model exhibits hysteresis-loop burst<strong>in</strong>g.If rest<strong>in</strong>g x does not push u <strong>in</strong>to the spik<strong>in</strong>g area, but leaves it <strong>in</strong> the bistable area,then the neuron exhibits burst excitability: It has quiescent excitable dynamics, but itsresponse to perturbations is not a s<strong>in</strong>gle spike, but a burst of spikes, as we illustrate <strong>in</strong>Fig. 9.18. Grade III bursters of hippocampus (Fig. 8.34Eb) produce such a response,often called complex spike response, to brief stimuli. In general, many bistable modelsare bistable only because they neglect slow currents and other homeostatic processespresent <strong>in</strong> real neurons. If the currents are taken <strong>in</strong>to account, then the models becomebistable on a short time scale and burst-excitable on a longer time scale. This justifieswhy many researchers refer to bistable systems as excitable, implicitly assum<strong>in</strong>g thatthe response to superthreshold perturbations is either a s<strong>in</strong>gle spike or a long tra<strong>in</strong> ofspikes.If the fast subsystem does not have a coexistence of rest<strong>in</strong>g and spik<strong>in</strong>g states,then the reduced slow subsystem (9.6) must be at least two-dimensional to exhibit


360 Burst<strong>in</strong>gI=0I=4.5425 ms 25 mVI=5I=7I=7.6I=7.7I=80IFigure 9.19: Bifurcations of burst<strong>in</strong>g solutions <strong>in</strong> the I Na,p +I K +I K(M) -model as themagnitude of the <strong>in</strong>jected dc-current I changes.susta<strong>in</strong>ed autonomous oscillation (however, see Ex. 6). Such an oscillation produces adepolarization wave that drives the fast subsystem to spik<strong>in</strong>g and back, as <strong>in</strong> Fig. 9.3.We refer to such bursters as slow-wave bursters. Quite often, however, the slow subsystemof a slow-wave burster needs the feedback from the fast subsystem to oscillate.For example, <strong>in</strong> Sect. 9.3.2 we consider slow-wave burst<strong>in</strong>g <strong>in</strong> the I Na,p +I K +I Na,slow +I K(M) -model, whose slow subsystem consists of two uncoupled equations, and hencecannot oscillate by itself unless the fast subsystem is present.9.2.6 Bifurcations “rest<strong>in</strong>g ↔ burst<strong>in</strong>g ↔ spik<strong>in</strong>g”Switch<strong>in</strong>g between spik<strong>in</strong>g and rest<strong>in</strong>g states dur<strong>in</strong>g burst<strong>in</strong>g occurs because the slowvariable drives the fast subsystem through bifurcations of equilibria and limit cycleattractors. These bifurcations play an important role <strong>in</strong> our classification of burstersand <strong>in</strong> understand<strong>in</strong>g their neuro-computational properties. We discuss them <strong>in</strong> detail<strong>in</strong> the next section.S<strong>in</strong>ce the fast subsystem goes through bifurcations, does this mean that the entiresystem (9.1) undergoes bifurcations dur<strong>in</strong>g burst<strong>in</strong>g? The answer is NO. As long asparameters of (9.1) are fixed, the system as a whole does not undergo any bifurcations,no matter how small µ is. The system can exhibit periodic, quasi-periodic or evenchaotic burst<strong>in</strong>g activity, but its (m + k)-dimensional phase portrait does not change.The only way to make system (9.1) undergo a bifurcation is to change its param-


Burst<strong>in</strong>g 361eters. For example, <strong>in</strong> Fig. 9.19 we change the magnitude of the <strong>in</strong>jected dc-currentI <strong>in</strong> the I Na,p +I K +I K(M) -model. Apparently, no burst<strong>in</strong>g exists when I = 0. Then,repetitive burst<strong>in</strong>g appears with a large <strong>in</strong>terburst period that decreases as I <strong>in</strong>creases.The value I = 5 was used to obta<strong>in</strong> burst<strong>in</strong>g solutions <strong>in</strong> Fig. 9.12 and Fig. 9.13.Increas<strong>in</strong>g I further <strong>in</strong>creases the duration of each burst, until it becomes <strong>in</strong>f<strong>in</strong>ite, i.e.,burst<strong>in</strong>g turns <strong>in</strong>to tonic spik<strong>in</strong>g. When I > 8, the slow K + current is not enough tostop spik<strong>in</strong>g.In Fig. 9.20 we depict the geometry of burst<strong>in</strong>g <strong>in</strong> the I Na,p +I K +I K(M) -model whenI = 3, i.e., just before periodic burst<strong>in</strong>g appears, and when I = 10, i.e., just afterburst<strong>in</strong>g turns <strong>in</strong>to tonic spik<strong>in</strong>g.When I = 3, the nullcl<strong>in</strong>e of the slow subsystem n slow = n ∞,slow (V ) <strong>in</strong>tersects thelocus of stable equilibria of the fast subsystem. The <strong>in</strong>tersection po<strong>in</strong>t is a globallystable equilibrium of the full system (9.1). Small perturbations, whether <strong>in</strong> the Vdirection, n direction, or n slow direction subside, whereas a large perturbation, e.g., <strong>in</strong>the V direction, that moves the membrane potential to the open square <strong>in</strong> the figure,<strong>in</strong>itiates a transient (phasic) burst of 7 spikes. Increas<strong>in</strong>g the magnitude of the <strong>in</strong>jectedcurrent I shifts the saddle-node parabola to the right. When I ≈ 4.54, the nullcl<strong>in</strong>eof the slow subsystem does not <strong>in</strong>tersect the locus of stable equilibria, and the rest<strong>in</strong>gstate no longer exists, as <strong>in</strong> Fig. 9.16, right. (There is still a global steady state, but itis not stable.)Further <strong>in</strong>crease of the magnitude of the <strong>in</strong>jected current I results <strong>in</strong> the <strong>in</strong>tersectionof the nullcl<strong>in</strong>e of the slow subsystem with the equivalent voltage function V equiv (n slow ).The <strong>in</strong>tersection, marked by the black circle <strong>in</strong> Fig. 9.20, right, corresponds to a globallystable (spik<strong>in</strong>g) limit cycle of the full system (9.1). A sufficiently strong perturbationcan push the state of the fast subsystem <strong>in</strong>to the attraction doma<strong>in</strong> of the stable(rest<strong>in</strong>g) equilibrium. While the fast subsystem is rest<strong>in</strong>g, the slow variable decreases,i.e., K + current deactivates, the rest<strong>in</strong>g equilibrium disappears and repetitive spik<strong>in</strong>gresumes.Figures 9.19 and Fig. 9.20 illustrate possible transitions between burst<strong>in</strong>g and rest<strong>in</strong>g,and burst<strong>in</strong>g and tonic spik<strong>in</strong>g. There could be other routes of emergence ofburst<strong>in</strong>g solutions from rest<strong>in</strong>g or spik<strong>in</strong>g, some of them are <strong>in</strong> Fig. 9.21. Each suchroute corresponds to a bifurcation <strong>in</strong> the full system (9.1) with some µ > 0. For example,the case a → 0 corresponds to supercritical Andronov-Hopf bifurcation; the casec → ∞ corresponds to a saddle-node on <strong>in</strong>variant circle or saddle homocl<strong>in</strong>ic orbit bifurcation;the case d → ∞ corresponds to a periodic orbit with a homocl<strong>in</strong>ic structure,e.g., blue-sky catastrophe, fold limit cycle on homocl<strong>in</strong>ic torus bifurcation, or someth<strong>in</strong>gmore complicated. The transitions “burst<strong>in</strong>g ↔ spik<strong>in</strong>g often exhibit chaotic(irregular) activity, so Fig. 9.21 if probably a great oversimplification. Understand<strong>in</strong>gand classify<strong>in</strong>g all possible bifurcations lead<strong>in</strong>g to burst<strong>in</strong>g dynamics is an importantbut open problem; see Ex. 27.


362 Burst<strong>in</strong>gmembrane potential, V (mV)I=3 I=100-20-40-600 50 100 150time (ms)membrane potential, V (mV)0-20-40-600 50 100 150time (ms)spik<strong>in</strong>gstable (spik<strong>in</strong>g)limit cyclethresholdVfastnglobalrest<strong>in</strong>gstaterest<strong>in</strong>gn slowmembrane potential, V (mV)0-20-40-60n slow (V)saddlenode0 0.05 0.1slow K + gat<strong>in</strong>g variable, n slowmembrane potential, V (mV)0-20-40-60limit cycle (max)V equiv (n slow )limit cycle (m<strong>in</strong>)n slow (V)stable (spik<strong>in</strong>g) limit cyclesaddlenode0 0.05 0.1slow K + gat<strong>in</strong>g variable, n slowFigure 9.20: Burst excitability (I = 3, left) and periodic spik<strong>in</strong>g (I = 10, right) <strong>in</strong> theI Na,p +I K +I K(M) -model.


Burst<strong>in</strong>g 363abdcburst<strong>in</strong>ga 0 cc 0 b 0 d 0drest<strong>in</strong>gspik<strong>in</strong>gFigure 9.21: Possible transitions between repetitive burst<strong>in</strong>g and rest<strong>in</strong>g, and repetitiveburst<strong>in</strong>g and repetitive spik<strong>in</strong>g.spik<strong>in</strong>g state(limit cycle)bifurcation of equilibrium(transition to spik<strong>in</strong>g)bifurcation of limit cycle(transition to rest<strong>in</strong>g)rest<strong>in</strong>g state (equilibrium)rest<strong>in</strong>g state (equilibrium)Figure 9.22: Two important bifurcations associated with fast-slow burst<strong>in</strong>g.


364 Burst<strong>in</strong>g9.3 ClassificationIn Fig. 9.22 we identify two important bifurcations of the fast subsystem that areassociated with burst<strong>in</strong>g activity <strong>in</strong> the fast-slow burster (9.1):• (rest<strong>in</strong>g → spik<strong>in</strong>g) Bifurcation of an equilibrium attractor that results <strong>in</strong>transition from rest<strong>in</strong>g to repetitive spik<strong>in</strong>g.• (spik<strong>in</strong>g → rest<strong>in</strong>g) Bifurcation of the limit cycle attractor that results <strong>in</strong>transition from spik<strong>in</strong>g to rest<strong>in</strong>g.The ionic basis of burst<strong>in</strong>g, i.e., the f<strong>in</strong>e electrophysiological details, determ<strong>in</strong>e thek<strong>in</strong>d of bifurcations <strong>in</strong> Fig. 9.22. The bifurcations, <strong>in</strong> turn, determ<strong>in</strong>e the neurocomputationalproperties of fast-slow bursters, discussed <strong>in</strong> Sect. 9.4.A complete topological classification of bursters based on these two bifurcations isprovided by Izhikevich (2000), who identified 120 different topological types. Here, weconsider only 16 planar po<strong>in</strong>t-cycle co-dimension-1 fast-slow bursters. We say that afast-slow burster is planar when its fast subsystem is two-dimensional. We emphasizeplanar bursters because they have a greater chance to be encountered <strong>in</strong> computersimulations (but not necessarily <strong>in</strong> nature). We say that a burster is of the po<strong>in</strong>tcycletype when its rest<strong>in</strong>g state is a stable equilibrium po<strong>in</strong>t and its spik<strong>in</strong>g state is astable limit cycle. All bursters considered so far, <strong>in</strong>clud<strong>in</strong>g those <strong>in</strong> Fig. 9.1, are of thepo<strong>in</strong>t-cycle type. Other, less common types, such as cycle-cycle and po<strong>in</strong>t-po<strong>in</strong>t, areconsidered as exercises.We consider here only bifurcations of co-dimension 1, i.e. those that need onlyone parameter and hence are more likely to be encountered <strong>in</strong> nature. Hav<strong>in</strong>g a twodimensionalfast subsystem imposes severe restriction on possible co-dimension-1 bifurcationsof the rest<strong>in</strong>g and spik<strong>in</strong>g states. In particular, there are only 4 bifurcations ofequilibria and 4 bifurcations of limit cycles, which we consider <strong>in</strong> Chap. 6 and summarize<strong>in</strong> Figures 6.46 and 6.47. Any comb<strong>in</strong>ation of them results <strong>in</strong> a dist<strong>in</strong>ct topologicaltype of fast-slow burst<strong>in</strong>g, hence there are 4 × 4 = 16 such bursters, summarized <strong>in</strong>Fig. 9.23.We name the bursters accord<strong>in</strong>g to the type of the bifurcations of the rest<strong>in</strong>g andspik<strong>in</strong>g states. To keep the names short, we refer to saddle-node on <strong>in</strong>variant circlebifurcation as just a “circle” bifurcation because it is the only co-dimension-1 bifurcationon a circle manifold S 1 . We refer to supercritical Andronov-Hopf bifurcationas just “Hopf” bifurcation, to subcritical Andronov-Hopf as “subHopf”, to fold limitcycle bifurcation as “fold cycle”, and to saddle homocl<strong>in</strong>ic orbit bifurcation as “homocl<strong>in</strong>ic”bifurcation. Thus, the burst<strong>in</strong>g pattern exhibited by the I Na,p +I K +I K(M) -model<strong>in</strong> Fig. 9.13 is of the “fold/homocl<strong>in</strong>ic” type because the rest<strong>in</strong>g state disappears via“fold” bifurcation and the spik<strong>in</strong>g limit cycle attractor disappears via saddle “homocl<strong>in</strong>ic”orbit bifurcation.Similarly to Fig. 9.13, we depict the geometry of the other bursters <strong>in</strong> Fig. 9.24.This figure gives only examples and it does not exhaust all possibilities. Let us considersome most common burst<strong>in</strong>g types <strong>in</strong> detail.


Burst<strong>in</strong>g 365bifurcations of limit cyclessaddle-nodeon <strong>in</strong>variantcirclesaddlehomocl<strong>in</strong>icorbitsupercriticalAndronov-Hopffoldlimitcyclesaddle-node(fold)fold/circlefold/homocl<strong>in</strong>icfold/Hopffold/fold cyclebifurcations of equilibriasaddle-nodeon <strong>in</strong>variantcirclesupercriticalAndronov-HopfsubcriticalAndronov-Hopfcircle/circleHopf/circlesubHopf/circlecircle/homocl<strong>in</strong>icHopf/homocl<strong>in</strong>icsubHopf/homocl<strong>in</strong>iccircle/HopfHopf/HopfsubHopf/Hopfcircle/fold cycleHopf/fold cyclesubHopf/fold cycleFigure 9.23: Classification of planar po<strong>in</strong>t-cycle fast-slow bursters based on the codimention-1bifurcations of the rest<strong>in</strong>g and spik<strong>in</strong>g states of the fast subsystem.9.3.1 fold/homocl<strong>in</strong>icWhen the rest<strong>in</strong>g state disappears via a saddle-node (fold) bifurcation and the spik<strong>in</strong>glimit cycle disappears via saddle homocl<strong>in</strong>ic orbit bifurcation, the burster is said to beof the “fold/homocl<strong>in</strong>ic” type depicted <strong>in</strong> Fig. 9.25. Notice the bistability of rest<strong>in</strong>gand spik<strong>in</strong>g states, result<strong>in</strong>g <strong>in</strong> hysteresis loop oscillation of the slow subsystem.“Fold/homocl<strong>in</strong>ic” burst<strong>in</strong>g is quite common <strong>in</strong> neuronal models, for example <strong>in</strong>the I Na,p +I K +I K(M) -model considered <strong>in</strong> this chapter; see Fig. 9.13. It was first characterized<strong>in</strong> the context of the <strong>in</strong>sul<strong>in</strong>-produc<strong>in</strong>g pancreatic β-cells <strong>in</strong> Fig. 9.26, with<strong>in</strong>tracellular concentration of Ca 2+ ions be<strong>in</strong>g the slow resonant variable (Chay andKeizer 1983). Neurons located <strong>in</strong> the pre-Botz<strong>in</strong>ger complex, a region that is associatedwith generat<strong>in</strong>g the rhythm for breath<strong>in</strong>g, also exhibit this k<strong>in</strong>d of burst<strong>in</strong>g(Butera et al. 1999), as shown <strong>in</strong> Fig. 9.27. Intr<strong>in</strong>sic burst<strong>in</strong>g (IB) and chatter<strong>in</strong>g(CH) behavior of the simple model <strong>in</strong> Sect. 8.2 could be of the “fold/homocl<strong>in</strong>ic” typetoo, provided that the parameter a is sufficiently small. Because of the dist<strong>in</strong>ct squarewaveshape of oscillations of the membrane potential <strong>in</strong> Figures 9.26 and Fig. 9.27,this burst<strong>in</strong>g was called “square-wave” burst<strong>in</strong>g <strong>in</strong> earlier studies. S<strong>in</strong>ce many types ofbursters resemble square waves, referr<strong>in</strong>g to a burster by its shape is mislead<strong>in</strong>g andshould be avoided.In Fig. 9.25, bottom, we depict a typical configuration of nullcl<strong>in</strong>es of the fastsubsystem dur<strong>in</strong>g “fold/homocl<strong>in</strong>ic” burst<strong>in</strong>g. The rest<strong>in</strong>g state of the membrane po-


366 Burst<strong>in</strong>gbifurcation of spik<strong>in</strong>g statesaddle-node on<strong>in</strong>variant circlesaddlehomocl<strong>in</strong>ic orbitsupercriticalAndronov-Hopffold limit cycle"Fold/Circle" Burst<strong>in</strong>g"Fold/Homocl<strong>in</strong>ic" Burst<strong>in</strong>g "Fold/Hopf" Burst<strong>in</strong>g "Fold/Fold Cycle" Burst<strong>in</strong>gsaddle-node (fold)"Circle/Circle" Burst<strong>in</strong>g "Circle/Homocl<strong>in</strong>ic" Burst<strong>in</strong>g "Circle/Hopf" Burst<strong>in</strong>g "Circle/Fold Cycle" Burst<strong>in</strong>gbifurcation of rest<strong>in</strong>g statesaddle-node on<strong>in</strong>variant circlesupercriticalAndronov-Hopf"Hopf/Circle" Burst<strong>in</strong>g "Hopf/Homocl<strong>in</strong>ic" Burst<strong>in</strong>g "Hopf/Hopf" Burst<strong>in</strong>g "Hopf/Fold Cycle" Burst<strong>in</strong>g"SubHopf/Circle" Burst<strong>in</strong>g"SubHopf/Homocl<strong>in</strong>ic"Burst<strong>in</strong>g"SubHopf/Hopf" Burst<strong>in</strong>g"SubHopf/Fold Cycle"Burst<strong>in</strong>gsubcriticalAndronov-HopfFigure 9.24: Examples of “2+1” po<strong>in</strong>t-cycle fast-slow co-dimension-1 bursters ofhysteresis-loop type (modified from Izhikevich 2000). Dashed cha<strong>in</strong>s of arrows showtransitions that might <strong>in</strong>volve other bifurcations not relevant to the burst<strong>in</strong>g type.


spik<strong>in</strong>gBurst<strong>in</strong>g 367foldbifurcationxurestsaddlehomocl<strong>in</strong>icorbitbifurcationfold bifurcationsaddle homocl<strong>in</strong>icorbit bifurcationFigure 9.25: “Fold/homocl<strong>in</strong>ic” burst<strong>in</strong>g: The rest<strong>in</strong>g state disappears via saddle-node(fold) bifurcation and the spik<strong>in</strong>g limit cycle disappears via saddle homocl<strong>in</strong>ic orbitbifurcation.tential corresponds to the left stable equilibrium, which is the <strong>in</strong>tersection of the leftknee of the fast N-shaped nullcl<strong>in</strong>e with the slow nullcl<strong>in</strong>e. Dur<strong>in</strong>g rest<strong>in</strong>g, the N-shapenullcl<strong>in</strong>e slowly moves up, until its knee touches the slow nullcl<strong>in</strong>e at a saddle-nodepo<strong>in</strong>t. Right after this moment, the rest<strong>in</strong>g state disappears via saddle-node (fold) bifurcation,hence the part “fold” <strong>in</strong> the name of the burster. After the fold bifurcation,the membrane potential jumps up to the stable limit cycle correspond<strong>in</strong>g to repetitivespik<strong>in</strong>g. Dur<strong>in</strong>g the spik<strong>in</strong>g state, the N-shaped nullcl<strong>in</strong>e slowly moves down, and themiddle (saddle) equilibrium moves away from the rest<strong>in</strong>g state toward the limit cycle.After a while, the limit cycle becomes a homocl<strong>in</strong>ic trajectory to the saddle, and thenthe cycle disappears via saddle homocl<strong>in</strong>ic orbit bifurcation, hence the part “homocl<strong>in</strong>ic”<strong>in</strong> the name of the burster. After this bifurcation, the membrane potentialjumps down to the rest<strong>in</strong>g state and closes the hysteresis loop.


368 Burst<strong>in</strong>gFigure 9.26: Putative “fold/homocl<strong>in</strong>ic” burst<strong>in</strong>g <strong>in</strong> a pancreatic β-cell (modified fromK<strong>in</strong>ard et al. 1999).0 mV20 mV1 secFigure 9.27: Putative ”fold/homocl<strong>in</strong>ic” burst<strong>in</strong>g <strong>in</strong> a cell located <strong>in</strong> pre-Botz<strong>in</strong>gercomplex of rat bra<strong>in</strong> stem (data k<strong>in</strong>dly shared by Christopher A. Del Negro and JackL. Feldman, <strong>Systems</strong> Neurobiology Laboratory, Department of Neurobiology, UCLA.)Suppose that the hysteresis loop oscillation of the slow variable u has a small amplitude.That is, the saddle-node bifurcation and the saddle homocl<strong>in</strong>ic orbit bifurcationoccur for nearby values of the parameter u. In this case, the fast subsystem of (9.1)is near co-dimension-2 saddle-node homocl<strong>in</strong>ic orbit bifurcation, depicted <strong>in</strong> Fig. 9.28and studied <strong>in</strong> Sect. 6.3.6. The figure shows a two-parameter unfold<strong>in</strong>g of the bifurcation,treat<strong>in</strong>g u ∈ R 2 as the parameter. A stable equilibrium (rest<strong>in</strong>g state) exists <strong>in</strong>the left half-plane, and a stable limit cycle (spik<strong>in</strong>g state) exists <strong>in</strong> the right half-planeof the figure and <strong>in</strong> the shaded (bistable) region. “Fold/homocl<strong>in</strong>ic” burst<strong>in</strong>g occurswhen the bifurcation parameter, be<strong>in</strong>g a slow variable, oscillates between the rest<strong>in</strong>gand spik<strong>in</strong>g states through the shaded region. Due to the bistability, the parametercould be one-dimensional. Other trajectories of the slow parameter correspond to othertypes of burst<strong>in</strong>g.In Ex. 16 we prove that there is a piece-wise cont<strong>in</strong>uous change of variables thattransforms any “fold/homocl<strong>in</strong>ic” burster with fast subsystem near such a bifurcation<strong>in</strong>to the canonical model (see Sect. 8.1.5)˙v = I + v 2 − u ,˙u = −µu ,(9.7)


Burst<strong>in</strong>g 369homocl<strong>in</strong>ic orbit bifurcationsaddle-node bifurcation saddle-node on<strong>in</strong>variant circle bifurcationsaddle-nodehomocl<strong>in</strong>ic orbitbifurcationcircle/homocl<strong>in</strong>icburst<strong>in</strong>ghomocl<strong>in</strong>iccircle/circleburst<strong>in</strong>gfold/homocl<strong>in</strong>icburst<strong>in</strong>gcirclefoldfold/circleburst<strong>in</strong>gFigure 9.28: A neural system near co-dimension-2 saddle-node homocl<strong>in</strong>ic orbit bifurcation(center dot) can exhibit four different types of fast-slow burst<strong>in</strong>g, depend<strong>in</strong>g onthe trajectory of the slow variable u ∈ R 2 <strong>in</strong> the two-dimensional parameter space.Solid (dotted) l<strong>in</strong>es correspond to spik<strong>in</strong>g (rest<strong>in</strong>g) regimes.with an after-spike resett<strong>in</strong>g:if v = +∞, then v ← 1 and u ← u + d.Here v ∈ R is the re-scaled membrane potential of the neuron, u ∈ R is the re-scalednet outward (resonant) current that provides a negative feedback to v, and I, d andµ ≪ 1 are parameters. This model is related to the canonical model considered <strong>in</strong>Sect. 8.1.4, and it is simplified further <strong>in</strong> Ex. 15.The fast subsystem ˙v = (I − u) + v 2 is the normal form for the saddle-node bifurcation,and with the resett<strong>in</strong>g it is known as the quadratic <strong>in</strong>tegrate-and-fire neuron(Sect. 3.3.8). When u > I, there is a stable equilibrium v rest = − √ u − I correspond<strong>in</strong>gto the rest<strong>in</strong>g state. While the parameter u slowly decreases toward u = 0, the stableequilibrium and the saddle equilibrium v thresh = + √ u − I approach and annihilateeach other at u = I via saddle-node (fold) bifurcation. When u < I, the membranepotential v <strong>in</strong>creases and escapes to <strong>in</strong>f<strong>in</strong>ity <strong>in</strong> a f<strong>in</strong>ite time, i.e., it fires a spike. (Insteadof <strong>in</strong>f<strong>in</strong>ity, any large value can be used <strong>in</strong> simulations.) The spike activates fastoutward currents and resets v to 1, as <strong>in</strong> Fig. 9.29. It also activates slow currents and<strong>in</strong>crements u by d. If the reset value 1 is greater than the threshold potential v thresh ,the fast subsystem fires another spike, and so on, even when u > I; see Fig. 9.29.S<strong>in</strong>ce each spike <strong>in</strong>creases u, the repetitive spik<strong>in</strong>g stops when u = I + 1 via saddlehomocl<strong>in</strong>ic orbit bifurcation. The membrane potential jumps down to the rest<strong>in</strong>gstate, the hysteresis loop is closed, and the variable u decreases (recovers) to <strong>in</strong>itiateanother “fold/homocl<strong>in</strong>ic” burst. One can vary I <strong>in</strong> the canonical model (9.7) to studytransitions from quiescence to burst<strong>in</strong>g to tonic spik<strong>in</strong>g, as <strong>in</strong> Fig. 9.20.


370 Burst<strong>in</strong>g108spik<strong>in</strong>gslow variable, u fast variable, v64homocl<strong>in</strong>ic2reset0fold3 0.5 1 1.5 2 2.52homocl<strong>in</strong>icslow variable, u1fold00 50time100 150thresholdrest<strong>in</strong>gFigure 9.29: “Fold/homocl<strong>in</strong>ic” burst<strong>in</strong>g <strong>in</strong> the canonical model (9.7) with parametersµ = 0.02, I = 1 and d = 0.2.saddle-node on<strong>in</strong>variant circle bifurcationspik<strong>in</strong>gxurestsaddle-node on <strong>in</strong>variantcircle bifurcationsaddle-nodeon <strong>in</strong>variantcircle bifurcationsaddle-nodeon <strong>in</strong>variantcircle bifurcationFigure 9.30: “Circle/circle” burst<strong>in</strong>g: The rest<strong>in</strong>g state disappears via saddle-node on<strong>in</strong>variant circle bifurcation and the spik<strong>in</strong>g limit cycle disappears via saddle-node on<strong>in</strong>variant circle bifurcation.


Burst<strong>in</strong>g 371Figure 9.31: Putative “circle/circle” burst<strong>in</strong>g pacemaker activity of neuron R 15 <strong>in</strong>abdom<strong>in</strong>al ganglion of the mollusk Aplysia (modified from Levitan and Levitan 1988).<strong>in</strong>terspike period (sec)21.510.500 5 10 15spike number<strong>in</strong>terspike frequency (Hz)2.521.510.500 5 10 15spike numberFigure 9.32: The <strong>in</strong>terspike period <strong>in</strong> the “circle/circle” burst<strong>in</strong>g <strong>in</strong> Fig. 9.31 resemblesa parabola, which motivates the name “parabolic burst<strong>in</strong>g” used <strong>in</strong> earlier studies.9.3.2 circle/circleWhen the equilibrium correspond<strong>in</strong>g to the rest<strong>in</strong>g state disappears via a saddle-nodeon <strong>in</strong>variant circle bifurcation, and the limit cycle attractor correspond<strong>in</strong>g to spik<strong>in</strong>gstate disappears via another saddle-node on <strong>in</strong>variant circle bifurcation, the burster issaid to be of the “circle/circle” type shown <strong>in</strong> Fig. 9.30. S<strong>in</strong>ce the bifurcation doesnot produce a co-existence of attractors, there is usually no hysteresis loop, and theburst<strong>in</strong>g is of the slow-wave type with at least two slow variables. (An example of“circle/circle” hysteresis loop burst<strong>in</strong>g <strong>in</strong> a “2+1” system is provided by Izhikevich(2000)).“Circle/circle” burst<strong>in</strong>g is a prom<strong>in</strong>ent feature of the R 15 cells <strong>in</strong> the abdom<strong>in</strong>al ganglionof the mollusk Aplysia, shown <strong>in</strong> Fig. 9.31 (Plant 1981). It was called “parabolic”burst<strong>in</strong>g <strong>in</strong> earlier studies because the <strong>in</strong>terspike period depicted <strong>in</strong> Fig. 9.32 was erroneouslythought to be a parabola. In Sect. 6.1.2 we showed that when a systemundergoes a saddle-node on <strong>in</strong>variant circle bifurcation, its period scales as 1/ √ λ,where λ is the distance to the bifurcation. Two pieces of this function, put togetheras <strong>in</strong> Fig. 9.32, do <strong>in</strong>deed resemble a parabola. But so does the <strong>in</strong>terspike period of a“circle/homocl<strong>in</strong>ic” burster.To transform the I Na,p +I K -model to a “circle/circle” burster, we take the parametersas <strong>in</strong> Fig. 4.1a so that there is a saddle-node on <strong>in</strong>variant circle bifurcation whenI = 4.51 (see Sect. 6.1.2). Its nullcl<strong>in</strong>es and phase portrait look similar to those <strong>in</strong>Fig. 9.30. Then, we add one amplify<strong>in</strong>g and one resonant current with gat<strong>in</strong>g variablesṁ slow = (m ∞,slow (V ) − m slow )/τ Na,slow (V ) (slow I Na,slow ),ṅ slow = (n ∞,slow (V ) − n slow )/τ K(M) (V ) (slow I K(M) ),


372 Burst<strong>in</strong>g(a)membrane potential, V (mV)(b)activation gate0-20-40-60-800 100 200 300 400time (ms)0.250.20.150.10.05n slow (t)m slow (t)00 100 200 300 400time (ms)(c)K + activation gate, nslow(d)membrane potential, V (mV)-20-40-60-800.120.10.080.060.040.0220000 0.05 0.1 0.15 0.2 0.25Na + activation gate, m slow0.1rest<strong>in</strong>gK + activation gate, n slow0.05rest<strong>in</strong>gspik<strong>in</strong>g00spik<strong>in</strong>g0.2Na + activationgate, m slowFigure 9.33: “Circle/circle” burst<strong>in</strong>g <strong>in</strong> the I Na,p +I K +I Na,slow +I K(M) -model. Parametersof the fast I Na,p +I K -subsystem are the same as <strong>in</strong> Fig. 4.1a with I = 5. Slow Na +current has V 1/2 = −40 mV, k = 5 mV, g Na,slow = 3, τ Na,slow (V ) = 20 ms. Slow K +current has V 1/2 = −20 mV, k = 5 mV, g K(M) = 20, τ K(M) (V ) = 50 ms.hav<strong>in</strong>g parameters as <strong>in</strong> Fig. 9.33. Notice that these equations are uncoupled and hencecannot oscillate by themselves without the feedback from variable V .Let us describe the burst<strong>in</strong>g mechanism <strong>in</strong> the full I Na,p +I K +I Na,slow + I K(M) -modelwith I = 5. S<strong>in</strong>ce I > 4.51, the rest<strong>in</strong>g state of the fast subsystem does not exist,and the model generates action potentials, depicted <strong>in</strong> Fig. 9.33a. Each spike activatesI Na,slow , produces even more <strong>in</strong>ward current and hence more spikes. This, however,activates a much slower K + current, see Fig. 9.33b, and produces a net outward currentthat moves the fast nullcl<strong>in</strong>e down and eventually term<strong>in</strong>ates spik<strong>in</strong>g. The transitionfrom spik<strong>in</strong>g to rest<strong>in</strong>g occurs via saddle-node on <strong>in</strong>variant circle bifurcation. While atrest, both currents deactivate and the fast nullcl<strong>in</strong>e slowly moves up. The net <strong>in</strong>wardcurrent, consist<strong>in</strong>g mostly of the <strong>in</strong>jected dc-current I = 5, drives the fast subsystemvia the same saddle-node on <strong>in</strong>variant circle bifurcation and <strong>in</strong>itiates another burst, asshown <strong>in</strong> Fig. 9.33a.


Burst<strong>in</strong>g 373fast variable, v10864200s<strong>in</strong>spik<strong>in</strong>gunstableatollstablerest<strong>in</strong>gphase of slow oscillation,Figure 9.34: “Circle/circle” burst<strong>in</strong>g <strong>in</strong>the Ermentrout-Kopell canonical model(9.8) with r(ψ) = s<strong>in</strong> ψ and ω = 0.1.The fast variable fires spikes whiles<strong>in</strong> ψ > 0 and quiescent while s<strong>in</strong> ψ < 0.Shaded atoll is surrounded by theequilibria curves ± √ | s<strong>in</strong> ψ|. The fastsubsystem undergoes saddle-node on <strong>in</strong>variantcircle bifurcation when s<strong>in</strong> ψ = 0.Us<strong>in</strong>g the averag<strong>in</strong>g technique described above <strong>in</strong> Sect. 9.2.3, we can reduce the fourdimensionalI Na,p + I K + I Na,slow + I K(M) -model to a simpler, two-dimensional slowI Na,slow +I K(M) -subsystem of the form (9.4). Burst<strong>in</strong>g of the full model corresponds toa limit cycle attractor of the averaged slow subsystem depicted as a bold curve onthe (m slow , n slow ) plane <strong>in</strong> Fig. 9.33c. Superimposed is the projection of the burst<strong>in</strong>gsolution of the full system (th<strong>in</strong> wobbly curve). In Fig. 9.33d we project a fourdimensionalburst<strong>in</strong>g trajectory onto the three-dimensional subspace (V, m slow , n slow ).The I Na,p +I K +I Na,slow +I K(M) -model <strong>in</strong> Fig. 9.33 enjoys a remarkable property: Itgenerates slow-wave bursts even though its slow I Na,slow + I K(M) -subsystem consistsof two uncoupled equations, and hence cannot oscillate by itself! Another exampleof this phenomenon is presented <strong>in</strong> Ex. 12. Thus, the slow wave that drives the fastI Na,p +I K -subsystem through the two circle bifurcations is not autonomous: it needsa feedback from V . In particular, the oscillation would disappear <strong>in</strong> a voltage-clampexperiment, i.e., when the membrane potential is fixed.Now consider a “circle/circle” burster with a slow subsystem perform<strong>in</strong>g smallamplitudeoscillations so that the fast subsystem is always near the saddle-node on<strong>in</strong>variant circle bifurcation. If the slow subsystem has an autonomous limit cycle attractorthat exists without feedback from V , then such a burster can be reduced to theErmentrout-Kopell (1986) canonical model˙v = v 2 + r(ψ) , if v = +∞, then v = −1, (9.8)˙ψ = ω ,which was orig<strong>in</strong>ally written <strong>in</strong> the ϑ-form; see Ex. 13. Here, ψ is the phase of autonomousoscillation of the slow subsystem, ω ≈ 0 is the frequency of slow oscillation,and r(ψ) is a periodic function that changes sign and slowly drives the fast quadratic<strong>in</strong>tegrate-and-fire neuron (9.8) back and forth through the bifurcation, as we illustrate<strong>in</strong> Fig. 9.34.Alternatively, suppose that the slow subsystem cannot have susta<strong>in</strong>ed oscillationswithout the fast subsystem, i.e., the slow subsystem has a stable equilibrium if v isfixed. In Ex. 17 we prove that there is a piece-wise cont<strong>in</strong>uous change of variables thattransforms any such “circle/circle” burster <strong>in</strong>to one of the two canonical models below,


374 Burst<strong>in</strong>gdepend<strong>in</strong>g on the type of the equilibrium. If the equilibrium of the slow subsystem isa stable node, then the canonical model has the form˙v = I + v 2 + u 1 − u 2 ,˙u 1 = −µ 1 u 1 ,˙u 2 = −µ 2 u 2 .(9.9)If the equilibrium of the slow subsystem is a stable focus, the canonical model has theform˙v = I + v 2 + u 1 ,˙u 1 = −µ 1 u 2 ,(9.10)˙u 2 = −µ 2 (u 2 − u 1 ) ,with µ 2 < 4µ 1 . In both cases, there is an after-spike resett<strong>in</strong>g:if v = +∞, then v ← −1, and (u 1 , u 2 ) ← (u 1 , u 2 ) + (d 1 , d 2 ).Similarly to (9.7), the variable v ∈ R is the re-scaled membrane potential of the neuron.The positive feedback variable u 1 ∈ R describes activation of slow amplify<strong>in</strong>g currentsor potential at a dendritic compartment, whereas the negative feedback variable u 2 ∈ Rdescribes activation of slow resonant currents. I, d 1 , d 2 , and µ 1 , µ 2 ≪ 1 are parameters.When µ 2 > 4µ 1 , the equilibrium of the slow subsystem <strong>in</strong> (9.10) is a stable node,so (9.10) can be transformed <strong>in</strong>to (9.9) by a l<strong>in</strong>ear change of slow variables. If d 1 = 0,then u 1 → 0 and (9.9) is equivalent to (9.7).Both canonical models above exhibit “circle/circle” slow-wave burst<strong>in</strong>g, as depicted<strong>in</strong> Fig. 9.35. When I > 0, the equilibrium of the slow subsystem is <strong>in</strong> the shadedarea correspond<strong>in</strong>g to spik<strong>in</strong>g dynamics of the fast subsystem. When the slow vector(u 1 , u 2 ) enters the shaded area, the fast subsystem fires spikes, prevents the vectorfrom converg<strong>in</strong>g to the equilibrium, and eventually pushes it out of the area. Whileoutside, the vector follows the curved trajectory of the l<strong>in</strong>ear slow subsystem and thenreenters the shaded area aga<strong>in</strong>. Such a slow wave oscillation corresponds to the thicklimit cycle attractor <strong>in</strong> Fig. 9.35, which looks remarkably similar to the one for theI Na,p +I K +I Na,slow +I K(M) -model <strong>in</strong> Fig. 9.33.9.3.3 subHopf/fold cycleWhen the rest<strong>in</strong>g state loses stability via subcritical Andronov-Hopf bifurcation, andthe spik<strong>in</strong>g state disappears via fold limit cycle bifurcation, the burster is said to be ofthe “subHopf/fold cycle” type depicted <strong>in</strong> Fig. 9.36. Because there is a co-existence ofrest<strong>in</strong>g and spik<strong>in</strong>g states, such burst<strong>in</strong>g usually occurs via a hysteresis loop with onlyone slow variable.This k<strong>in</strong>d of burst<strong>in</strong>g was one of the three basic types identified by R<strong>in</strong>zel (1987). Itwas called “elliptic” <strong>in</strong> earlier studies because the profile of oscillation of the membranepotential resembles ellipses, or at least half-ellipses; see Fig. 9.37. Rodent trigem<strong>in</strong>al<strong>in</strong>terneurons <strong>in</strong> Fig. 9.38, dorsal root ganglion and mesV neurons <strong>in</strong> Fig. 9.39 are all


Burst<strong>in</strong>g 37510stable node4fast variable, vslow variables86420-24resetslow variable, u220rest<strong>in</strong>gu 2 =I+u 1equilibirum ofslow subsystemspik<strong>in</strong>g2 u 20 u 1-20 2 40 50 100 150 200 250timeslow variable, u 1108stable focus4rest<strong>in</strong>gspik<strong>in</strong>gfast variable, v6420resetslow variable, u220slow variablesu22u 10-20 20 40 60 80time-2u 1 =-I-4 -2 0 2 4slow variable, u 1Figure 9.35: “Circle/circle” burst<strong>in</strong>g <strong>in</strong> the canonical models (9.9) (top, parameters:I = 1, µ 1 = 0.1, µ 2 = 0.02, d 1 = 1, d 2 = 0.5) and (9.10) (bottom, parameters:I = 1, µ 1 = 0.2, µ 2 = 0.1, d 1 = d 2 = 0.5).“subHopf/fold cycle” bursters, yet the burst<strong>in</strong>g profiles do not look like ellipses. Manymodels of “subHopf/fold cycle” bursters do not generate elliptic profiles either, hencereferr<strong>in</strong>g to this type of burst<strong>in</strong>g by its shape is mislead<strong>in</strong>g and should be avoided.It is quite easy to transform the I Na,p +I K -model <strong>in</strong>to a “subHopf/fold cycle” burster.First, we chose the parameters of the model as <strong>in</strong> Fig. 6.16 so that the phase portraitdepicted <strong>in</strong> Fig. 9.40 is the same as <strong>in</strong> Fig. 9.36, bottom. The co-existence of the stableequilibrium, an unstable limit cycle and a stable limit cycle is essential for produc<strong>in</strong>g thehysteresis loop oscillation. Then, we add a slow K + M-current that activates while thefast subsystem fires spikes and deactivates while it is rest<strong>in</strong>g. Such a resonant currentprovides a negative feedback to the fast subsystem, and the full I Na,p +I K +I K(M) -modelexhibits “subHopf/fold cycle” burst<strong>in</strong>g, shown <strong>in</strong> Fig. 9.41.As <strong>in</strong> the previous examples, the burster <strong>in</strong> this figure is conditional: It needs an<strong>in</strong>jection of a dc-current I, so that the equilibrium correspond<strong>in</strong>g to the rest<strong>in</strong>g state of


376 Burst<strong>in</strong>gsubcriticalAndronov-Hopfbifurcationfold limitcycle bifurcationsubcriticalAndronov-Hopfbifurcationfold limit cyclebifurcationFigure 9.36: “SubHopf/fold cycle” burster: The middle equilibrium correspond<strong>in</strong>g tothe rest<strong>in</strong>g state loses stability via subcritical Andronov-Hopf bifurcation, and the outerlimit cycle attractor correspond<strong>in</strong>g to repetitive spik<strong>in</strong>g disappears via fold limit cyclebifurcation. Top two images are different views of the same 3-D structure.r1r1-1-10 uslow passageeffect0 ur(t)Re z(t)r(t)Re z(t)Figure 9.37: Phase portrait and solution of the canonical model (9.11) for µ = 0.1, ω =3, and a = 0.25 (top) and a = 0.8 (bottom), where r = |z| is the amplitude of oscillation(modified from Izhikevich 2000).


Burst<strong>in</strong>g 37710 mV-39 mVFigure 9.38: Putative “subHopf/fold cycle” burst<strong>in</strong>g <strong>in</strong> rodent trigem<strong>in</strong>al neurons(modified from Del Negro et al. 1998).Figure 9.39: Putative “subHopf/fold cycle” burst<strong>in</strong>g <strong>in</strong> (a) <strong>in</strong>jured dorsal root ganglion(data modified from Jian et al. 2004) and <strong>in</strong> (b) rat mesencephalic layer 5 neurons.1K+ activation variable, n0.5stablelimit cycleunstable limitcyclen-nullcl<strong>in</strong>estable equilibrium (rest)0-80 -70 -60 -50 -40 -30 -20 -10 0 10membrane voltage, V (mV)V-nullcl<strong>in</strong>eFigure 9.40: Phase portrait ofthe I Na,p +I K -model with parameterscorrespond<strong>in</strong>g to subcriticalAndronov-Hopf bifurcation andfold limit cycle bifurcation.


378 Burst<strong>in</strong>g(a)membrane potential, V (mV)(b)0-20-40-60fold cyclesubHopfdelayed transition-800 50 100 150 200 250 300time (ms)membrane potential, V (mV)(c)0-20-40-60-80(d)limit cycle (max)delayedtransitionsubHopflimit cycle (m<strong>in</strong>)fold cycle0 0.05 0.1 0.15slow K + activation gate, n slowslow K + activation gate, nslow0.20.150.10.05fold limit cyclebifurcationsubcriticalAndronov-Hopfbifurcation00 50 100 150 200 250 300time (ms)membrane potential, V (mV)0-20-40-60-800slow K + activationgate, n slow0.10.2 00.5fast K + activationgate, n1Figure 9.41: “SubHopf/fold cycle” burst<strong>in</strong>g <strong>in</strong> the I Na,p +I K +I K(M) -model. Parametersof the fast I Na,p +I K -subsystem are the same as <strong>in</strong> Fig. 6.16 with I = 55. Slow K +M-current has V 1/2 = −20 mV, k = 5 mV, τ(V ) = 60 ms and g K(M) = 1.5.the fast subsystem is unstable. If the subsystem is near such an equilibrium, it slowlydiverges from the equilibrium and jumps to the large-amplitude limit cycle attractorcorrespond<strong>in</strong>g to spik<strong>in</strong>g behavior, as one can see <strong>in</strong> Fig. 9.41a. Each spike activatesslow K + M-current, see Fig. 9.41b, and results <strong>in</strong> build-up of a net outward current thatmakes the fast subsystem less and less excitable. Geometrically, the large-amplitudelimit cycle attractor is approached by a smaller amplitude unstable limit cycle, theycoalesce, and annihilate each other via fold limit cycle bifurcation at n slow ≈ 0.14, seeFig. 9.41c. The trajectory jumps to the stable equilibrium correspond<strong>in</strong>g to the rest<strong>in</strong>gstate. At this moment, the slow K + current starts to deactivate and the net outwardcurrent decreases. S<strong>in</strong>ce the activation gate n slow moves <strong>in</strong> the opposite direction, thefold limit cycle bifurcation gives birth to large-amplitude stable and unstable limitcycles, but the trajectory rema<strong>in</strong>s on the steady-state branch. The unstable limitcycle slowly shr<strong>in</strong>ks and makes the rest<strong>in</strong>g equilibrium lose stability via subcriticalAndronov-Hopf bifurcation. Once the rest<strong>in</strong>g state becomes unstable, the trajectorydiverges from it and jumps back to the large-amplitude limit cycle, thereby clos<strong>in</strong>g thehysteresis loop.


Burst<strong>in</strong>g 379stable equilibriumBaut<strong>in</strong>bifurcationunstalimitcyclestable limit cycle|V|unstable equilibriumusupercritical Andronov-HopfbifurcationHopf/fold cycleburst<strong>in</strong>gHopfHopf/Hopfburst<strong>in</strong>gsubHopf/Hopfburst<strong>in</strong>gfold limit cyclebifurcationsubcriticalAndronov-HopfbifurcationsubHopf/fold cycleburst<strong>in</strong>gfold cyclesubHopfFigure 9.42: A neural system near co-dimension-2 Baut<strong>in</strong> bifurcation (central dot) canexhibit 4 different types of fast-slow burst<strong>in</strong>g, depend<strong>in</strong>g on the trajectory of the slowvariable u ∈ R 2 <strong>in</strong> the parameter space. The “subHopf/fold cycle” burst<strong>in</strong>g occurs viaa hysteresis loop and requites only one slow variable. Solid (dotted) l<strong>in</strong>es correspondto spik<strong>in</strong>g (rest<strong>in</strong>g) regimes (modified from Izhikevich 2000).A prom<strong>in</strong>ent feature of “subHopf/fold cycle” burst<strong>in</strong>g, as well as any other typeof fast-slow burst<strong>in</strong>g <strong>in</strong>volv<strong>in</strong>g Andronov-Hopf bifurcation (“subHopf/*” or “Hopf/*”,where the wildcard “*” means any bifurcation) is that the transition from rest<strong>in</strong>g tospik<strong>in</strong>g does not occur at the moment the rest<strong>in</strong>g state becomes unstable. The fastsubsystem cont<strong>in</strong>ues to dwell at the unstable equilibrium for quite some time before itjumps rather abruptly to a spik<strong>in</strong>g state, as we can clearly see <strong>in</strong> Fig. 9.41. This delayedtransition is due to the slow passage through Andronov-Hopf bifurcation, discussed <strong>in</strong>Sect. 6.1.4. Delayed transitions through Andronov-Hopf bifurcation are ubiquitous <strong>in</strong>neuronal models, but they have never been seen <strong>in</strong> real neurons. Conductance noise,always present at physiological temperatures, constantly kicks the membrane potentialaway from the stable equilibrium, as one can see <strong>in</strong> the <strong>in</strong>set <strong>in</strong> Fig. 9.38, so transitionto spik<strong>in</strong>g <strong>in</strong> real neurons is never delayed. Instead, it can occur even before theequilibrium becomes unstable, as we show <strong>in</strong> Sect. 6.1.4.Suppose that the hysteresis loop oscillation of the slow variable has a small amplitude.That is, the subcritical Andronov-Hopf bifurcation and the fold limit cycle


spik<strong>in</strong>g380 Burst<strong>in</strong>gfoldbifurcationsaddlehomocl<strong>in</strong>icorbitbifurcationfold limit cyclebifurcationxurestfoldsaddlehomocl<strong>in</strong>ic orbitfold limitcycleFigure 9.43: “Fold/fold cycle” burst<strong>in</strong>g: The rest<strong>in</strong>g state disappears via saddle-node(fold) bifurcation and the spik<strong>in</strong>g limit cycle disappears via fold limit cycle bifurcation(modified from Izhikevich 2000).bifurcation of the fast subsystem <strong>in</strong> (9.1) occur for nearby values of the parameter u.In this case, the fast subsystem is near a co-dimension-2 Baut<strong>in</strong> bifurcation, whichwas studied <strong>in</strong> Sect. 6.3.5. Its two-parameter unfold<strong>in</strong>g is depicted <strong>in</strong> Fig. 9.42, left.“SubHopf/fold cycle” burst<strong>in</strong>g occurs when the bifurcation parameter, be<strong>in</strong>g a slowvariable, oscillates between the rest<strong>in</strong>g and spik<strong>in</strong>g regions through the shaded region.Due to the bistability, the parameter could be one-dimensional. Other trajectories ofthe slow parameter correspond to other types of burst<strong>in</strong>g shown <strong>in</strong> Fig. 9.42, right.If the slow variable has an equilibrium near the Baut<strong>in</strong> bifurcation po<strong>in</strong>t, then thefast-slow burster (9.1) can be transformed <strong>in</strong>to the follow<strong>in</strong>g canonical “2+1” modelby a cont<strong>in</strong>uous change of variablesż = (u + iω)z + 2z|z| 2 − z|z| 4 ,˙u = µ(a − |z| 2 ) ,(9.11)where z ∈ C and u ∈ R are the canonical fast and slow variables, respectively, and a, ωand µ ≪ 1 are parameters. In Ex. 14 we show that the model exhibits hysteresis loopperiodic po<strong>in</strong>t-cycle burst<strong>in</strong>g behavior depicted <strong>in</strong> Fig. 9.37 when 0 < a < 1.


Burst<strong>in</strong>g 3819.3.4 fold/fold cycleWhen the stable equilibrium correspond<strong>in</strong>g to the rest<strong>in</strong>g state disappears via saddlenode(fold) bifurcation and the limit cycle attractor correspond<strong>in</strong>g to the spik<strong>in</strong>g statedisappears via fold limit cycle bifurcation, the burster is said to be of the “fold/foldcycle” type, as <strong>in</strong> Fig. 9.43. This type was first discovered <strong>in</strong> the Chay-Cook (1988)model of a pancreatic β-cell by Bertram et al. (1995), who referred to it as be<strong>in</strong>g TypeIV burst<strong>in</strong>g (the three bursters we considered so far were referred to as be<strong>in</strong>g TypeI, II, and III, respectively). S<strong>in</strong>ce both bifurcations result <strong>in</strong> a co-existence of rest<strong>in</strong>gand spik<strong>in</strong>g states, the “fold/fold cycle” burst<strong>in</strong>g can occur via a hysteresis loop <strong>in</strong> a“2+1” system.An <strong>in</strong>terest<strong>in</strong>g geometrical feature of the “fold/fold cycle” burst<strong>in</strong>g is that thereis an unstable limit cycle that appears <strong>in</strong> the middle of a burst and that participates<strong>in</strong> the “fold cycle” bifurcation to term<strong>in</strong>ate the burst. The cycle appears via saddlehomocl<strong>in</strong>ic orbit bifurcation <strong>in</strong> Fig. 9.43, but other scenarios are possible too. It is agood exercise of one’s geometrical <strong>in</strong>tuition and understand<strong>in</strong>g of the fast-slow burst<strong>in</strong>gmechanisms to come up with alternative scenarios of the “fold/fold cycle” burst<strong>in</strong>g.For example, consider the case of the unstable limit cycle be<strong>in</strong>g <strong>in</strong>side the stable one.9.3.5 fold/HopfWhen the stable equilibrium correspond<strong>in</strong>g to the rest<strong>in</strong>g state disappears via saddlenode(fold) bifurcation and the limit cycle attractor correspond<strong>in</strong>g to the spik<strong>in</strong>g stateshr<strong>in</strong>ks to a po<strong>in</strong>t via supercritical Andronov-Hopf bifurcation, the burster is said to beof the “fold/Hopf” type; see Fig. 9.44. This type of burst<strong>in</strong>g, called “tapered” <strong>in</strong> someearlier studies, was found <strong>in</strong> models of <strong>in</strong>sul<strong>in</strong>-produc<strong>in</strong>g pancreatic β-cells (Smolen etal. 1993, Pernarowski 1994) and <strong>in</strong> models of certa<strong>in</strong> enzymatic systems (Holden andErneux 1993a,b).As one can see <strong>in</strong> the figure, the fast subsystem undergoes two bifurcations whileit is <strong>in</strong> the excited state: One corresponds to the term<strong>in</strong>ation of repetitive spik<strong>in</strong>gvia supercritical Andronov-Hopf bifurcation, and the other one corresponds to thetransition from the excited equilibrium to rest<strong>in</strong>g equilibrium via saddle-node (fold)bifurcation. The first bifurcation, i.e., bifurcation of a spik<strong>in</strong>g limit cycle attractor,determ<strong>in</strong>es the topological type of burst<strong>in</strong>g. The second bifurcation is essential forthe “fold/fold” hysteresis loop, and it only determ<strong>in</strong>es the subtype of the “fold/Hopf”burst<strong>in</strong>g. Us<strong>in</strong>g ideas described <strong>in</strong> Ex. 19, one can come up with another subtype of“fold/Hopf” burster hav<strong>in</strong>g “fold/subHopf” hysteresis loop.9.3.6 fold/circleWhen the stable equilibrium correspond<strong>in</strong>g to the rest<strong>in</strong>g state disappears via saddlenode(fold) bifurcation and the limit cycle attractor correspond<strong>in</strong>g to the spik<strong>in</strong>g statedisappears via saddle-node on <strong>in</strong>variant circle bifurcation, the burster is said to be ofthe “fold/circle” type, as <strong>in</strong> Fig. 9.45. This type was first discovered <strong>in</strong> the model of


382 Burst<strong>in</strong>gspik<strong>in</strong>gsupercriticalAndronov-Hopfbifurcationfoldbifurcationfoldbifurcationxurest<strong>in</strong>gfoldsupercriticalAndronov-HopffoldFigure 9.44: “Fold/Hopf” burst<strong>in</strong>g: The rest<strong>in</strong>g state disappears via saddle-node (fold)bifurcation and the spik<strong>in</strong>g limit cycle shr<strong>in</strong>ks to a po<strong>in</strong>t via supercritical Andronov-Hopf bifurcation (modified from Izhikevich 2000).thalamo-cortical relay neuron by Rush and R<strong>in</strong>zel (1994), and it was called “triangular”<strong>in</strong> earlier studies (Wang and R<strong>in</strong>zel 1995) because of the shape of the voltage envelope.As one can see <strong>in</strong> the figure, the fast subsystem can have five equilibria, two ofwhich are stable nodes. This is a consequence of the qu<strong>in</strong>tic shape of the V -nullcl<strong>in</strong>eof the fast subsystem. While the trajectory is at the lower equilibrium, the V -nullcl<strong>in</strong>emoves up, the equilibrium disappears via fold bifurcation, and the fast subsystemstarts to fire spikes. Dur<strong>in</strong>g this active period, the V -nullcl<strong>in</strong>e slowly moves down,and the spik<strong>in</strong>g limit cycle disappears via saddle-node on <strong>in</strong>variant circle bifurcation.The fast subsystem, however, is at the second stable equilibrium correspond<strong>in</strong>g to adepolarized state. The slow V -nullcl<strong>in</strong>e cont<strong>in</strong>ues to move down, and this equilibriumdisappears via another fold bifurcation, thereby clos<strong>in</strong>g the “fold/fold” hysteresis loop.Alternatively, “fold/circle” burst<strong>in</strong>g can be of the slow-wave type depicted <strong>in</strong> Fig. 9.28hav<strong>in</strong>g only three equilibria. The slow subsystem needs to be at least two-dimensional<strong>in</strong> this case, though.


spik<strong>in</strong>grest<strong>in</strong>gBurst<strong>in</strong>g 383saddle-node on<strong>in</strong>variant circlebifurcationfoldbifurcationxufoldbifurcationSaddle-Node onInvariant CircleFoldFoldFigure 9.45: “Fold/circle” burst<strong>in</strong>g: The rest<strong>in</strong>g state disappears via fold bifurcationand the spik<strong>in</strong>g state disappears via saddle-node on <strong>in</strong>variant circle bifurcation (modifiedfrom Izhikevich 2000).9.4 Neuro-Computational PropertiesThere is more to the topological classification of bursters than just a mathematicalexercise. Indeed, <strong>in</strong> Chap. 7 we have shown that the neuro-computational propertiesof an excitable system depend on the type of bifurcation of the rest<strong>in</strong>g state. Thesame is valid for a burster: Its neuro-computational properties depend on the k<strong>in</strong>d ofbifurcations of the rest<strong>in</strong>g and spik<strong>in</strong>g states, that is, on the burster’s type. Know<strong>in</strong>gthe topological type of a given burst<strong>in</strong>g neuron, we know what the neuron can doand more importantly what it cannot do, regardless of the model that describes itsdynamics.9.4.1 How to dist<strong>in</strong>guish?First, we stress that the topological classification of bursters provided <strong>in</strong> the previoussection is def<strong>in</strong>ed for mathematical models, and not for real neurons. Moreover, the


384 Burst<strong>in</strong>gaction potentials cut2 mV100 msFigure 9.46: The conductance noise destabilizes the focusequilibrium <strong>in</strong> a mesencephalic V neuron before subcriticalAndronov-Hopf bifurcation takes place, therebygiv<strong>in</strong>g an impression of a supercritical Andronov-Hopfbifurcation (data modified from Wu et al. 2001).types are def<strong>in</strong>ed for models of the fast-slow form (9.1) assum<strong>in</strong>g that the ratio of timescales, µ, is sufficiently small. Not all neurons can be described adequately by suchmodels, hence extend<strong>in</strong>g the classification to those neurons may be worthless. Typicalexample when the classification fails is the model of burst<strong>in</strong>g of the sensory process<strong>in</strong>gneuron <strong>in</strong> weakly electric fish, known as the “ghostburster” (Doiron et al. 2002) <strong>in</strong>which µ > 0.1.If a burst<strong>in</strong>g neuron can be described accurately by a model hav<strong>in</strong>g fast-slow form(9.1), then there is no problem to determ<strong>in</strong>e its topological type — just freeze theslow subsystem, i.e., set µ = 0, and f<strong>in</strong>d bifurcations of the fast subsystem treat<strong>in</strong>gu as a parameter. Software packages, such as XPPAUT, AUTO, or MATLAB basedMATCONT, are helpful <strong>in</strong> bifurcation analyses of such systems.What if a neuron has an apparent fast-slow dynamics but its model is not knownat present? To determ<strong>in</strong>e the types of bifurcations of the fast subsystem, we firstuse non-<strong>in</strong>vasive observations: presence or absence of fast subthreshold oscillations,changes <strong>in</strong> <strong>in</strong>traburst (<strong>in</strong>terspike) frequency, changes <strong>in</strong> spike amplitudes, etc. Eachpiece of <strong>in</strong>formation excludes some bifurcations and narrows the set of possible typesof burst<strong>in</strong>g. Then, we can use <strong>in</strong>vasive methods, e.g., small perturbations, to test theco-existence of rest<strong>in</strong>g and spik<strong>in</strong>g states, and narrow down the choice of bifurcationsfurther. With some luck, we can exclude sufficiently many bifurcations and determ<strong>in</strong>eexactly the type of burst<strong>in</strong>g without even know<strong>in</strong>g the details of the mathematicalmodel that describes it.9.4.2 Integrators vs. ResonatorsA conspicuous feature of neuronal systems near Andronov-Hopf bifurcation, whethersubcritical or supercritical, is the existence of fast subthreshold oscillations of themembrane potential. Quite often, these oscillations are visible <strong>in</strong> record<strong>in</strong>gs of themembrane potential. If not, then they could be evoked by a brief small pulse ofcurrent. Apparently, a burst<strong>in</strong>g neuron exhibit<strong>in</strong>g such oscillations <strong>in</strong> the quiescentstate is either of the “Hopf/*” type of “subHopf/*” type, where the wildcard “*”denotes any appropriate bifurcation of the spik<strong>in</strong>g state. All such bursters are <strong>in</strong> thelower half of the table <strong>in</strong> Fig. 9.23.To discern whether the bifurcation is supercritical or subcritical, one needs to study


Burst<strong>in</strong>g 385bifurcations of limit cyclessaddle-nodeon <strong>in</strong>variantcirclesaddlehomocl<strong>in</strong>icorbitsupercriticalAndronov-Hopffoldlimitcyclesaddle-node(fold)fold/circlefold/homocl<strong>in</strong>icfold/Hopffold/fold cyclebifurcations of equilibriasaddle-nodeon <strong>in</strong>variantcirclesupercriticalAndronov-HopfsubcriticalAndronov-Hopfcircle/circleHopf/circlesubHopf/circlecircle/homocl<strong>in</strong>icHopf/homocl<strong>in</strong>icsubHopf/homocl<strong>in</strong>iccircle/HopfHopf/HopfsubHopf/Hopfcircle/fold cycleHopf/fold cyclesubHopf/fold cyclebistabilitybeforethe burstbistabilityat the endof the burstFigure 9.47: Bistability, i.e., co-existence of rest<strong>in</strong>g and spik<strong>in</strong>g states, depends on thetopological type of burst<strong>in</strong>g.the amplitude of emerg<strong>in</strong>g oscillations, which could be tricky. In models, slow passagethrough supercritical Andronov-Hopf bifurcations often results <strong>in</strong> a delayed transitionto oscillations with an <strong>in</strong>termediate or large amplitude, hence such a bifurcation maylook like subcritical. In record<strong>in</strong>gs, like the one <strong>in</strong> Fig. 9.39a or <strong>in</strong> Fig. 9.46, noisedestabilizes the focus equilibrium before the subcritical Andronov-Hopf bifurcationtakes place and gives the impression that the amplitude <strong>in</strong>creases gradually, i.e., as ifthe bifurcation were supercritical.The existence of fast subthreshold oscillations <strong>in</strong>dicates that the burst<strong>in</strong>g neuronacts as a resonator, at least right before the onset of a burst. In Sect. 7.2.2 we showedthat such neurons prefer a certa<strong>in</strong> resonant frequency of stimulation that matches thefrequency of subthreshold oscillations. A resonant <strong>in</strong>put may excite the neuron and<strong>in</strong>itiate a burst or it may delay the transition to the burst, depend<strong>in</strong>g on its phaserelative to the phase of subthreshold oscillations.In contrast, all bursters <strong>in</strong> the upper half of the table <strong>in</strong> Fig. 9.23, i.e., “fold/*” and“circle/*” types, do not have fast subthreshold oscillations, at least before the onsetof each burst (see Ex. 5). The fast subsystem of such bursters acts as an <strong>in</strong>tegrator:It prefers high-frequency <strong>in</strong>puts; the higher the frequency, the sooner the transition tothe spik<strong>in</strong>g state. The phase of the <strong>in</strong>put does not play any role here.


386 Burst<strong>in</strong>g9.4.3 BistabilitySuppose the transition from rest<strong>in</strong>g to spik<strong>in</strong>g state occurs via saddle-node bifurcation(off an <strong>in</strong>variant circle) or subcritical Andronov-Hopf bifurcation of the fast subsystem,as <strong>in</strong> Fig. 6.46. In these cases, the trajectory jumps to a pre-exist<strong>in</strong>g limit cycleattractor correspond<strong>in</strong>g to the spik<strong>in</strong>g state, not shown <strong>in</strong> the figure. In contrast,saddle-node on <strong>in</strong>variant circle bifurcation or supercritical Andronov-Hopf bifurcationcreates such a limit cycle attractor. Thus, there must be a co-existence of stable rest<strong>in</strong>gand stable spik<strong>in</strong>g states <strong>in</strong> the former case, but not necessarily <strong>in</strong> the latter case. Thissimple observation has far reach<strong>in</strong>g consequences described below. In particular, itimplies that all “fold/*” and “subHopf/*” bursters exhibit bistability, at least beforethe onset of a burst, while “circle/*” and “Hopf/*” bursters may not; see Fig. 9.47.Similarly, if the transition from spik<strong>in</strong>g to rest<strong>in</strong>g state of the fast subsystem occursvia saddle homocl<strong>in</strong>ic orbit bifurcation or fold limit cycle bifurcation, then there is apre-exist<strong>in</strong>g stable equilibrium, and hence a co-existence of attractors. Thus, “*/homocl<strong>in</strong>ic”and “*/fold cycle” bursters also exhibit bistability, at least at the end of aburst, while “*/circle” and “*/Hopf” bursters may not, as we summarize <strong>in</strong> Fig. 9.47.An obvious consequence of bistability is that an appropriate stimulus can switch thesystem from rest<strong>in</strong>g to spik<strong>in</strong>g and back. We illustrate this phenomenon <strong>in</strong> Fig. 9.48us<strong>in</strong>g the I Na,p +I K +I K(M) -model, which exhibits a hysteresis loop “fold/homocl<strong>in</strong>ic”burst<strong>in</strong>g when I = 5. All three simulations <strong>in</strong> the figure start with the same <strong>in</strong>itialconditions. In Fig. 9.48b we apply a brief pulse of current while the fast subsystem is atthe rest<strong>in</strong>g state. This stimulation pushes the membrane potential over the thresholdstate <strong>in</strong>to the attraction doma<strong>in</strong> of the spik<strong>in</strong>g limit cycle of the fast subsystem, therebyevok<strong>in</strong>g a burst.Notice that the evoked burst is one spike shorter than the control one <strong>in</strong> Fig. 9.48a.This is expected, s<strong>in</strong>ce the K + M-current did not have enough time to recover fromthe previous burst (not shown <strong>in</strong> the figure), therefore, there is a residual outwardcurrent that shortens the active phase. From the geometrical po<strong>in</strong>t of view, this occursbecause the transition to the spik<strong>in</strong>g manifold <strong>in</strong> Fig. 9.48b, right, occurs before theslow variable reaches the fold knee, hence the distance to the homocl<strong>in</strong>ic bifurcation isshorter. An <strong>in</strong>terest<strong>in</strong>g observation is that the first spike <strong>in</strong> the evoked burst actuallycorresponds to the second spike <strong>in</strong> the control burst <strong>in</strong> Fig. 9.48a. The earlier thestimulation acts, the sooner the trajectory jumps to the spik<strong>in</strong>g manifold and thefewer spikes the evoked burst has.In Fig. 9.48c we applied a brief pulse of current <strong>in</strong> the middle of a burst to switchthe system to the rest<strong>in</strong>g state. Notice that the quiescent period, i.e., the time periodto the second burst, is shorter than the control one <strong>in</strong> Fig. 9.48a or <strong>in</strong> Fig. 9.48b. Thisis also to be expected, s<strong>in</strong>ce the K + M- current was not fully activated dur<strong>in</strong>g the<strong>in</strong>terrupted burst, therefore it does not need that much time to deactivate dur<strong>in</strong>g therest<strong>in</strong>g period. Geometrically, the short duration of the rest<strong>in</strong>g phase is a consequenceof the distance the slow variable needs to travel to get to the fold knee be<strong>in</strong>g small.


Burst<strong>in</strong>g 387spik<strong>in</strong>g(a)Vfastnfoldthresholdhomocl<strong>in</strong>icrest<strong>in</strong>gn slow(b)stimulationstimulationstimulation(c)stimulationFigure 9.48: Bistability of rest<strong>in</strong>g and spik<strong>in</strong>g states <strong>in</strong> a “fold/homocl<strong>in</strong>ic” burster. Abrief stimulus can <strong>in</strong>itiate a premature transition to spik<strong>in</strong>g state (b) or to quiescentstate (c). Shown are simulations of the I Na,p +I K +I K(M) -model with parameters as <strong>in</strong>Fig. 9.4b.9.4.4 Bursts as a unit of neuronal <strong>in</strong>formationMammalian neurons may fire bursts to <strong>in</strong>crease the reliability of synaptic transmission(Lisman 1997). Indeed, if a presynaptic neuron sends a burst of spikes <strong>in</strong>stead of a s<strong>in</strong>glespike, then the chances that at least one of them overcomes the synaptic transmissionfailure <strong>in</strong>crease. If two or more spikes go through, then the postsynaptic effect is muchstronger than the one for a s<strong>in</strong>gle spike.In addition, important <strong>in</strong>formation may be carried <strong>in</strong> the <strong>in</strong>traburst frequency.Consider the effect of a burst on a postsynaptic resonator neuron, i.e., a neuron withrest<strong>in</strong>g state near an Andronov-Hopf bifurcation. Such a neuron cares about the frequencycontent of the burst, i.e., whether it is resonant or not, as we discussed <strong>in</strong>Sect. 7.2.2. Some types of bursters have relatively constant <strong>in</strong>traburst (<strong>in</strong>stantaneous<strong>in</strong>terspike) frequencies, as <strong>in</strong> Fig. 9.49b, which may be resonant for some postsynaptic


388 Burst<strong>in</strong>gFigure 9.49: The <strong>in</strong>stantaneous spike frequency of a trigem<strong>in</strong>al motor neuron (a) andtrigem<strong>in</strong>al <strong>in</strong>terneuron (b) of a rodent (modified from Del Negro et al. 1998).bifurcations of limit cyclessaddle-nodeon <strong>in</strong>variantcirclesaddlehomocl<strong>in</strong>icorbitsupercriticalAndronov-Hopffoldlimitcyclesaddle-node(fold)fold/circlefold/homocl<strong>in</strong>icfold/Hopffold/fold cyclebifurcations of equilibriasaddle-nodeon <strong>in</strong>variantcirclesupercriticalAndronov-HopfsubcriticalAndronov-Hopfcircle/circleHopf/circlesubHopf/circlecircle/homocl<strong>in</strong>icHopf/homocl<strong>in</strong>icsubHopf/homocl<strong>in</strong>iccircle/HopfHopf/HopfsubHopf/Hopfcircle/fold cycleHopf/fold cyclesubHopf/fold cycle<strong>in</strong>creas<strong>in</strong>gfrequencyat thebeg<strong>in</strong>n<strong>in</strong>gdecreas<strong>in</strong>gfrequencyat theendFigure 9.50: Topological types of bursters <strong>in</strong> the shaded regions can produce chirpburststhat sweep a frequency range.


Burst<strong>in</strong>g 389spike synchronizationburst synchronizationFigure 9.51: Various regimes of synchronization of bursters.neurons but not for others. In contrast, other topological types of bursters have widelyvary<strong>in</strong>g <strong>in</strong>stantaneous <strong>in</strong>terspike frequencies, as <strong>in</strong> Fig. 9.49a, that scan or sweep abroad frequency range go<strong>in</strong>g all the way to zero.When the bifurcation from rest<strong>in</strong>g to spik<strong>in</strong>g state is of the saddle-node on <strong>in</strong>variantcircle type, i.e., the system is Class 1 excitable, the frequency of emerg<strong>in</strong>g spik<strong>in</strong>g isfirst small, and then <strong>in</strong>creases. Therefore, all “circle/*” bursters generate chirps with<strong>in</strong>stantaneous <strong>in</strong>terspike frequencies <strong>in</strong>creas<strong>in</strong>g from zero to a relatively large value,at least at the beg<strong>in</strong>n<strong>in</strong>g of the burst. Similarly, when the bifurcation of the spik<strong>in</strong>gstate is of the saddle-node on <strong>in</strong>variant circle or saddle homocl<strong>in</strong>ic orbit type, thefrequency of spik<strong>in</strong>g at the end of the burst decreases to zero, so all “*/circle” and“*/homocl<strong>in</strong>ic” bursters also generate chirps, as <strong>in</strong> Fig. 9.49a. In summary, all shadedbursters <strong>in</strong> Fig. 9.50 have sweep<strong>in</strong>g <strong>in</strong>terspike frequencies, so that one part of the burstis resonant for one neuron and another part of the same burst is resonant for anotherneuron.9.4.5 SynchronizationConsider two coupled burst<strong>in</strong>g neurons of the fast-slow type. S<strong>in</strong>ce each burster hastwo times scales, one for rhythmic spik<strong>in</strong>g and one for repetitive burst<strong>in</strong>g, there aretwo synchronization regimes:• Spike synchronization, as <strong>in</strong> Fig. 9.51, left.• Burst synchronization, as <strong>in</strong> Fig. 9.51, right.One of them does not imply the other. Of course, there is an additional regime whenspikes and bursts are synchronized. We will study synchronization phenomena <strong>in</strong> detail<strong>in</strong> Chap. 10; here we just mention how they depend on the topological type of burst<strong>in</strong>g.Let us consider spike synchronization first. S<strong>in</strong>ce we are <strong>in</strong>terested <strong>in</strong> the fast timescale, we neglect the slow variable dynamics for a while and treat two bursters ascoupled oscillators. A necessary condition for synchronization of two weakly coupledoscillators is that they have nearly equal frequencies. How near is “near” dependson the strength of the coupl<strong>in</strong>g. Thus, spike synchronization depends crucially on


est<strong>in</strong>g390 Burst<strong>in</strong>gspikesynchronizationspikede-synchronizationspik<strong>in</strong>gABexcitatory coupl<strong>in</strong>gburster B<strong>in</strong>hibitorystimulationexcitatorystimulationAB<strong>in</strong>hibitory coupl<strong>in</strong>gFigure 9.52: Burst synchronization and de-synchronization of two coupled“fold/homocl<strong>in</strong>ic” bursters (modified from Izhikevich 2000).the <strong>in</strong>stantaneous <strong>in</strong>terspike frequency, which may vary substantially dur<strong>in</strong>g a burst.Indeed, a small perturbation of the slow variable may result <strong>in</strong> large perturbations ofthe <strong>in</strong>terspike frequency <strong>in</strong> any shaded burster <strong>in</strong> Fig. 9.50, hence such a burster wouldbe reluctant to exhibit spike synchronization, unless the coupl<strong>in</strong>g is strong.Study<strong>in</strong>g burst synchronization of weakly coupled neurons <strong>in</strong>volves the same mathematicalmethods as study<strong>in</strong>g synchronization of strongly coupled relaxation oscillators,which we consider <strong>in</strong> detail <strong>in</strong> Chap. 10. The mechanisms of synchronization dependon whether the burst<strong>in</strong>g is of the hysteresis loop type or of the slow wave type, andwhether the rest<strong>in</strong>g state is an <strong>in</strong>tegrator or a resonator.In Fig. 9.52 we illustrate the geometry of burst synchronization of two coupled“fold/homocl<strong>in</strong>ic” bursters of hysteresis loop type. Burster A is slightly ahead ofburster B so that A starts the spik<strong>in</strong>g phase while B is still rest<strong>in</strong>g. If the synapticconnections between the bursters are excitatory, fir<strong>in</strong>g of A causes B to jump to thespik<strong>in</strong>g state prematurely, thereby shorten<strong>in</strong>g the time difference between the bursts.In addition, the evoked burst of B is shorter, which also speeds up the synchronizationprocess. In contrast, when the connections are <strong>in</strong>hibitory, fir<strong>in</strong>g of A delays thetransition of B to the spik<strong>in</strong>g state, thereby <strong>in</strong>creas<strong>in</strong>g the time difference betweenthe bursts and desynchroniz<strong>in</strong>g the bursters. Thus, the “fold/homocl<strong>in</strong>ic” burster behavesaccord<strong>in</strong>g to the pr<strong>in</strong>ciple excitation means synchronization, <strong>in</strong>hibition meansde-synchronization. S<strong>in</strong>ce the <strong>in</strong>stantaneous <strong>in</strong>terspike frequency of “fold/homocl<strong>in</strong>ic”burst<strong>in</strong>g decays to zero, small deviations of the slow variable result <strong>in</strong> large deviationsof the period of oscillation. Typically, the periods of fast oscillations of the two bursterscan slowly diverge from each other. As a result, spikes start synchronized and thende-synchronize dur<strong>in</strong>g the burst, as we <strong>in</strong>dicate <strong>in</strong> the figure.If the burst<strong>in</strong>g neuron is a resonator, i.e., it is of the “Hopf/*” or “subHopf/*” type,then both excitation and <strong>in</strong>hibition may evoke premature spik<strong>in</strong>g, as we have shown


Burst<strong>in</strong>g 391<strong>in</strong> Chap. 7, and lead to burst synchronization. An important feature here is that the<strong>in</strong>terspike frequency of one burster be resonant to the subthreshold oscillations of theother one. We study these and other issues related to synchronization <strong>in</strong> Chap. 10.


392 Burst<strong>in</strong>gReview of Important Concepts• A burst of spikes is a tra<strong>in</strong> of action potentials followed by a periodof quiescence.• Burst<strong>in</strong>g activity typically <strong>in</strong>volves two time scales: fast spik<strong>in</strong>g andslow modulation via a resonant current.• Many mathematical models of bursters have fast-slow formẋ = f(x, u) (fast spik<strong>in</strong>g),˙u = µg(x, u) (slow modulation).• To dissect a burster, one freezes its slow subsystem (i.e., sets µ = 0)and uses the slow variable u as a bifurcation parameter to study thefast subsystem.• The fast subsystem undergoes two important bifurcations dur<strong>in</strong>g aburst: (1) bifurcation of an equilibrium result<strong>in</strong>g <strong>in</strong> transition to spik<strong>in</strong>gstate, and (2) bifurcation of a limit cycle attractor result<strong>in</strong>g <strong>in</strong>transition to rest<strong>in</strong>g state.• Different types of bifurcations result <strong>in</strong> different topological types ofburst<strong>in</strong>g.• There are 16 basic types of burst<strong>in</strong>g, summarized <strong>in</strong> Fig. 9.23.• Different topological types of bursters have different neurocomputationalproperties.Bibliographical NotesThe history of formal classification of burst<strong>in</strong>g starts with the sem<strong>in</strong>al paper by R<strong>in</strong>zel(1987), who contrasted the bifurcation mechanism of the “square-wave”, “parabolic”,and “elliptic” bursters. Then, Bertram et al. (1995) followed R<strong>in</strong>zel’s suggestionand referred to the bursters us<strong>in</strong>g Roman numbers, add<strong>in</strong>g a new, Type IV burster.Another, “tapered” type of burst<strong>in</strong>g was studied simultaneously and <strong>in</strong>dependently byHolden and Erneux (1993a,b), Smolen et al. (1993), and Pernarowski (1994). Laterde Vries (1998) suggested to refer to it as Type V burster. Yet another, “triangular”type of burst<strong>in</strong>g was studied by Rush and R<strong>in</strong>zel (1994), mak<strong>in</strong>g the total number ofidentified bursters to be 6. To honor these pioneers, we described these six classicalbursters <strong>in</strong> the order consisted with the number<strong>in</strong>g nomenclature of Bertram et al.(1995). Their bifurcation mechanisms are summarized <strong>in</strong> Fig. 9.53.


Burst<strong>in</strong>g 393Saddle-Node Saddle Supercritical FoldBifurcations on Invariant Homocl<strong>in</strong>ic Andronov- LimitCircle Orbit Hopf Cycletriangular square-wave taperedFold Type I Type V Type IVSaddle-Nodeon InvariantCircleparabolicType IISupercriticalAndronov-HopfSubcriticalAndronov-HopfellipticType IIIFigure 9.53: Bifurcation mechanisms and classical nomenclature of the 6 burstersknown <strong>in</strong> the XX century. Compare with Fig. 9.23 and Fig. 9.24.The complete classification of bursters was provided by Izhikevich (2000), and itwas actually motivated by the paper of Guckenheimer et al. (1997). There is a drasticdifference between Izhikevich’s approach, and that of the scientists mentioned above.They used a bottom-up approach; that is, they considered biophysically plausible conductancebased models describ<strong>in</strong>g experimentally observable cellular behavior and thenthey determ<strong>in</strong>ed the types of burst<strong>in</strong>g these models exhibited. In contrast, Izhikevich(2000) used the top-down approach and considered all possible pairs of co-dimension-1bifurcations of rest and spik<strong>in</strong>g states, which resulted <strong>in</strong> different types of burst<strong>in</strong>g. Itwas an easy task to provide a conductance-based model exhibit<strong>in</strong>g each burst<strong>in</strong>g type.Thus, many of the bursters are “theoretical” <strong>in</strong> the sense that they have yet to be seen<strong>in</strong> experiments.Interest<strong>in</strong>gly, a challeng<strong>in</strong>g problem was to suggest a nam<strong>in</strong>g scheme for the bursters.The names should be self-explanatory and easy to remember and understand. Thus,the number<strong>in</strong>g scheme suggested by Bertram et al. (1995) would lead, e.g., to burstersof Type XXVII, Type LXIII, Type LCXVI, etc. We cannot use descriptions such as“elliptic”, “parabolic”, “hyperbolic”, “triangular”, “rectangular”, etc., s<strong>in</strong>ce they aremislead<strong>in</strong>g. In this book we follow Izhikevich (2000) and name the bursters accord<strong>in</strong>gto the two bifurcations <strong>in</strong>volved, as <strong>in</strong> Fig. 9.23.Not all bursters can be represented <strong>in</strong> the fast-slow form with a clear separation


394 Burst<strong>in</strong>gspik<strong>in</strong>grest<strong>in</strong>gFigure 9.54: A hedgehog-like limit cycle attractor results <strong>in</strong> burst<strong>in</strong>g dynamics even <strong>in</strong>two-dimensional systems; see Ex. 1 (modified from Hoppensteadt and Izhikevich 1997).of the time scales. Those that cannot are referred to as hedgehog bursters (Izhikevich2000), s<strong>in</strong>ce they have a limit cycle (or a more complicated attractor) with some spikyparts correspond<strong>in</strong>g to repetitive spik<strong>in</strong>g and some smooth parts correspond<strong>in</strong>g toquiescence, as <strong>in</strong> Fig. 9.54. An <strong>in</strong>terest<strong>in</strong>g example of the hedgehog burster is themodel of sensory process<strong>in</strong>g neuron of weakly electric fish (Doiron et al. 2002). Theauthors refer to the model as “ghostburster” because repetitive spik<strong>in</strong>g corresponds toa slow transition of the full system through the ghost of a fold limit cycle attractor.As a dynamical system, the ghostburster is near a co-dimention-2 bifurcation of limitcycle attractor, and it exhibits chaotic dynamics.Betram et al. (1995) noticed that burst<strong>in</strong>g often occurs when the fast subsystemis near a co-dimension-2 bifurcation. Izhikevich (2000) suggested that many simplemodels of bursters could be obta<strong>in</strong>ed by consider<strong>in</strong>g unfold<strong>in</strong>gs of various degeneratebifurcations of high co-dimension (organiz<strong>in</strong>g centers) and treat<strong>in</strong>g the unfold<strong>in</strong>g parametersas slow variables rotat<strong>in</strong>g around the bifurcation po<strong>in</strong>t, as <strong>in</strong> Fig. 9.28 orFig. 9.42. Consider<strong>in</strong>g the Baut<strong>in</strong> bifurcation, Izhikevich (2001) obta<strong>in</strong>ed the canonicalmodel for “elliptic” burster (9.11). Golubitsky et al. (2001) applied this idea toother local bifurcations (spik<strong>in</strong>g with <strong>in</strong>f<strong>in</strong>itesimal amplitude). Global bifurcations areconsidered <strong>in</strong> Ex. 26.Izhikevich and Hoppensteadt (2004) extend the classification of bursters to oneandtwo-dimensional mapp<strong>in</strong>gs, identify<strong>in</strong>g 3 and 20 different classes, respectively. Acollection of chapters “Burst<strong>in</strong>g: The Genesis of Rhythm <strong>in</strong> the Nervous System” editedby Coombes and Bressloff (2005) provides recent developments <strong>in</strong> the field of burst<strong>in</strong>gdynamics.Study<strong>in</strong>g burst<strong>in</strong>g dynamics is still one of the hardest problems <strong>in</strong> applied mathematics.The method of dissection of fast-slow bursters of the form (9.1), pioneeredby R<strong>in</strong>zel (1987), is part of the asymptotic theory of s<strong>in</strong>gularly perturbed dynamicalsystems (Mishchenko et al. 1994). One would expect the theory to suggest other,quantitative methods of analyses of fast-slow bursters. However, the basic assumptionof the theory is that the fast subsystem has only equilibria, e.g., up- and down-states as


Burst<strong>in</strong>g 3951K + activation gate, n0.80.60.40.20-80 -60 -40 -20 0membrane potential, V (mV)Figure 9.55: Burst<strong>in</strong>g <strong>in</strong> two-dimensional I Na,p +I K -model with parameters as <strong>in</strong>Fig. 6.16 and I = 43.<strong>in</strong> the po<strong>in</strong>t-po<strong>in</strong>t hysteresis loops <strong>in</strong> Ex. 19. This assumption is violated when the neuronfires a burst of spikes. Thus, the theory is helpless <strong>in</strong> study<strong>in</strong>g fast-slow po<strong>in</strong>t-cyclebursters. An exception is Pontryag<strong>in</strong>’s problem, which is related to “fold cycle/foldcycle” burst<strong>in</strong>g; see Ex. 21 below and Sect. 7 <strong>in</strong> Mishchenko et al. (1994). Pontryag<strong>in</strong>and Rodyg<strong>in</strong> (1960) pioneered the method of averag<strong>in</strong>g of the fast subsystem, whichwas used <strong>in</strong> the context of bursters by R<strong>in</strong>zel and Lee (1986), Pernarowski et al. (1992),Smolen et al. (1993), Baer et al. (1995). Shilnikov et al. (2005) <strong>in</strong>troduced an averagenullcl<strong>in</strong>e of the slow subsystem, and showed how the averag<strong>in</strong>g method can be used tostudy co-existence of spik<strong>in</strong>g and burst<strong>in</strong>g states <strong>in</strong> a model neuron, and bifurcations<strong>in</strong> bursters <strong>in</strong> general. Some of the transitions “rest<strong>in</strong>g ↔ burst<strong>in</strong>g ↔ tonic spik<strong>in</strong>g”were also considered by Ermentrout and Kopell (1986a), Terman (1991), Destexhe andGaspard (1993), Shilnikov and Cymbalyuk (2004, 2005), and Medvedev (2005).The averag<strong>in</strong>g method, as many other classical methods of analysis of dynamicalsystems, breaks down when the fast subsystem slowly passes a bifurcation po<strong>in</strong>t. Thedevelopment of early dynamical system theory was largely motivated by studies ofperiodic oscillators. It is reasonable to expect that the next major developments ofthis theory will be com<strong>in</strong>g from studies of bursters.Exercises1. (Planar burster) Invent a planar system of ODEs hav<strong>in</strong>g a hedgehog limit cycleattractor, as <strong>in</strong> Fig. 9.54, and capable of exhibit<strong>in</strong>g periodic burst<strong>in</strong>g activity.2. (Noise-<strong>in</strong>duced burst<strong>in</strong>g) Expla<strong>in</strong> why the I Na,p +I K -model with the phase portraitas <strong>in</strong> Fig. 9.55 bursts even though it has only two dimensions.3. (Noise-<strong>in</strong>duced burst<strong>in</strong>g) Explore numerically the I Na,p +I K -model with phaseportrait as <strong>in</strong> Fig. 6.7, top, and make it burst as <strong>in</strong> Fig. 9.56 without add<strong>in</strong>gany new current or gat<strong>in</strong>g variable.


396 Burst<strong>in</strong>g0-10membrane potential, V (mV)-20-30-40-50-60-700 10 20 30 40 50 60 70time (ms)Figure 9.56: Burst<strong>in</strong>g <strong>in</strong> a twodimensionalI Na,p +I K -model; see Ex. 3.0.21V-nullcl<strong>in</strong>ew0.1w-nullcl<strong>in</strong>e0.50V(t)0-0.5 0 0.5 1VI(t)-0.50 1000 2000 3000 4000timeFigure 9.57: Rebound burst<strong>in</strong>g <strong>in</strong> the FitzHugh-Nagumo oscillator; see Ex. 4.


Burst<strong>in</strong>g 3970.5x 1Figure 9.58: Hopf/Hopf burst<strong>in</strong>gtwithout co-existence of attractors;see Ex. 6 (modified from Hoppensteadtand Izhikevich 1997).4. (Rebound burst<strong>in</strong>g) Expla<strong>in</strong> the mechanism of rebound burst<strong>in</strong>g <strong>in</strong> the twodimensionalFitzHugh-Nagumo oscillator (4.11,4.12) shown <strong>in</strong> Fig. 9.57.5. Can “circle/*” and “fold/*” bursters have fast subthreshold oscillations of membranepotential? Expla<strong>in</strong>.6. (Hopf/Hopf burst<strong>in</strong>g) The systemẋ = (y + i)x − x|x| 2 , x = x 1 + ix 2 ∈ C ,has a unique attractor for any value of the parameter y ∈ R. Ifẏ = µ(2aS( y a − a) − |x|) , µ = 0.05 , a = √ µ/20 , S(u) =11 + e −uthen the “2+1” system above can burst, as we show <strong>in</strong> Fig. 9.58. Explore thesystem numerically and expla<strong>in</strong> the orig<strong>in</strong> of burst<strong>in</strong>g.7. (Hopf/Hopf canonical model) Consider the “2+1” fast-slow burster (9.1) andsuppose that x 0 is the supercritical Andronov-Hopf bifurcation po<strong>in</strong>t of the fastsubsystem when u = u 0 . Also suppose that u 0 is a stable equilibrium of theslow subsystem when x = x 0 is fixed. Show that there is a cont<strong>in</strong>uous change ofvariables that transforms (9.1) <strong>in</strong>to the canonical modelz ′ = (u + iω)z − z|z| 2 ,u ′ = µ(±1 ± u − a|z| 2 ) ,where z ∈ C is the new fast variable, u ∈ R is a slow variable, and ω, a and µ areparameters.8. (Burst<strong>in</strong>g <strong>in</strong> the I Na,t +I Na,slow -model) Take advantage of the phenomenon of<strong>in</strong>hibition-<strong>in</strong>duced spik<strong>in</strong>g described <strong>in</strong> Sect. 7.2.8 to show that a slow persistent<strong>in</strong>ward current, say I Na,slow , can make a spik<strong>in</strong>g model burst.9. Modify the example above to obta<strong>in</strong> repetitive burst<strong>in</strong>g <strong>in</strong> a model consist<strong>in</strong>g offast I Na,t current, leak current, and a slow passive dendritic compartment.


398 Burst<strong>in</strong>g0membrane potential, V (mV)-10-20-30-40-50-60-700 50 100 150 200 250 300 350time (ms)Figure 9.59: Burst<strong>in</strong>g <strong>in</strong> theI Na,p +I K +I Na,slow -model; see Ex. 10.y21.510.5y-nullcl<strong>in</strong>e10x-1-20-0.5-1-3 -2 -1 0 1 2 3xx-nullcl<strong>in</strong>e1I(t)0-10 200 400 600timeFigure 9.60: The phase portrait of the system <strong>in</strong> Ex. 11 shows that there is only onestable equilibrium for any value of I. Yet, the system bursts when I is periodicallymodulated.10. (Burst<strong>in</strong>g <strong>in</strong> the I Na,p +I K +I Na,slow -model) Explore numerically this model withthe fast subsystem as <strong>in</strong> Fig. 6.16 and a slow Na + current with parameters:g Na,slow = 0.5, m ∞,slow (V ) with V 1/2 = −50 mV and k = 10 mV, and τ slow (V ) =5 + 100 exp(−(V + 20) 2 /25 2 ). Expla<strong>in</strong> the orig<strong>in</strong> of burst<strong>in</strong>g oscillations whenI = 27 <strong>in</strong> Fig. 9.59.11. The Bonhoeffer–van der Pol oscillatorẋ = I + x − x 3 /3 − y ,ẏ = 0.2(1.05 + x) ,with nullcl<strong>in</strong>es as <strong>in</strong> Fig. 9.60, is Class 3 excitable. It has a unique stable equilibriumfor any value of I (check). Periodic modulations of I shift the x-nullcl<strong>in</strong>eup and down but do not change the stability of the equilibrium. Why does thesystem burst <strong>in</strong> Fig. 9.60? Explore the phenomenon numerically and expla<strong>in</strong> theexistence of repetitive spikes without a limit cycle.12. Prove that the fast-slow “2+2” systemż = (1 + u + iω)z − z|z| 2 , z ∈ C ,


Burst<strong>in</strong>g 399˙u = µ(u − u 3 − w) ,ẇ = µ(|z| 2 − 1) ,is a slow-wave burster, even though the slow subsystem cannot oscillate for anyfixed value of the fast subsystem z.13. (Ermentrout and Kopell 1986) Consider the system˙ϑ = 1 − cos ϑ + (1 + cos ϑ)r(ψ) ,˙ψ = ω ,with ϑ and ψ be<strong>in</strong>g phase variables on the unit circle S 1 and r(ψ) be<strong>in</strong>g anycont<strong>in</strong>uous function that changes sign. Show that this system exhibits burst<strong>in</strong>gactivity when ω is sufficiently small but positive. What type of burst<strong>in</strong>g is that?14. Prove that the canonical model for “subHopf/fold cycle” burst<strong>in</strong>g (9.11) exhibitssusta<strong>in</strong>ed burst<strong>in</strong>g activity when 0 < a < 1. What happens when a approaches0 or 1?15. Show that the canonical model for “fold/homocl<strong>in</strong>ic” burst<strong>in</strong>g (9.7) is equivalentto a simpler model (eq.27 <strong>in</strong> Izhikevich 2000 and Chap. 8)˙v = v 2 + w ,ẇ = µ ,with after-spike (v = +∞) resett<strong>in</strong>g v ← 1 and w ← w − d, when I is sufficientlylarge and µ and d are sufficiently small.16. Derive the canonical model for “fold/homocl<strong>in</strong>ic” burst<strong>in</strong>g (9.7) assum<strong>in</strong>g thatthe fast subsystem is near a saddle-node homocl<strong>in</strong>ic orbit bifurcation po<strong>in</strong>t atsome u = u 0 , which is an equilibrium of the slow subsystem.17. Derive the canonical models (9.9) and (9.10) for “circle/circle” burst<strong>in</strong>g.18. Show that the averaged slow subsystems of the canonical models for “circle/circle”bursters (9.9) and (9.10) have the formandrespectively, where˙u 1 = −µ 1 u 1 + d 1 f(I + u 1 − u 2 ) ,˙u 2 = −µ 2 u 2 + d 2 f(I + u 1 − u 2 ) ,˙u 1 = −µ 1 u 2 + d 1 f(I + u 1 ) ,˙u 2 = −µ 2 (u 2 − u 1 ) + d 2 f(I + u 1 ) ,f(u) =√ uπ/2 + arcot √ uis the frequency of spik<strong>in</strong>g of the fast subsystem (<strong>in</strong> Hz).


400 Burst<strong>in</strong>gFigure 9.61: A cycle-cycle burst<strong>in</strong>g:The rest<strong>in</strong>g state is not an equilibrium,but a small-amplitude limit cycle attractor."Fold/Fold" Burst<strong>in</strong>g"Fold/SubHopf" Burst<strong>in</strong>gUp-State(Rest)FoldBifurcationUp-State(Rest)FoldBifurcationFoldBifurcationSaddle Homocl<strong>in</strong>icOrbit BifurcationSubcriticalAndronov-HopfBifurcationxuDown-State(Rest)xuDown-State(Rest)FoldSubcriticalAndronov-HopfBifurcationFoldFoldSaddle Homocl<strong>in</strong>icOrbit BifurcationFigure 9.62: Two examples of po<strong>in</strong>t-po<strong>in</strong>t (not fast-slow) bursters (modified fromIzhikevich 2000).19. (Po<strong>in</strong>t-po<strong>in</strong>t hysteresis loops) Consider (9.1) and suppose that the fast subsystemhas only equilibria for any value of the one-dimensional slow variable u. If there isa co-existence of equilibrium po<strong>in</strong>ts of the fast subsystem, then (9.1) can exhibitpo<strong>in</strong>t-po<strong>in</strong>t hysteresis loop oscillation. Classify all co-dimension-1 po<strong>in</strong>t-po<strong>in</strong>thysteresis loops.20. (Po<strong>in</strong>t-po<strong>in</strong>t burst<strong>in</strong>g) In Fig. 9.62 we present two geometrical examples of po<strong>in</strong>tpo<strong>in</strong>tbursters that have no limit cycle attractors, yet are capable of exhibit<strong>in</strong>gspike-like dynamics <strong>in</strong> the active phase. Construct a model for each type of po<strong>in</strong>tpo<strong>in</strong>tburster <strong>in</strong> the figure. Use phase portrait snapshots at the bottom of thefigure as h<strong>in</strong>ts. What makes such burst<strong>in</strong>g possible?21. (Cycle-cycle bursters) Consider a fast-slow burster (9.1) and suppose that therest<strong>in</strong>g state is not an equilibrium, but a limit cycle attractor, as <strong>in</strong> Fig. 9.61.


Burst<strong>in</strong>g 401bifurcations of limit cyclessaddle-nodeon <strong>in</strong>variantcirclesaddlehomocl<strong>in</strong>icorbitsupercriticalAndronov-Hopffoldlimitcyclebifurcations of equilibriasaddle-node(fold)saddle-nodeon <strong>in</strong>variantcirclesupercriticalAndronov-HopfsubcriticalAndronov-Hopfsaddle-nodehomocl<strong>in</strong>ic orbitv' = I+v2+u 1u 1 ' = -µ 1 u 2u 2 ' = -µ 2 (u 2 -u 1 )v' = I+v2-uu' = -µuBaut<strong>in</strong>z'=(u+iω)z+2z|z| 2 -z|z| 4u'=µ(a-|z| 2 )Figure 9.63: Some canonical models of fast-slow bursters; see Ex. 26.Such a burst<strong>in</strong>g is called cycle-cycle. Classify all co-dimension-1 planar cyclecyclefast-slow bursters. Is burst<strong>in</strong>g <strong>in</strong> Fig. 9.10 of cycle-cycle type?22. (M<strong>in</strong>imal models for burst<strong>in</strong>g) Fill <strong>in</strong> the blank squares <strong>in</strong> Fig. 9.8.23. Choose a m<strong>in</strong>imal model from Fig. 9.8 and simulate it. Change the parametersto get as many different burst<strong>in</strong>g types as possible.24. [M.S.] Determ<strong>in</strong>e the bifurcation diagram of the canonical model for “fold/homocl<strong>in</strong>ic”burst<strong>in</strong>g (9.7).25. [M.S.] Determ<strong>in</strong>e the bifurcation diagrams of the canonical models for “circle/circle”bursters (9.9) and (9.10).26. [Ph.D.] Consider fast-slow bursters of the form (9.1) and assume that the fastsubsystem is near a bifurcation of high co-dimension, as <strong>in</strong> Fig. 9.28 or <strong>in</strong> Fig. 9.42.Treat<strong>in</strong>g the bifurcation po<strong>in</strong>t as an organiz<strong>in</strong>g center for the fast subsystem(Bertram et al. 1995, Izhikevich 2000, Golubitsky et al. 2001), use unfold<strong>in</strong>gtheory to derive canonical models for the rema<strong>in</strong><strong>in</strong>g fast-slow bursters <strong>in</strong> Fig. 9.63.Do not assume that the slow subsystem has an autonomous oscillation or thatthe fast oscillations have small amplitude.27. [Ph.D.] Classify all possible mechanisms of emergence of burst<strong>in</strong>g oscillationsfrom rest<strong>in</strong>g or spik<strong>in</strong>g, as <strong>in</strong> Fig. 9.19.


402 Burst<strong>in</strong>g28. [Ph.D.] Develop an asymptotic theory of s<strong>in</strong>gularly perturbed systems of theformẋ = f(x, u) (fast subsystem),˙u = µg(x, u) (slow modulation),that can deal with transitions between equilibria and limit cycle attractors of thefast subsystem.


Chapter 10Synchronization (seewww.izhikevich.com)This chapter, found on the author’s webpage www.izhikevich.com, considers networksof tonically spik<strong>in</strong>g neurons. As any other k<strong>in</strong>d of physical, chemical, or biologicaloscillators, such neurons could synchronize and exhibit collective behavior that isnot <strong>in</strong>tr<strong>in</strong>sic to any <strong>in</strong>dividual neuron. For example, partial synchrony <strong>in</strong> cortical networksis believed to generate various bra<strong>in</strong> oscillations, such as the alpha and gammaEEG rhythms. Increased synchrony may result <strong>in</strong> pathological types of activity, suchas epilepsy. Coord<strong>in</strong>ated synchrony is needed for locomotion and swim pattern generation<strong>in</strong> fish. There is an ongo<strong>in</strong>g debate on the role of synchrony <strong>in</strong> neural computation,see e.g., the special issue of Neuron (September 1999) devoted to the b<strong>in</strong>d<strong>in</strong>g problem.Depend<strong>in</strong>g on the circumstances, synchrony could be good or bad, and it is importantto know what factors contribute to synchrony and how to control it. This is thesubject of the present chapter – the most advanced chapter of the book. It provides anice application of the theory developed earlier and hopefully gives some <strong>in</strong>sight <strong>in</strong>towhy the previous chapters might be worth master<strong>in</strong>g. Unfortunately, it is too long tobe <strong>in</strong>cluded <strong>in</strong>to the book, so reviewers recommended to put it on the web.Our goal is to understand how the behavior of two coupled neurons depends on their<strong>in</strong>tr<strong>in</strong>sic dynamics. First, we <strong>in</strong>troduce the method of description of an oscillation byits phase. Then, we describe various methods of reduction of coupled oscillators tophase models. The reduction method and the exact form of the phase model dependson the type of coupl<strong>in</strong>g, i.e., whether it is pulsed, weak, or slow, and on the type of<strong>in</strong>-phase anti-phase out-of-phaseFigure 10.1: Different types of synchronization.403


404 Synchronization (see www.izhikevich.com)bifurcations of the limit cycle attractor generat<strong>in</strong>g tonic spik<strong>in</strong>g. F<strong>in</strong>ally, we show howto use phase models to understand the collective dynamics of many coupled oscillators.


Synchronization (see www.izhikevich.com) 405Review of Important Concepts• Oscillations are described by their phase variables ϑ rotat<strong>in</strong>g on a circleS 1 . We def<strong>in</strong>e ϑ as the time s<strong>in</strong>ce the last spike.• The phase response curve, PRC (ϑ), describes the magnitude of the phaseshift of an oscillator caused by a strong pulsed <strong>in</strong>put arriv<strong>in</strong>g at phase ϑ.• PRC depends on the bifurcations of spik<strong>in</strong>g limit cycle, and it def<strong>in</strong>essynchronization properties of an oscillator.• Two oscillators are synchronized <strong>in</strong>-phase, anti-phase, or out-of-phase,when their phase difference, ϑ 2 − ϑ 1 , equals 0, half-period, or some othervalue, respectively; see Fig. 10.1.• Synchronized states of pulse-coupled oscillators are fixed po<strong>in</strong>ts of thecorrespond<strong>in</strong>g Po<strong>in</strong>care phase map.• Weakly coupled oscillatorscan be reduced to phase modelsẋ i = f(x i ) + ε ∑ g ij (x j )˙ϑ i = 1 + ε Q(ϑ i ) ∑ g ij (x j (ϑ j )) ,where Q(ϑ) is the <strong>in</strong>f<strong>in</strong>itesimal PRC def<strong>in</strong>ed by Malk<strong>in</strong>’s equation.• Weak coupl<strong>in</strong>g <strong>in</strong>duces a slow phase deviation of the natural oscillation,ϑ i (t) = t + ϕ i , described by the averaged model(˙ϕ i = ε ω i + ∑ )H ij (ϕ j − ϕ i ) ,where the ω i denote the frequency deviations, andH ij (ϕ j − ϕ i ) = 1 T∫ Tdescribe the <strong>in</strong>teractions between the phases.0Q(t) g ij (x j (t + ϕ j − ϕ i )) dt• Synchronization of two coupled oscillators correspond to equilibria of theone-dimensional system˙χ = ε(ω + G(χ)) , χ = ϕ 2 − ϕ 1 ,where G(χ) = H 21 (−χ) − H 12 (χ) describes how the phase difference χcompensates for the frequency mismatch ω = ω 2 − ω 1 .


406 Synchronization (see www.izhikevich.com)


Solutions to ExercisesSolutions to Chapter 21. T = 20 ◦ C ≈ 293 ◦ F.E Ion = RTzF ln [Ion] out 8315 · 293 · ln 10=[Ion] <strong>in</strong> z · 96480when z = ±1. Therefore,log 10[Ion] out[Ion] <strong>in</strong>E K = 58 log(20/430) = −77 mVE Na = 58 log(440/50) = 55 mVE Cl = −58 log(560/65) = −54 mV= ± 58 log 10[Ion] out[Ion] <strong>in</strong>2.I = ḡ Na p (V − E Na ) + ḡ K p (V − E K ) = p{(ḡ Na + ḡ K ) V − ḡ Na E Na − ḡ K E K }()= (ḡ Na + ḡ K ) p V − ḡNa E Na + ḡ K E Kḡ Na + ḡ K} {{ } } {{ }ḡE3. The answer follows from the equation4. See Fig. 10.1.I − g L (V − E L ) = −g L (V − ÊL) , where Ê L = E L + I/g L .Function V 1/2 k Function V max σ C amp C basen ∞ (V ) 12 15 τ n (V ) −14 50 4.7 1.1m ∞ (V ) 25 9 τ m (V ) 27 30 0.46 0.04h ∞ (V ) 3 −7 τ n (V ) −2 20 7.4 1.2Remark: Hodgk<strong>in</strong> and Huxley shifted V 1/2 and V max by 65 mV so that the rest<strong>in</strong>g potential isat V = 0 mV.5. (Willms et al. 1999)Ṽ 1/2 = V 1/2 − k ln(2 1/p − 1) ,˜k =k2p(1 − 2 −1/p ) .The first equation is obta<strong>in</strong>ed from the condition m p ∞(Ṽ1/2) = 1/2. The second equation isobta<strong>in</strong>ed from the condition that the two functions have the same slope at V = Ṽ1/2.6. See author’s webpage.7. See author’s webpage.407


408 Solutions to Exercises, Chap. 310.90.8h (V)987τ (V) h0.760.60.50.40.30.20.1n (V)m (V)0-40 -20 0 20 40 60 80V (mV)54321τ (V)m0-40 -20 0 20 40 60 80 100V (mV)τ (V) nFigure 10.1: Open dots: The steady-state (<strong>in</strong>)activation functions and voltage-sensitive time constants<strong>in</strong> the Hodgk<strong>in</strong>-Huxley model. Filled dots: steady-state Na + activation function m ∞ (V ) <strong>in</strong>the squid giant axon (experimental results by Hodgk<strong>in</strong> and Huxley, 1952, Fig. 8). Cont<strong>in</strong>uous curves:Approximations by Boltzmann and Gaussian functions. See Ex. 4.Solutions to Chapter 31. Consider the limit case: (1) activation of Na + current is <strong>in</strong>stantaneous, and (2) conductancek<strong>in</strong>etics of the other currents are frozen. Then, the Na + current will result <strong>in</strong> the nonl<strong>in</strong>earterm g Na m ∞ (V ) (V − E Na ) with the parameter h ∞ (V rest ) <strong>in</strong>corporated <strong>in</strong>to g Na , and all theother currents will result <strong>in</strong> the l<strong>in</strong>ear leak term.In Fig. 3.15, the activation of the Na + current is not <strong>in</strong>stantaneous, hence the sag right afterthe pulses. In addition, its <strong>in</strong>activation, as well as the k<strong>in</strong>etics of the other currents are notslow enough, hence the membrane potential quickly reaches the excited state and then slowlyrepolarizes back to the rest<strong>in</strong>g state.2. See Fig. 10.2. The eigenvalues are negative at each equilibrium marked as filled circle (stable),and positive at each equilibrium marked as open circle (unstable). The eigenvalue at thebifurcation po<strong>in</strong>t (left equilibrium <strong>in</strong> Fig. 10.2b) is zero.F(V) F(V) F(V)V V Va b cFigure 10.2: Phase portraits of the system ˙V = F (V ) with given F (V ).3. Phase portraits are shown <strong>in</strong> Fig. 10.3.(a) The equation 0 = −1 + x 2 has two solutions: x = −1 and x = +1, hence there aretwo equilibria <strong>in</strong> the system (a). The eigenvalues are the derivatives at each equilibrium,λ = (−1 + x 2 ) ′ = 2x, where x = ±1. Equilibrium x = −1 is stable because λ = −2 < 0.Equilibrium x = +1 is unstable because λ = +2 > 0. The same fact follows from thegeometrical analysis <strong>in</strong> Fig. 10.3.


Solutions to Exercises, Chap. 3 409(b) The equation 0 = x − x 3 has three solutions: x = ±1 and x = 0, hence there are threeequilibria <strong>in</strong> the system (b). The eigenvalues are the derivatives at each equilibrium, λ =(x − x 3 ) ′ = 1 − 3x 2 . The equilibria x = ±1 are stable because λ = 1 − 3(±1) 2 = −2 < 0.The equilibrium x = 0 is unstable because λ = 1 > 0. The same conclusions also followfrom the geometrical analysis <strong>in</strong> Fig. 10.3.1.51x2 x -110.5xx-x 30.50x0x-0.5-0.5-1-2 -1 0 1 2a-1-2 -1 0 1 2bFigure 10.3: Phase portraits of the systems (a) ẋ = −1 + x 2 , (b) ẋ = x − x 3 .4. The equilibrium x = 0 is stable <strong>in</strong> all three cases.5. See Fig. 10.4. Topologically equivalent systems are <strong>in</strong> (a), (b), and (c). In (d) there aredifferent numbers of equilibria; no stretch<strong>in</strong>g or shr<strong>in</strong>k<strong>in</strong>g of the rubber phase l<strong>in</strong>e can producenew equilibria. In (e) the right equilibrium is unstable <strong>in</strong> ˙V = F 1 (V ), but stable <strong>in</strong> ˙V = F 2 (V );no stretch<strong>in</strong>g or shr<strong>in</strong>k<strong>in</strong>g can change the stability of an equilibrium. In (f) the flow betweenthe two equilibria is directed rightward <strong>in</strong> ˙V = F 1 (V ) and leftward <strong>in</strong> ˙V = F 2 (V ); no stretch<strong>in</strong>gor shr<strong>in</strong>k<strong>in</strong>g can change the direction of the flow.6. (Saddle-node (fold) bifurcation <strong>in</strong> ẋ = a + x 2 ) The equation 0 = a + x 2 has no real solutionswhen a > 0, and two solutions x = ± √ |a| when a ≤ 0. Hence there are two branches ofequilibria, depicted <strong>in</strong> Fig. 10.5. The eigenvalues are7.λ = (a + x 2 ) ′ = 2x = ±2 √ |a| .The lower branch − √ |a| is stable (λ < 0), and the upper branch + √ |a| is unstable (λ > 0).They meet at the saddle-node (fold) bifurcation po<strong>in</strong>t a = 0.(a) x = −1 at a = 1, (b) x = −1/2 at a = 1/4 (c) x = 1/2 at a = 1/4(d) x = ±1/ √ 3 at a = ±2/(3 √ 3) (e) x = ±1 at a = ∓2 (f) x = −1 at a = 18. (Pitchfork bifurcation <strong>in</strong> ẋ = bx − x 3 ) The equation 0 = bx − x 3 has one solution x = 0 whenb ≤ 0, and three solutions x = 0, x = ± √ b when b > 0. Hence there is only one branch ofequilibria for b < 0 and three branches for b > 0 of the pitchfork curve depicted <strong>in</strong> Fig. 10.6.The eigenvalues areλ = (bx − x 3 ) ′ = b − 3x 2 .The branch x = 0 exists for any b and its eigenvalue is λ = b. Thus, it is stable for b < 0 andunstable for b > 0. The two branches x = ± √ b exist only for b > 0, but they are always stablebecause λ = b−3(± √ b) 2 = −2b < 0. We see that the branch x = 0 loses stability when b passesthe pitchfork bifurcation value b = 0, at which po<strong>in</strong>t a pair of new stable branches bifurcates(hence the name bifurcation). In other words, the stable branch x = 0 divides (bifurcates) <strong>in</strong>totwo stable branches when b passes 0.


410 Solutions to Exercises, Chap. 3TopologicallyEquivalentVTopologicallyEquivalentF1(V)VTopologicallyEquivalentVF (V) 2F (V)1VF (V) 2VF (V) 2F (V)1VabcTopologically NOTEquivalentF1(V)F (V) 2?VVTopologically NOTEquivalentF1(V)LeftF (V) 2Topologically NOTF1(V)Equivalent? ??RightUnstableStableRightLeftVVdeFigure 10.4: Answer to Chap. 3, Ex. 5.f9. Recall that the current I Kir is turned off by depolarization and turned on by hyperpolarization.The dynamics of the I Kir -model is similar to that of the I Na,p -model <strong>in</strong> many respects. In particular,this system can also have co-existence of two stable equilibria separated by an unstableequilibrium, which follows from the N-shaped I-V relation. Indeed, when V is hyperpolarized,the current I Kir is turned on (de<strong>in</strong>activated) and it pulls V toward E K . In contrast, when V isdepolarized, the current is turned off (<strong>in</strong>activated) and does not obstruct further depolarizationof V .We use (3.11) to f<strong>in</strong>d the curveI = g L (V − E L ) + g Kir h ∞ (V )(V − E K ) ,<strong>in</strong> Fig. 10.8. (The curve might not be S-shaped if a different bifurcation parameter is used, as<strong>in</strong> Ex. 12a).The bifurcation diagram of the I Kir -model (3.11) <strong>in</strong> Fig. 10.8 has three branches correspond<strong>in</strong>gto the three equilibria. When the parameter I is relatively small, the outward I Kir currentdom<strong>in</strong>ates and the system has only one equilibrium <strong>in</strong> the low voltage range — the “downstate”.When the parameter I is relatively large, the <strong>in</strong>jected <strong>in</strong>ward current I dom<strong>in</strong>ates, andthe system has one equilibrium <strong>in</strong> the <strong>in</strong>termediate voltage range — the “up-state”. Whenthe parameter I is <strong>in</strong> neighborhood of I = 6, the system exhibits bistability of the “up-” and“down-states”. The states appear and disappear via saddle-node bifurcations. We see thatthe behavior of the I Kir -model is conceptually (and qualitatively) similar to the behavior ofthe I Na,p -model (3.5) even though the models have completely different ionic mechanisms forbistability.10. The equilibrium satisfies the one-dimensional equation0 = I − g K n 4 ∞(V )(V − E K ) − g Na m 3 ∞(V )h ∞ (V )(V − E Na ) − g L (V − E L ) ,


Solutions to Exercises, Chap. 3 411Unstableequilibria10-1StableequilibriuaxBifurcation DiagramaaSaddle-node(fold)bifurcation-2 -1 0 0.5 1aRepresentative Phase Portraits1.51xa+x 21.51x1.5a+x2xa+x210.50x0.50x0.50x-0.5-0.5-0.5-1-2 -1 0 1 2a=-1-1-2 -1 0 1 2a=0-1-2 -1 0 1 2a=0.5Figure 10.5: Saddle-node (fold) bifurcation diagram and representative phase portraits of the systemẋ = a + x 2 (see Chap. 3, Ex. 6).where all gat<strong>in</strong>g variables assume their asymptotic values. The solutionI = g K n 4 ∞(V )(V − E K ) + g Na m 3 ∞(V )h ∞ (V )(V − E Na ) + g L (V − E L )is depicted <strong>in</strong> Fig. 10.9. S<strong>in</strong>ce the curve <strong>in</strong> Fig. 10.9 does not have folds, there are no saddle-nodebifurcations <strong>in</strong> the Hodgk<strong>in</strong>-Huxley model (with the orig<strong>in</strong>al values of parameters).11. The curvesand(b)are depicted <strong>in</strong> Fig. 10.7.12. The curves13.and(a) g L (V ) = −g Na m ∞ (V )(V − E Na )/(V − E L )E L (V ) = V + g Na m ∞ (V )(V − E Na )(V − E L )/g L(a) g L (V ) = {I − g Kir h ∞ (V )(V − E K )}/(V − E L )(b) g Kir (V ) = {I − g L (V − E L )}/{h ∞ (V )(V − E K )}are depicted <strong>in</strong> Fig. 10.10. Notice that the curve <strong>in</strong> Fig. 10.10a does not have the S shape.F ′ (V ) = −g L − g K m 4 ∞(V ) − g K 4m 3 ∞(V )m ′ ∞(V )(V − E K ) < 014.because g L > 0, m ∞ (V ) > 0, m ′ ∞(V ) > 0, and V − E K > 0 for all V > E K .F ′ (V ) = −g L − g h h ∞ (V ) − g h h ′ ∞(V )(V − E h ) < 0because g L > 0, h ∞ (V ) > 0, but h ′ ∞(V ) < 0 and V − E h < 0 for all V < E h .


412 Solutions to Exercises, Chap. 321.51xBifurcationDiagramstableb0.50stableunstableb-0.5-1Pitchforkbifurcationstable-1.5b-2-2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5 3Representative Phase Portraits42xbx-x 321xbx-x 310.5xbx-x 30x0x0x-2-1-0.5-4-2 0 2b=-1-2-2 0 2-1-2 0 2b=0 b=+1Figure 10.6: Pitchfork bifurcation diagram and representative phase portraits of the system ẋ =bx − x 3 (see Chap. 3, Ex. 8).Voltage V40200-20-40-60excitedthresholdrest-800.5 1 1.5 2 2.5leak conductance g Lasaddle-node(fold)bifurcationsVoltage V500-50-100thresholdrestexcited-150-150 -100 -50 0leak reverse potential E Lbsaddle-node(fold)bifurcationsFigure 10.7: Bifurcation diagrams of the I Na,p -model (3.5) with bifurcation parameters (a) g L and(b) E L (see Chap. 3, Ex. 11).


Solutions to Exercises, Chap. 3 413-10MonostabilityBistabilityMonostabilityMembrane Voltage, V (mV)-20-30-40-50-60Saddle-node(fold) bifurcationRest ("down") stateThreshold-704 4.5 5 5.5 6 6.5 7 7.5 8Injected Curent, IRest ("up") stateSaddle-node(fold) bifurcationFigure 10.8: Bifurcation diagram of the I Kir -model (3.11).12010020membrane voltage, V806040200membrane voltage, V100-10magnification-200 1000 2000 3000 4000 5000<strong>in</strong>jected dc-current, I-200 50 100 150 200<strong>in</strong>jected dc-current, IFigure 10.9: Dependence of the position of equilibrium <strong>in</strong> the Hodgk<strong>in</strong>-Huxley model on the <strong>in</strong>jecteddc-current; see Ex. 10.15. When V is sufficiently large, ˙V ≈ V 2 . The solution of ˙V = V 2 is V (t) = 1/(c − t) (check bydifferentiat<strong>in</strong>g), where c = 1/V (0). Another way to show this is to solve (3.9) for V and f<strong>in</strong>dthe asymptote of the solution.16. Each equilibrium of the system ẋ = a + bx − x 3 is a solution to the equation 0 = a + bx − x 3 .Treat<strong>in</strong>g x and b as free parameters, the set of all equilibria is given by a = −bx + x 3 , and itlooks like the cusp surface <strong>in</strong> Fig. 6.34. Each po<strong>in</strong>t where the cusp surface folds corresponds toa saddle-node (fold) bifurcation. The derivative with respect to x at each such po<strong>in</strong>t is zero;alternatively, the tangent vector to the cusp surface at each such po<strong>in</strong>t is parallel to the x-axis.The set of all bifurcation po<strong>in</strong>ts is projected to the (a, b)-plane at the bottom of the figure, andit looks like a curve hav<strong>in</strong>g two branches. To f<strong>in</strong>d the equation for the bifurcation curves oneneeds to remember that each bifurcation po<strong>in</strong>t satisfies two conditions:• It is an equilibrium; that is, a + bx − x 3 = 0.• The derivative of a + bx − x 3 with respect to x is zero; that is, b − 3x 2 = 0.Solv<strong>in</strong>g the second equation for x and us<strong>in</strong>g the solution x = ± √ b/3 <strong>in</strong> the first equation yieldsa = ∓2(b/3) 3/2 . The po<strong>in</strong>t a = b = 0 is called a cusp bifurcation po<strong>in</strong>t.


414 Solutions to Exercises, Chap. 4Voltage V40200-20-40-60thresholdup-statedown-statesaddle-node(fold)bifurcation-800 0.1 0.2 0.3 0.4E LVoltage V-30-40-50-60saddle-node(fold)bifurcationup-statethresholddown-statesaddle-node(fold)bifurcation-701.7 1.8 1.9 2 2.1 2.2leak conductance g LaKir conductance g KirbFigure 10.10: Bifurcation diagrams of the I Kir -model (3.11), I = 6, with bifurcation parameters (a)g L and (b) g Kir (see Chap. 3, Ex. 12).17. (Gradient systems) For ˙V = F (V ) takewhere c is any constant.∫ VE(V ) = − F (v) dv ,ca. E(V ) = 1 b. E(V ) = −V c. E(V ) = V 2 /2d. E(V ) = V − V 3 /3 e. E(V ) = −V 2 /2 + V 4 /4 f. E(V ) = − cos V18. (c) implies (b) because |x(t) − y| < exp(−at) implies that x(t) → y as t → ∞. (b) implies (a)accord<strong>in</strong>g to the def<strong>in</strong>ition.(a) does not imply (b) because x(t) might not approach y. For example, y = 0 is an equilibrium<strong>in</strong> the system ẋ = 0 (any other po<strong>in</strong>t is also an equilibrium). It is stable, s<strong>in</strong>ce |x(t)−0| < ε for all|x 0 −0| < ε and all t ≥ 0. However, it is not asymptotically stable because lim t→∞ x(t) = x 0 ≠ 0regardless of how close x 0 is to 0 (unless x 0 = 0).(b) does not imply (c). For example, the equilibrium y = 0 <strong>in</strong> the system ẋ = −x 3 is asymptoticallystable (check by differentiat<strong>in</strong>g that x(t) = (2t + x −20 )−1/2 → 0 is a solution withx(0) = x 0 ), however x(t) approaches 0 with a slower than exponential rate, exp(−at), for anyconstant a > 0.Solutions to Chapter 41. See figures 10.11 through 10.15.2. See Fig. 10.16.3. See figures 10.17 through 10.21.4. The diagram follows from the form of the eigenvaluesλ = τ ± √ τ 2 − 4∆2.


Solutions to Exercises, Chap. 4 415Figure 10.11: Nullcl<strong>in</strong>es of the vector field; see also Fig. 10.17.Figure 10.12: Nullcl<strong>in</strong>es of the vector field; see also Fig. 10.18.Figure 10.13: Nullcl<strong>in</strong>es of the vector field; see also Fig. 10.19.


416 Solutions to Exercises, Chap. 4Figure 10.14: Nullcl<strong>in</strong>es of the vector field; see also Fig. 10.20.Figure 10.15: Nullcl<strong>in</strong>es of the vector field; see also Fig. 10.21.


Solutions to Exercises, Chap. 4 417abcdFigure 10.16: Approximate directions of the vector <strong>in</strong> each region between the nullcl<strong>in</strong>es.If ∆ < 0 (left half-plane <strong>in</strong> Fig. 4.15), then the eigenvalues have opposite signs. Indeed,√τ2− 4∆ > √ τ 2 = |τ| ,whenceτ + √ τ 2 − 4∆ > 0 and τ − √ τ 2 − 4∆ < 0 .The equilibrium is a saddle <strong>in</strong> this case. Now consider the case ∆ > 0. When τ 2 < 4∆ (<strong>in</strong>sidethe parabola <strong>in</strong> Fig. 4.15), the eigenvalues are complex-conjugate, hence the equilibrium is afocus. It is stable (unstable) when τ < 0 (τ > 0). When τ 2 > 4∆ (outside the parabola <strong>in</strong>Fig. 4.15), the eigenvalues are real. They both are negative (positive) when τ < 0 (τ > 0).5. (van der Pol oscillator) The nullcl<strong>in</strong>es of the van der Pol oscillator,y = x − x 3 /3 (x-nullcl<strong>in</strong>e) ,x = 0 (y-nullcl<strong>in</strong>e) ,are depicted <strong>in</strong> Fig. 10.22. There is a unique equilibrium (0, 0). The Jacobian matrix at theequilibrium has the form( )1 −1L =.b 0S<strong>in</strong>ce tr L = 1 > 0 and det L = b > 0, the equilibrium is always an unstable focus.6. (Bonhoeffer–van der Pol oscillator) The nullcl<strong>in</strong>es of the Bonhoeffer–van der Pol oscillator withc = 0 have the formy = x − x 3 /3 (x-nullcl<strong>in</strong>e) ,x = a (y-nullcl<strong>in</strong>e) ,


418 Solutions to Exercises, Chap. 4Figure 10.17: Left: No equilibria. Right: Saddle equilibrium.Figure 10.18: Left: Stable node. Right: Stable focus.Figure 10.19: Left: Excitable system hav<strong>in</strong>g one stable equilibrium. Right: Two stable nodesseparated by a saddle equilibrium.


Solutions to Exercises, Chap. 4 419Figure 10.20: Left: Unstable focus <strong>in</strong>side a stable limit cycle. Right: Stable focus <strong>in</strong>side an unstablelimit cycle.Figure 10.21: Left: Saddle-node equilibrium. Right: Stable node and saddle equilibria connectedby two heterocl<strong>in</strong>ic trajectories, which form an <strong>in</strong>variant circle with an unstable focus <strong>in</strong>side.1y0-1-2 -1 0x1 2Figure 10.22: Nullcl<strong>in</strong>es and phase portrait of the van der Pol oscillator (b = 0.1).


420 Solutions to Exercises, Chap. 4a=-1.2a=-0.8y10.50-0.5-2 -1 0 1 2 -2 -1 0 1 2xxFigure 10.23: Nullcl<strong>in</strong>es and phase portrait of the Bonhoeffer–van der Pol oscillator (b = 0.05 andc = 0).-1shown <strong>in</strong> Fig. 10.23. They <strong>in</strong>tersect at the po<strong>in</strong>t x = a, y = a − a 3 /3. The Jacobian matrix atthe equilibrium (a, a − a 3 /3) has the form( )1 − a2−1L =.b 0S<strong>in</strong>ce tr L = 1 − a 2 and det L = b > 0, the equilibrium is a stable (unstable) focus when |a| > 1(|a| < 1), as we illustrate <strong>in</strong> Fig. 10.23.7. (H<strong>in</strong>dmarsh-Rose spik<strong>in</strong>g neuron) The Jacobian matrix at the equilibrium (¯x, ȳ) is( )f′−1L =g ′ ,−1thereforetr L = f ′ − 1 and det L = −f ′ + g ′ .The equilibrium is a saddle (det L < 0) when g ′ < f ′ , i.e., <strong>in</strong> the region below the diagonal <strong>in</strong>Fig. 10.24. When g ′ > f ′ , the equilibrium is stable (tr L < 0) when f ′ < 1, which is the lefthalf-plane <strong>in</strong> Fig. 10.24. Us<strong>in</strong>g the classification <strong>in</strong> Fig. 4.15, we conclude that it is focus when(f ′ − 1) 2 − 4(g ′ − f ′ ) < 0, i.e., wheng ′ > 1 4 (f ′ + 1) 2which is the upper part of the parabola <strong>in</strong> Fig. 10.24.8. (I K -model) The steady-state I-V relation of the I K -model is monotone, hence it has a uniqueequilibrium, which we denote here as ( ¯V , ¯m) ∈ R 2 , where ¯V > E K and ¯m = m ∞ ( ¯V ). TheJacobian at the equilibrium has the form(−(gL + ḡL =K ¯m 4 )/C −4ḡ K ¯m 3 ( ¯V)− E K )/Cm ′ ∞( ¯V )/τ( ¯V ) −1/τ( ¯V ,)with the signsL =(− −+ −Obviously, det L > 0 and tr L < 0, hence the equilibrium (focus or node) is always stable.).


Solutions to Exercises, Chap. 5 4216543g'stable focusf'=1unstable focusg'=(f'+1) 2unstable nodef'=g'210stable nodesaddle-1-2 -1 0 1 2 3 4f'Figure 10.24: Stability diagram ofthe H<strong>in</strong>dmarsh-Rose spik<strong>in</strong>g neuronmodel; see Ex. 7.9. (I h -model) The steady-state I-V relation of the I h -model is monotone, hence it has a uniqueequilibrium denote here as ( ¯V , ¯h) ∈ R 2 , where ¯V < E h and ¯h = h ∞ ( ¯V ). The Jacobian at theequilibrium has the form(−(gL + ḡL =h¯h)/C −ḡh ( ¯V)− E h )/Ch ′ ∞( ¯V )/τ( ¯V ) −1/τ( ¯V ,)with the signs( )− +L =.− −Obviously, det L > 0 and tr L < 0, hence the equilibrium is always stable.10. (Bendixson’s criterion) The divergence of the vector field of the I K -model∂f(x,y)/∂x{ }} {(−g L − ḡ K m 4 )/C +∂g(x,y)/∂y{ }} {−1/τ(V )is always negative, hence the model cannot have a periodic orbit. Therefore, it cannot havesusta<strong>in</strong>ed oscillations.11. The x-nullcl<strong>in</strong>e is y = a + x 2 and the y-nullcl<strong>in</strong>e is y = bx/c, as <strong>in</strong> Fig. 10.25. The equilibria(<strong>in</strong>tersections of the nullcl<strong>in</strong>es) are¯x = b/c ± √ (b/c) 2 − 4a2, ȳ = b¯x/c ,provided that a < 1 4 (b/c)2 . The Jacobian matrix at (¯x, ȳ) has the formwith tr L = 2¯x − c and( )2¯x −1L =b −cdet L = −2¯xc + b = ∓ √ b 2 − 4ac 2 .Thus, the right equilibrium (i.e., (b/c + √ (b/c) 2 − 4a)/2) is always a saddle and the left equilibrium(i.e., (b/c − √ (b/c) 2 − 4a)/2) is always a focus or a node. It is always stable when itlies on the left branch of the parabola y = a + x 2 (i.e., when ¯x < 0), and can also be stable onthe right branch if it is not too far from the parabola knee (i.e., if ¯x < c/2); see Fig. 10.25.


422 Solutions to Exercises, Chap. 5yy=a+x2y=bx/cunstable (saddle)stable if x


Solutions to Exercises, Chap. 6 423-30membrane voltage, V (mV)-40-50-60-70(-24, -41)(0.45, -63)-80-30 -20 -10 0 10<strong>in</strong>jected dc-current, II Na,t -modelmembrane voltage, V (mV)0I A -model-20-40(10.75, -19)(12.7, -42)-600 5 10 15<strong>in</strong>jected dc-current, IFigure 10.27: The saddle-node bifurcation diagrams of the I Na,t - and I A -m<strong>in</strong>imal models.3. The curvesandare depicted <strong>in</strong> Fig. 10.27.I = g L (V − E L ) + ḡ Na m 3 ∞(V )h ∞ (V )(V − E Na )I = g L (V − E L ) + ḡ A m ∞ (V ) h ∞ (V )(V − E K )4. g is not an absolute conductance, but is taken relative to the conductance at rest<strong>in</strong>g state.Negative values occur because the <strong>in</strong>itial hold<strong>in</strong>g voltage value <strong>in</strong> the voltage-clamp experimentdescribed <strong>in</strong> Fig. 5.22a corresponds to the rest<strong>in</strong>g potential, at which the K + conductance ispartially activated. Indeed, <strong>in</strong> the I Na,p +I K -model the K + gat<strong>in</strong>g variable n ≈ 0.04, hencethe K + conductance is approximately 0.4 (because ḡ K = 10). Accord<strong>in</strong>g to the procedure,this value corresponds to g = 0. Any small decrease <strong>in</strong> conductance would result <strong>in</strong> negativevalues of g. If the <strong>in</strong>itial hold<strong>in</strong>g voltage were very negative, say below −100 mV, then the slowconductance g would have non-negative values <strong>in</strong> the relevant voltage range (above −100 mV).5. The curve I slow (V ) def<strong>in</strong>es slow changes of the membrane voltage. The curve I −I fast (V ) def<strong>in</strong>esfast changes. Its middle part, which has positive slope, is unstable. If the I-V curves <strong>in</strong>tersect<strong>in</strong> the middle part, the equilibrium is unstable, and the system exhibits periodic spik<strong>in</strong>g: Thevoltage slides down slowly along the left branch of the fast I-V curve toward the slow I-V curveuntil it reaches the left knee, and then it jumps quickly to the right branch. After the jump,the voltage slides up slowly along the right branch until it reaches the right knee, and then itquickly jumps to the left branch along the straight l<strong>in</strong>e that connects the knee and the po<strong>in</strong>t(E K , 0) (see also previous exercise). Notice that the direction of the jump is not horizontal,as <strong>in</strong> relaxation oscillators, but along a sloped l<strong>in</strong>e. On that l<strong>in</strong>e the slow conductance g isconstant, but the slow current I slow (V ) = g(V − E K ) changes fast because the driv<strong>in</strong>g forceV − E K changes fast. When the I-V curves <strong>in</strong>tersect at the stable po<strong>in</strong>t (negative slope ofI − I fast (V )), the voltage variable may produce a s<strong>in</strong>gle action potential, and then slides slowlytoward the <strong>in</strong>tersection, which is a stable equilibrium.Solutions to Chapter 61. There are two equilibria: x = 0 and x = b. The stability is determ<strong>in</strong>ed by the sign of thederivativeλ = (x(b − x)) ′ x = b − 2xat the equilibrium. S<strong>in</strong>ce λ = b when x = 0, this equilibrium is stable (unstable) when b < 0(b > 0). S<strong>in</strong>ce λ = −b when x = b, this equilibrium is unstable (stable) when b < 0 (b > 0).


424 Solutions to Exercises, Chap. 62. (a) The systemẋ = bx 2 , b ≠ 0cannot exhibit saddle-node bifurcation: It has one equilibrium for any non-zero b, or an<strong>in</strong>f<strong>in</strong>ite number of equilibria when b = 0. The equilibrium x = 0 is non-hyperbolic and thenon-degeneracy condition is satisfied (a = b ≠ 0). However, the transversality conditionis not satisfied at the equilibrium x = 0. Another example is ẋ = b 2 + x 2 .(b) The systemẋ = b − x 3has a s<strong>in</strong>gle stable equilibrium for any b. However, the po<strong>in</strong>t x = 0 is non-hyperbolic whenb = 0, and the transversality condition is also satisfied. The non-degeneracy condition isviolated, though.3. It is easy to check (by differentiat<strong>in</strong>g) that√c(b − bsn )V (t) = √ tan( √ ac(b − b sn )t)ais a solution to the system. S<strong>in</strong>ce tan(−π/2) = −∞ and tan(+π/2) = +∞, it takesfor the solution to go from −∞ to +∞.T =π√ac(b − bsn )4. The first system can be transformed <strong>in</strong>to the second one if we use complex coord<strong>in</strong>ates z = u+iv.To obta<strong>in</strong> the third system, we use polar coord<strong>in</strong>atesso thatre iϕ = z = u + iv ∈ C ,ż(c(b)+iω(b))z+(a+id)z|z| 2{ }} { { }} {ṙe iϕ + re iϕ i ˙ϕ = (c(b) + iω(b))re iϕ + (a + id)r 3 e iϕ .Next, we divide both sides of this equation by e iϕ and separate the real and imag<strong>in</strong>ary partsto obta<strong>in</strong>{ṙ − c(b)r − ar 3 } + ir{ ˙ϕ − ω(b) − dr 2 } = 0 ,which we can write <strong>in</strong> the polar-coord<strong>in</strong>ates form.5. (a) The equilibrium r = 0 of the systemṙ = br 3 ,˙ϕ = 1 ,has a pair of complex-conjugate eigenvalues ±i for any b, and the non-degeneracy conditionis satisfied for any b ≠ 0. However, the transversality condition is violated, andthe system does not exhibit Andronov-Hopf bifurcation (no limit cycle exists near theequilibrium).(b) The equilibrium r = 0 for b = 0ṙ = br ,˙ϕ = 1 ,has a pair of complex-conjugate eigenvalues ±i and the transversality condition is satisfied.However, the bifurcation is not of the Andronov-Hopf type because no limit cycleexists near the equilibrium for any b.


Solutions to Exercises, Chap. 6 4256. The Jacobian matrix at the equilibrium (u, v) = (0, 0) has the form( )b −1L =.1 bIt has eigenvalues b±i. Therefore, the loss of stability occurs at b = 0, and the non-hyperbolicityand transversality conditions are satisfied. S<strong>in</strong>ce the model can be reduced to the polarcoord<strong>in</strong>atesystem (see Ex. 4), and a ≠ 0, the non-degeneracy condition is also satisfied, andthe system undergoes an Andronov-Hopf bifurcation.7. S<strong>in</strong>ce(cr + ar 3 ) ′ r = c + 3ar 2 = c + 3a|c/a| =the limit cycle is stable when a < 0{c + 3|c| when a > 0,c − 3|c| when a < 0,8. The sequence of bifurcations is similar to that of the RS neuron <strong>in</strong> Fig. 8.15. The rest<strong>in</strong>gstate is a globally asymptotically stable equilibrium for I < 5.64. At this value a stable(spik<strong>in</strong>g) limit cycle appears via a big saddle homocl<strong>in</strong>ic orbit bifurcation. At I = 5.8 asmall-amplitude unstable limit cycle is born via another saddle homocl<strong>in</strong>ic orbit bifurcation.This cycle shr<strong>in</strong>ks to the equilibrium and makes it lose stability via subcritical Andronov-Hopfbifurcation at I = 6.5. This unstable focus becomes an unstable node when I <strong>in</strong>creases, andthen it coalesces with the saddle (at I = 7.3) and disappears. Notice that there is a saddle-nodebifurcation accord<strong>in</strong>g to the I-V relation, but it corresponds to the disappearance of an unstableequilibrium.9. The Jacobian matrix of partial derivatives has the form( )−I′L = V (V, x) −I x(V, ′ x)x ′ ,∞(V )/τ(V ) −1/τ(V )so thatandThe characteristic equationhas two solutionstr L = −{I ′ V (V, x) + 1/τ(V )}det L = {I ′ V (V, x) + I ′ x(V, x)x ′ ∞(V )}/τ(V ) = I ′ ∞(V )/τ(V ) .which might be complex-conjugate.10. Let z = re ϕi , thenλ 2 − λ tr L + det L = 0c{ }} { { √ }} {(tr L)/2 ± {(tr L)/2}2 − det Lωr ′ = ar + r 3 − r 5 ,ϕ ′ = ω .Any limit cycle is an equilibrium of the amplitude equation, i.e.,a + r 2 − r 4 = 0 .The system undergoes fold limit cycle bifurcation when the amplitude equation undergoes asaddle-node bifurcation, i.e., whena + 3r 2 − 5r 4 = 0(check the non-degeneracy and transversality conditions). The two equations have the nontrivialsolution (a, r) = (−1/4, 1/ √ 2).


426 Solutions to Exercises, Chap. 611. The projection onto the v 1 -axis is described by the equationẋ = λ 1 x , x(0) = a .The trajectory leaves the square when x(t) = ae λ1t = 1; that is, whent = − 1 ln a = − 1 ln τ(I − I b ) .λ 1 λ 112. Equation (6.13) has two bifurcation parameters, b and v, and the saddle-node homocl<strong>in</strong>icbifurcation occurs when b = b sn and v = V sn . The saddle-node bifurcation curve is the straightl<strong>in</strong>e b = b sn (any v). This bifurcation is on an <strong>in</strong>variant circle when v < V sn and off otherwise.When b > b sn , there are no equilibria and the normal form exhibits periodic spik<strong>in</strong>g. Whenb < b sn , the normal form has two equilibria,node{ }} {V sn − √ c|b − b sn |/aandsaddle{ }} {V sn + √ c|b − b sn |/a .The saddle homocl<strong>in</strong>ic orbit bifurcation occurs when the voltage is reset to the saddle, i.e.,whenv = V sn + √ c|b − b sn |/a .13. The Jacobian matrix at an equilibrium is( 2v −1L =ab −a).Saddle-node condition det L = −2va + ab = 0, results <strong>in</strong> v = b/2. S<strong>in</strong>ce v is an equilibrium, itsatisfies v 2 − bv + I = 0, hence b 2 = 4I. Andronov-Hopf condition tr L = 2v − a = 0 results <strong>in</strong>v = a/2, hence a 2 /4 − ab/2 + I = 0. The bifurcation occurs when det L > 0, result<strong>in</strong>g <strong>in</strong> a < b.Comb<strong>in</strong><strong>in</strong>g the saddle-node and Andronov-Hopf conditions results <strong>in</strong> the Bogdanov-Takensconditions.14. Change of variables (6.5), v = x and u = √ µy, transforms the relaxation oscillator <strong>in</strong>to theformẋ = f(x) − √ (µyfẏ = √ with the Jacobian L =′ (b) − √ )µ√µ(x − b)µ 0at the equilibrium v = x = b, u = √ µy = f(b). The Andronov-Hopf bifurcation occurs whentr L = f ′ (b) = 0 and det L = µ > 0. Us<strong>in</strong>g (6.7), we f<strong>in</strong>d that it is supercritical when f ′′′ (b) < 0and subcritical when f ′′′ (b) > 0.15. The Jacobian matrix at the equilibrium, which satisfies F (v) − bv = 0, is( )F′−1L =.µb −µThe Andronov-Hopf bifurcation occurs when tr L = F ′ − µ = 0 (hence F ′ = µ) and det L =ω 2 = µb − µ 2 > 0 (hence b > µ). The change of variables (6.5), v = x and u = µx + ωy,transforms the system <strong>in</strong>to the formẋ = −ωy + f(x)ẏ = ωx + g(x)where f(x) = F (x) − µx and g(x) = µ[bx − F (x)]/ω. The result follows from (6.7).


Solutions to Exercises, Chap. 7 42716. The change of variables (6.5) converts the system <strong>in</strong>to the formThe result follows from (6.7).ẋ = F (x) + l<strong>in</strong>ear termsẏ = µ(G(x) − F (x))/ω + l<strong>in</strong>ear terms17. The change of variables (6.5), v = x and u = µx + ωy, converts the system <strong>in</strong>toThe result follows from (6.7).ẋ = F (x) − x(µx + ωy) + l<strong>in</strong>ear termsẏ = µ[G(x) − F (x) + x(µx + ωy)]/ω18. The system undergoes Andronov-Hopf bifurcation when F v = −µG u and F u G v < −µG 2 u. Weperform all the steps from (6.4) to (6.7) disregard<strong>in</strong>g l<strong>in</strong>ear terms (they do not <strong>in</strong>fluence a)and the terms of the order o(µ). Let ω = √ −µF u G v + O(µ), then u = (µG u x − ωy)/F u =−ωy/F u + O(µ), andandf(x, y) = F (x, −ωy/F u + O(µ)) = F (x, 0) − F u (x, 0)ωy/F u + O(µ)The result follows from (6.7).g(x, y) = (µ/ω)[G u F (x, 0) − F u G(x, 0)] + O(µ) .Solutions to Chapter 71. Take c < 0 so that the slow w-nullcl<strong>in</strong>e has a negative slope.2. The quasi-threshold conta<strong>in</strong>s the union of canard solutions.3. The change of variables z = e iωt u transforms the system <strong>in</strong>to the formwhich can be averaged, yield<strong>in</strong>g˙u = ε{−u + e −iωt I(t)} ,˙u = ε{−u + I ∗ (ω)} .Apparently, the stable equilibrium u = I ∗ (ω) corresponds to the susta<strong>in</strong>ed oscillation z =e iωt I ∗ (ω).4. The existence of damped oscillations with frequency ω implies that the system has a focusequilibrium with eigenvalues −ε ± iω, where ε > 0. The local dynamics near the focus can berepresented <strong>in</strong> the form (7.3). The rest of the proof is the same as the one for Ex. 3.5. Even though the slow and the fast nullcl<strong>in</strong>es <strong>in</strong> Fig. 5.21 <strong>in</strong>tersect only <strong>in</strong> one po<strong>in</strong>t, theycont<strong>in</strong>ue to be close and parallel to each other <strong>in</strong> the voltage range 10 mV to 30 mV. Such aproximity creates a tunnel<strong>in</strong>g effect (Rush and R<strong>in</strong>zel 1996) that prolongs the time spent nearthose nullcl<strong>in</strong>es.6. (Shilnikov-Hopf bifurcation) The model is near a co-dimension-2 bifurcation hav<strong>in</strong>g a homocl<strong>in</strong>icorbit to an equilibrium undergo<strong>in</strong>g subcritical Andronov-Hopf bifurcation, as we illustrate<strong>in</strong> Fig. 10.28. Many weird phenomena could happen near bifurcations of co-dimension 2or higher.


428 Solutions to Exercises, Chap. 8fast K +activationslow K +activationVhomocl<strong>in</strong>ic orbitAndronov-HopfFigure 10.28: Co-dimension-2 Shilnikov-Hopf bifurcation.Solutions to Chapter 81. Consider two mutually coupled neurons fir<strong>in</strong>g together.2. The equationcan be written <strong>in</strong> the formwithprovided that b < b sn .˙V = c(b − b sn ) + a(V − V sn ) 2 ,˙V = a(V − V rest )(V − V thresh ) .V rest = V sn − √ c(b sn − b)/a and V thresh = V sn + √ c(b sn − b)/a3. The system ˙v = b + v 2 with b > 0 and the <strong>in</strong>itial condition v(0) = v reset has the solution (checkby differentiat<strong>in</strong>g)v(t) = √ b tan( √ b(t + t 0 ))wheret 0 =1 √batan v reset√b.From the condition v(T ) = v peak = 1, we f<strong>in</strong>dT = √ 1 atan √ 1 − t 0 = 1 (√ atan √ 1 − atan v )reset√ ,b b b b bwhich can be alternatively written asT =1 √batan( )√b v reset − 1.v reset + b4. The system ˙v = −|b| + v 2 with the <strong>in</strong>itial condition v(0) = v reset > √ |b| has the solution (checkby differentiat<strong>in</strong>g)v(t) = √ |b| 1 + e2√ b(t+t 0 )1 − e ,2√ b(t+t 0)wheret 0 = 12 √ |b| ln v reset − √ |b|v reset + √ |b| .From the condition v(T ) = 1, we f<strong>in</strong>d(T = 12 √ ln 1 − √ |b||b| 1 + √ |b| − ln v reset − √ )|b|v reset + √ |b|.


Solutions to Exercises, Chap. 9 4295. The saddle-node bifurcation occurs when b = 0 regardless of the value of v reset , which is astraight vertical l<strong>in</strong>e <strong>in</strong> Fig. 8.3. If v reset < 0, then the saddle-node bifurcation is on an<strong>in</strong>variant circle. When b < 0, the unstable node (saddle) equilibrium is at v = √ |b|. Hence,the saddle homocl<strong>in</strong>ic orbit bifurcation occurs when v reset = √ |b|.6. The change of variables v = g/2 + V , b = g 2 /4 + B transforms ˙v = b − gv + v 2 to ˙V = B + V 2with V reset = −∞ and V peak = +∞. It has threshold V = √ B, rheobase B = 0 and thesame F-I curve as <strong>in</strong> the orig<strong>in</strong>al model with g = 0. In v-coord<strong>in</strong>ates, the threshold is v =g/2 + √ b − g 2 /4, which is greater than √ b, the new rheobase is b = g 2 /4, which is greater thanb = 0, and the new F-I curve is the same as the old one, just shifted to the right by g 2 /4.7. Let b = εr with ε ≪ 1 be a small parameter. The change of variablestransforms (8.2) <strong>in</strong>to the theta-neuron formv = √ ε tan ϑ 2˙ϑ = √ ε{(1 − cos ϑ) + (1 + cos ϑ)r} .uniformly on the unit circle except the small <strong>in</strong>terval |ϑ−π| < 2 4√ ε correspond<strong>in</strong>g to the actionpotential (v > 1); see Hoppensteadt and Izhikevich (1997) for more details.8. Use the change of variablesv =√ εϑ1 − |ϑ| .To obta<strong>in</strong> other theta neuron models, use the change of variablesv = √ εh(ϑ) ,where the monotone function h maps (−π, π) to (−∞, ∞) and scales like 1/(ϑ ± π) whenϑ → ±π. The correspond<strong>in</strong>g model has the formϑ ′ = h 2 (ϑ)/h ′ (ϑ) + r/h ′ (ϑ) .In particular, h 2 (ϑ)/h ′ (ϑ) exists and is bounded and 1/h ′ (ϑ) = 0 when ϑ → ±π. These implya uniform velocity <strong>in</strong>dependent from the <strong>in</strong>put r when ϑ passes the value ±π correspond<strong>in</strong>g tofir<strong>in</strong>g a spike.9. The equilibrium v = I/(b + 1), u = bI/(b + 1) has the Jacobian matrix( ) −1 −1L =ab −awith trL = −(a + 1) and detL = a(b + 1). It is a stable node (<strong>in</strong>tegrator) when b < (a +1) 2 /(4a) − 1 and a stable focus (resonator) otherwise.10. The quadratic <strong>in</strong>tegrate-and-fire neuron with a dendritic compartment˙V = B + V 2 + g 1 (V d − V )˙V d = g leak (E leak − V d ) + g 2 (V − V d )can be written <strong>in</strong> the form (8.3, 8.4), with v = V − g 1 /2, u = −g 1 V d , I = B − g 2 1/4 − (g 2 1g 2 +g leak E leak )/(g leak + g 2 ), a = g leak + g 2 , and b = −g 1 g 2 /a.11. A MATLAB program generat<strong>in</strong>g the figure is provided on the author’s webpage.12. An example is <strong>in</strong> Fig. 10.29.13. An example is <strong>in</strong> Fig. 10.30.


430 Solutions to Exercises, Chap. 9FS neuron (<strong>in</strong> vitro)simple modelI=400 pA121I=230 pA21I=125 pA25 mV25 ms-55 mVresetAHPrecovery variable, u4003002001002 3 resetAHPv-nullcl<strong>in</strong>ev-nullcl<strong>in</strong>e, I=0u-nullcl<strong>in</strong>eI=100 pAspike0rest-60 -40 -20 0 20membrane potential, v (mV)Figure 10.29: Comparison of <strong>in</strong> vitro record<strong>in</strong>gs of a fast spik<strong>in</strong>g (FS) <strong>in</strong>terneuronof layer 5 rat’s visual cortex with simulations of the simple model with l<strong>in</strong>ear slownullcl<strong>in</strong>e 20 ˙v = (v + 55)(v + 40) − u + I, ˙u = 0.15{8(v + 55) − u}, if v ≥ 25, thenv ← −55, u ← u + 200.


Solutions to Exercises, Chap. 9 431layer 5 neuronsimple modelI=80 pA0I=50 pA0I=45 pA0I=40 pArecovery, u100u=I-5(v+60)500u=-2(v+60)u=I-5(v+60)+3(v+50) 2-50-60 -55 -50 -45 -40 -35membrane potential, v (mV)Figure 10.30: Comparison of <strong>in</strong> vitro record<strong>in</strong>gs of a regular spik<strong>in</strong>g (RS) neuronwith simulations of the simple model 100 ˙v = I − 5(v + 60) + 3(v + 50) 2 + − u, ˙u =0.02{−2(v + 60) − u}, if v ≥ 35, then v ← −50, u ← u + 70.


432 Solutions to Exercises, Chap. 90.4u0.200.20.40.60.8x12 1.5 1 0.5 0 0.5 1 1.5 2 2.5 332.521.510.500.511.5x(t)20 50 100 150 200 250 300 350 400 450Figure 10.31: Solution to Ex. 1. Nullcl<strong>in</strong>es,hedgehog limit cycle and a burst<strong>in</strong>gsolution of a planar system (modifiedfrom Izhikevich 2000).1K + activation gate, n0.80.60.40.20-80 -60 -40 -20 0membrane potential, V (mV)Figure 10.32: Noise-<strong>in</strong>duced burst<strong>in</strong>g<strong>in</strong> two-dimensional system; See Ex. 2.Solutions to Chapter 91. (Planar burster) Izhikevich (2000) suggested the systemẋ = x − x 3 /3 − u + 4S(x) cos 40u,˙u = µx ,with S(x) = 1/(1 + e 5(1−x) ) and µ = 0.01. It has a hedgehog limit cycle depicted <strong>in</strong> Fig. 10.31.2. (Noise-<strong>in</strong>duced burst<strong>in</strong>g) Noise can <strong>in</strong>duce burst<strong>in</strong>g <strong>in</strong> a two-dimensional system with coexistenceof rest<strong>in</strong>g and spik<strong>in</strong>g states. Indeed, noisy perturbations can randomly push thestate of the system <strong>in</strong>to the attraction doma<strong>in</strong> of the rest<strong>in</strong>g state or <strong>in</strong>to the attraction doma<strong>in</strong>of the limit cycle attractor, as <strong>in</strong> Fig. 10.32. The solution meanders between the states,exhibit<strong>in</strong>g a random burst<strong>in</strong>g pattern as <strong>in</strong> Fig. 9.55,right. neocortical neurons of RS and FStype, as well as stellate neurons of the entorh<strong>in</strong>al cortex exhibit such burst<strong>in</strong>g; see Chap. 8.


Solutions to Exercises, Chap. 9 4330.800.7-10K + gat<strong>in</strong>g variable, n0.60.50.40.30.2membrane potential, V (mV)-20-30-40-500.1-600-80 -60 -40 -20 0membrane potential, V (mV)-700 10 20 30 40 50 60 70time (ms)Figure 10.33: Noise-<strong>in</strong>duced burst<strong>in</strong>g a two-dimensional system with co-existence of an equilibriumand a limit cycle attractor; see Ex. 3.0.2w0.1ghost offold limit cycle10.50V(t)0-0.5 0 0.5 1VI(t)-0.50 1000 2000 3000 4000timeFigure 10.34: Rebound burst<strong>in</strong>g <strong>in</strong> the FitzHugh-Nagumo oscillator; see Ex. 4.3. (Noise-<strong>in</strong>duced burst<strong>in</strong>g) Burst<strong>in</strong>g occurs because noisy perturbations push the trajectory <strong>in</strong>toand out of the attraction doma<strong>in</strong> of the limit cycle attractor, which coexists with the rest<strong>in</strong>gequilibrium; see the phase portrait <strong>in</strong> Fig. 10.33.4. (Rebound burst<strong>in</strong>g <strong>in</strong> the FitzHugh-Nagumo oscillator) The oscillator is near fold limit cyclebifurcation. The solution makes a few rotations along the ghost of the cycle before return<strong>in</strong>gto rest; see Fig. 10.34.5. Yes, they can, at the end of a burst. Th<strong>in</strong>k of a “fold/Hopf” or “circle/Hopf” burster. Therest<strong>in</strong>g equilibrium is a stable focus right after the term<strong>in</strong>ation of a burst, and then it istransformed <strong>in</strong>to a stable node to be ready for the circle or fold bifurcation. Even “circle/circle”bursters could exhibit such oscillations, if the rest<strong>in</strong>g equilibrium turns <strong>in</strong>to a focus for a shortperiod of time somewhere <strong>in</strong> the middle of a quiescent phase. In any case, the oscillationsshould disappear just before the transition to the spik<strong>in</strong>g state.6. (Hopf/Hopf burst<strong>in</strong>g) Even though there is no co-existence of attractors, there is a hysteresisloop due to the slow passage effect through the supercritical Andronov-Hopf bifurcation; seeFig. 10.35. The delayed transition to spik<strong>in</strong>g creates the hysteresis loop and enables burst<strong>in</strong>g.7. (Hopf/Hopf canonical model) First, we restrict the fast subsystem to its center manifold andtransform it to the normal form for supercritical Andronov-Hopf bifurcation, which after appropriatere-scal<strong>in</strong>g, has the formż = (u + iω)z − z|z| 2 .


434 Solutions to Exercises, Chap. 90.5|x|yx1x2Syy0.5x 1Supercritical Andronov-Hopf BifurcationtSlow PassageEffectFigure 10.35: Hopf/Hopf burst<strong>in</strong>g without co-existence of attractors; see Ex. 6 (modified fromHoppensteadt and Izhikevich 1997).Here, u is the deviation from the slow equilibrium u 0 . The slow subsystem˙u = µg(ze iωt + complex-conjugate, u)can be averaged and transformed <strong>in</strong>to the canonical form.8. (Burst<strong>in</strong>g <strong>in</strong> the I Na,t +I Na,slow -model) First, determ<strong>in</strong>e the parameters of the I Na,t -model correspond<strong>in</strong>gto the subcritical Andronov-Hopf bifurcation, and hence the co-existence of therest<strong>in</strong>g and spik<strong>in</strong>g states. Then, add a slow high-threshold persistent Na + current that activatesdur<strong>in</strong>g spik<strong>in</strong>g, depolarizes the membrane potential and stops the spik<strong>in</strong>g. Dur<strong>in</strong>g rest<strong>in</strong>g,the current deactivates, the membrane potential hyperpolarizes and the neuron starts to fireaga<strong>in</strong>.9. Substitute the slow Na + current <strong>in</strong> the exercise above with a slow dendritic compartmentwith dendritic rest<strong>in</strong>g potential far below the somatic rest<strong>in</strong>g potential. As the dendriticcompartment hyperpolarizes the somatic compartment, the soma starts to fire (due to the<strong>in</strong>hibition-<strong>in</strong>duced fir<strong>in</strong>g described <strong>in</strong> Sect. 7.2.8). As the somatic compartment fires, dendriticcompartment slowly depolarizes, removes the hyperpolarization and stops fir<strong>in</strong>g.10. (Burst<strong>in</strong>g <strong>in</strong> the I Na,p +I K +I Na,slow -model) The time constant τ slow (V ) is relatively small <strong>in</strong>the voltage range correspond<strong>in</strong>g to the spike after-hyperpolarization (AHP). Deactivation ofthe Na + current dur<strong>in</strong>g each AHP is much stronger than its activation dur<strong>in</strong>g the spike peak.As a result, Na + current deactivates (turns off) dur<strong>in</strong>g the burst, and then slowly reactivatesto its basel<strong>in</strong>e level dur<strong>in</strong>g the rest<strong>in</strong>g period, as one can see <strong>in</strong> Fig. 10.36.11. The mechanism of spik<strong>in</strong>g, illustrated <strong>in</strong> Fig. 10.37, is closely related to the phenomenon ofaccommodation and anodal break excitation. The key feature is that this burst<strong>in</strong>g is notfast-slow.


Solutions to Exercises, Chap. 9 435membrane potential,V (mV)0-20-40-600 50 100 150 200 250 300 350 400 450Na + activationgate, mslow0.50.4deactivationreactivation0.30 50 100 150 200 250 300 350 400 450time (ms)Figure 10.36: Burst<strong>in</strong>g <strong>in</strong> the I Na,p +I K +I Na,slow -model. See Ex. 10.slow <strong>in</strong>crease of Ifast <strong>in</strong>crease of I2.52.5221.51.5110.50.50.50Rest0.50Spik<strong>in</strong>g12.5 2 1.5 1 0.5 0 0.5 1 1.5 2 2.512.5 2 1.5 1 0.5 0 0.5 1 1.5 2 2.5Figure 10.37: The system has a unique attractor — equilibrium, yet it can exhibit repetitive spik<strong>in</strong>gactivity when the N-shaped nullcl<strong>in</strong>e is moved up not very slow.


436 Solutions to Exercises, Chap. 9Saddle-Node onInvariant CircleBifurcationSpik<strong>in</strong>gSaddle-Node onInvariant CircleBifurcationRestψθFigure 10.38: Answer to Ex. 13.The system has a unique attractor — a stable equilibrium, and the solution always converges toit. The slow variable I controls the vertical position of the N-shaped nullcl<strong>in</strong>e. If I <strong>in</strong>creases, thenullcl<strong>in</strong>e moves up slowly, and so does the solution because it tracks the equilibrium. However,if the rate of change of I is not small enough, the solution cannot catch up with the equilibriumand starts to oscillate with a large amplitude. Thus, the system exhibits spik<strong>in</strong>g behavior eventhough it does not have a limit cycle attractor for any fixed I.12. From the first equation, we f<strong>in</strong>d the equivalent voltage{|z| 2 1 + u if 1 + u > 0 ,= |1 + u| + =0 if 1 + u ≤ 0 ,so that the reduced slow subsystem has the form˙u = µ[u − u 3 − w] ,ẇ = µ[|1 + u| + − 1] ,and it has essentially the same dynamics as the van der Pol oscillator.13. The fast equation˙ϑ = 1 − cos ϑ + (1 + cos ϑ)ris the Ermentrout-Kopell canonical model for Class 1 excitability, also known as the thetaneuron (Ermentrout 1996). It is quiescent when r < 0 and fires spikes when r > 0. As ψoscillates with frequency ω, the function r = r(ψ) changes sign. The fast equation undergoes asaddle-node on <strong>in</strong>variant circle bifurcation, hence the system is a “circle/circle” burster of theslow-wave type; see Fig. 10.38.14. To understand the burst<strong>in</strong>g dynamics of the canonical model, we rewrite it <strong>in</strong> polar coord<strong>in</strong>atesz = re iϕ :ṙ = ur + 2r 3 − r 5 ,˙u = µ(a − r 2 ) ,˙ϕ = ω .Apparently, it is enough to consider the first two equations, which determ<strong>in</strong>e the oscillationprofile. Nontrivial (r ≠ 0) equilibria of this system correspond to limit cycles of the canonicalmodel, which may look like periodic (tonic) spik<strong>in</strong>g with frequency ω. Limit cycles of thissystem correspond to quasi-periodic solutions of the canonical model, which look like burst<strong>in</strong>g;see Fig. 9.37.The first two equation above have a unique equilibrium( ) ( √ )r a=u a 2 − 2a


Solutions to Exercises, Chap. 9 437Saddle-NodeSeparatrix LoopSaddle-Nodeon InvariantCircleSpikeInvariantFoliationFigure 10.39: A small neighborhood of the saddle-node po<strong>in</strong>t can be <strong>in</strong>variantly foliated by stablesubmanifolds.for all µ and a > 0, which is stable when a > 1. When a decreases and passes an µ-neighborhoodof a = 1, the equilibrium loses stability via Andronov-Hopf bifurcation. When 0 < a < 1, thesystem has a limit cycle attractor. Therefore, the canonical model exhibits burst<strong>in</strong>g behavior.The smaller the value of a, the longer the <strong>in</strong>terburst period. When a → 0, the <strong>in</strong>terburst periodbecomes <strong>in</strong>f<strong>in</strong>ite.15. Take w = I − u. Then (9.7) becomes˙v = v 2 + w ,ẇ = µ(I − w) ≈ µI .16. Let us sketch the derivation. S<strong>in</strong>ce the fast subsystem is near saddle-node homocl<strong>in</strong>ic orbitbifurcation for some u = u 0 , a small neighborhood of the saddle-node po<strong>in</strong>t v 0 is <strong>in</strong>variantlyfoliated by stable submanifolds, as <strong>in</strong> Fig. 10.39. Because the contraction along the stablesubmanifolds is much stronger than the dynamics along the center manifold, the fast subsystemcan be mapped <strong>in</strong>to the normal form ˙v = q(u) + p(v − v 0 ) 2 by a cont<strong>in</strong>uos change of variables.When v escapes the small neighborhood of v 0 , the neuron is said to fire a spike, and v is resetv ← v 0 + c(u). Such a stereotypical spike also resets u by a constant d. If g(v 0 , u 0 ) ≈ 0, thenall functions are small, and l<strong>in</strong>earization and appropriate re-scal<strong>in</strong>g yields the canonical model.If g(v 0 , u 0 ) ≠ 0, then the canonical model has the same form as <strong>in</strong> the previous exercise.17. The derivation proceeds as <strong>in</strong> the previous exercise, yield<strong>in</strong>g˙v = I + v 2 + (a, u) ,˙u = µAu .where (a, u) is the scalar (dot) product of vectors a, u ∈ R 2 , and A is the Jacobian matrixat the equilibrium of the slow subsystem. If the equilibrium is a node, it has generically twodist<strong>in</strong>ct eigenvalues, and two real eigenvectors. In this case, the slow subsystem uncouples <strong>in</strong>totwo equations, each along the correspond<strong>in</strong>g eigenvector. Appropriate re-scal<strong>in</strong>g gives the firstcanonical model. If the equilibrium is a focus, the l<strong>in</strong>ear part can be made triangular to getthe second canonical model.18. The solution of the fast subsystem˙v = u + v 2 , v(0) = −1 ,with fixed u > 0 isv(t) = √ ( )√ut 1u tan − atan √u


438 Solutions to Exercises, Chap. 94slow variable u220-2averagedfull-4-4 -2 0 2 4slow variable u 1Figure 10.40: Ex. 18.The <strong>in</strong>terspike period, T , is def<strong>in</strong>ed by v(T ) = +∞, given by the formulaT (u) = √ 1 ( πu 2 + atan √ 1 ). uThe result follows from the <strong>in</strong>tegraland the relationships1T (u)∫ T (u)0d i δ(t − T (u)) dt = d 1T (u)f(u) = 1T (u)and atan 1 √ u= arcot √ u .Periodic solutions of the averaged system (focus case) and the full system are depicted <strong>in</strong>Fig. 10.40. The deviation is due to the f<strong>in</strong>ite size of the parameters µ 1 and µ 2 <strong>in</strong> Fig. 9.35.19. There are only two co-dimension-1 bifurcations of an equilibrium that result <strong>in</strong> transitions toanother equilibrium: saddle-node off limit cycle and subcritical Andronov-Hopf bifurcation.Hence, there are four po<strong>in</strong>t-po<strong>in</strong>t hysteresis loops, depicted <strong>in</strong> Fig. 10.41. More details areprovided <strong>in</strong> Izhikevich (2000).20. This figures are modified from (Izhikevich 2000), where one can f<strong>in</strong>d two models exhibit<strong>in</strong>gthis phenomenon. The key feature is that the slow subsystem is not too slow, and the rateof attraction to the upper equilibrium is relatively weak. The spikes are actually dampedoscillations that are generated by the fast subsystem while it converges to the equilibrium.Periodic burst<strong>in</strong>g is generated via the po<strong>in</strong>t-po<strong>in</strong>t hysteresis loop.21. There are only two co-dimension-1 bifurcations of a small limit cycle attractor (subthresholdoscillation) on a plane that result <strong>in</strong> sharp transitions to a large-amplitude limit cycle attractor(spik<strong>in</strong>g): Fold limit cycle bifurcation and saddle-homocl<strong>in</strong>ic orbit bifurcation; see Fig. 10.42.These two bifurcations paired with any of the four bifurcations of the large-amplitude limitcycle attractor result <strong>in</strong> 8 planar co-dimension-1 cycle-cycle bursters; see Fig. 10.43. Moredetails are provided by Izhikevich (2000).


Solutions to Exercises, Chap. 9 439FoldBifurcationFoldBifurcationFoldBifurcationSaddle Homocl<strong>in</strong>icOrbit BifurcationSubcriticalAndronov-HopfBifurcation"Fold/Fold"Hysteresis Loop"Fold/SubHopf"Hysteresis LoopSubcriticalAndrono-HopfBifurcationSaddle Homocl<strong>in</strong>icOrbit BifurcationFoldBifurcationSubcriticalAndrono-HopfBifurcationSaddleHomocl<strong>in</strong>icOrbitBifurcationsSubcriticalAndronov-HopfBifurcation"SubHopf/Fold"Hysteresis Loop"SubHopf/SubHopf"Hysteresis LoopFigure 10.41: Classification of po<strong>in</strong>t-po<strong>in</strong>t co-dimension-1 hysteresis loops.Fold Limit Cycle BifurcationSaddle Homocl<strong>in</strong>ic Orbit BifurcationFigure 10.42: Co-dimension-1 bifurcations of a stable limit cycle <strong>in</strong> planar systems that result <strong>in</strong>sharp loss of stability and transition to a large-amplitude (spik<strong>in</strong>g) limit cycle attractor, not shown<strong>in</strong> the figure. Fold limit cycle: Stable limit cycle is approached by an unstable one, they coalesce, andthen disappear. Saddle homocl<strong>in</strong>ic orbit: A limit cycle grows <strong>in</strong>to a saddle. The unstable manifold ofthe saddle makes a loop and returns via the stable manifold (separatrix).


440 Solutions to Exercises, Chap. 9Saddle-Node Saddle Supercritical FoldBifurcations on Invariant Homocl<strong>in</strong>ic Andronov- LimitCircle Orbit Hopf CycleFold fold cycle/ fold cycle/ fold cycle/ fold cycle/Limit circle homocl<strong>in</strong>ic Hopf fold cycleCycleSaddle homocl<strong>in</strong>ic/ homocl<strong>in</strong>ic/ homocl<strong>in</strong>ic/ homocl<strong>in</strong>ic/Homocl<strong>in</strong>ic circle homocl<strong>in</strong>ic Hopf fold cycleOrbitFigure 10.43: Classification of co-dimension-1 cycle-cycle planar bursters.


ReferencesAcebron J. A., Bonilla L.L., V<strong>in</strong>cente C.J. P., Ritort F., Spigler R. (2005) The Kuramoto model:a simple paradigm for synchronization phenomena. Review of Modern Physics. 77:137–185Alonso A. and Kl<strong>in</strong>k R. (1993) Differential Electroresponsiveness of Stellate and Pyramidal-LikeCells of Medial Entorh<strong>in</strong>al Cortex Layer II. Journal of Neurophysiology, 70:128–143.Alonso A and Ll<strong>in</strong>as R. R. (1989) Subthreshold Na + -dependent theta-like rhythmicity <strong>in</strong> stellatecells of entorh<strong>in</strong>al cortex layer II. Nature, 342:175-177.Am<strong>in</strong>i, B., Clark J. W. Jr., and Canavier C. C. (1999) Calcium Dynamics Underly<strong>in</strong>g Pacemaker-Like and Burst Fir<strong>in</strong>g Oscillations <strong>in</strong> Midbra<strong>in</strong> Dopam<strong>in</strong>ergic Neurons: A ComputationalStudy. J. Neurophysiol. 82:2249–2261,Amir R., Michaelis M., Devor M. (2002) Burst Discharge <strong>in</strong> Primary Sensory Neurons: Triggeredby Subthreshold Oscillations, Ma<strong>in</strong>ta<strong>in</strong>ed by Depolariz<strong>in</strong>g Afterpotentials. Journal of<strong>Neuroscience</strong>, 22:1187–1198.Armstrong C.M. and Hille B. (1998) Voltage-Gated Ion Channels and Electrical Excitability.Neuron 20:371–380.Arnold V.I. (1965) Small denom<strong>in</strong>ators. I. Mapp<strong>in</strong>gs of the circumference onto itself. Trans. ofthe AMS Series 2, 46:213–284.Arnold V.I. (1982) Geometrical Methods <strong>in</strong> the Theory of Ord<strong>in</strong>ary Differential Equations.Spr<strong>in</strong>ger-Verlag, New York, [Russian orig<strong>in</strong>al: Additional Chapters of the Theory of Ord<strong>in</strong>aryDifferential Equations, Moscow 1977].Arnold V.I., Afrajmovich V.S., Il’yashenko Yu.S., Shil’nikov L.P. (1994) <strong>Dynamical</strong> <strong>Systems</strong> V.Bifurcation Theory and Catastrophe Theory. Spr<strong>in</strong>ger-Verlag, Berl<strong>in</strong> Heidelberg.Aronson D.G., Ermentrout G.B., and Kopell N. (1990) Amplitude response of coupled oscillators.Physica D 41:403–449.Arshavsky Y.I., Berk<strong>in</strong>blit M.B., Kovalev S.A., Smolyan<strong>in</strong>ov V.V., and Chaylakhyan L.M.(1971) An analysis of the functional properties of dendrites <strong>in</strong> relation to their structure, <strong>in</strong>:I.M. Gelfand, Gurf<strong>in</strong>kel, S.V. Fom<strong>in</strong>, M.L. Zetl<strong>in</strong> (Eds.), Models of the Structural and FunctionalOrganization of Certa<strong>in</strong> Biological <strong>Systems</strong>, MIT Press, Cambridge, Massachusetts,1971, pp. 25-71.Bacci, A., Rudolph, U., Huguenard, J.R. and Pr<strong>in</strong>ce, D.A. (2003) Major differences <strong>in</strong> <strong>in</strong>hibitorysynaptic transmission onto two neocortical <strong>in</strong>terneuron subclasses. J. Neurosc. 23:9664–9674.Bacci A., Huguenard J.R., Pr<strong>in</strong>ce D.A. (2003) Long-last<strong>in</strong>g self-<strong>in</strong>hibition of neocortical <strong>in</strong>-441


442 Referencesterneurons mediated by endocannab<strong>in</strong>oids. nature, 431:312–316.Baer S.M., Erneux T., and R<strong>in</strong>zel J. (1989) The slow passage through a Hopf bifurcation: delay,memory effects, and resonances. SIAM Journal on Applied Mathematics 49:55–71.Baer S.M., R<strong>in</strong>zel J., and Carrillo H. (1995) Analysis of an autonomous phase model for neuronalparabolic burst<strong>in</strong>g. Journal of Mathematical Biology 33:309–333.Baesens, C., Guckenheimer, J., Kim, S. and MacKay, R. S. (1991) Three coupled oscillators:mode-lock<strong>in</strong>g, global bifurcations and toroidal chaos. Physica D, 49, 387–475.Baker M.D. and Bostock H. (1997) Low-threshold, Persistent Sodium Current <strong>in</strong> Rat LargeDorsal Root Ganglion Neurons <strong>in</strong> Culture, J. Neurophysiol.,77:1503-1513.Baker M.D. and Bostock H. (1998) Inactivation of Macroscopic Late Na + Current and Characteristicsof Unitary Late Na + Currents <strong>in</strong> Sensory Neurons, J. Neurophysiol., 80:2538-2549.Baut<strong>in</strong> N.N. (1949) The behavior of dynamical systems near to the boundaries of stability.Gostekhizdat., Moscow-Len<strong>in</strong>grad, 2nd edition, Nauka, Moscow (1984).Beierle<strong>in</strong> M., Gibson J.R., and Connors B.W. (2003) Two <strong>Dynamical</strong>ly Dist<strong>in</strong>ct InhibitoryNetworks <strong>in</strong> Layer 4 of the Neocortex. Journal of Neurophysiology, 90:2987–3000.Bekkers J.M. (2000) Properties of voltage-gated potassium currents <strong>in</strong> nucleated patches fromlarge layer 5 pyramidal neurons of the rat. Journal of Physiology, 525.3:593–609Benoit E. (1984) Canards de R 3 . These d’etat, Universite de Nice.Benoit E., Callot J.-L., Diener F., Diener M. (1981) Chasse au canards. Collectanea Mathematica,31–32(1–3):37–119.Bertram R., Butte M.J., Kiemel T., and Sherman A. (1995) Topological and phenomenologicalclassification of burst<strong>in</strong>g oscillations. Bullet<strong>in</strong> of Mathematical Biology 57:413–439.Blechman I.I. (1971) Synchronization of <strong>Dynamical</strong> <strong>Systems</strong>.D<strong>in</strong>amicheskich Sistem”, Science, Moscow].[<strong>in</strong> Russian: “S<strong>in</strong>chronizatziaBower J. M. and Beeman D. (1995) The Book of GENESIS. New York Spr<strong>in</strong>ger-Verlag.Bressloff P.C. and Coombes S. (2000) Dynamics of strongly coupled spik<strong>in</strong>g neurons. NeuralComputation, 12:91–129.Brizzi L., Meunier C., Zytnicki D., Donnet M., Hansel D., Lamotte d’Incamps B., and vanVreeswijk C. (2004) How shunt<strong>in</strong>g <strong>in</strong>hibition affects the discharge of lumbar motoneurons: adynamic clamp study <strong>in</strong> anaesthetized cats. J. Physiol, 558:671–683.Brown E., Moehlis J., Holmes P. (2004) On the phase reduction and response dynamics of neuraloscillator populations. Neural Computation 16:673–715.Brown T.G. (1911) The <strong>in</strong>tr<strong>in</strong>sic factors <strong>in</strong> the act of progression <strong>in</strong> the mammal. Proc. Roy.Soc. Lond. B, 84:308–319Brunel N., Hakim V., and Richardson M.J. (2003) Fir<strong>in</strong>g-rate resonance <strong>in</strong> a generalized<strong>in</strong>tegrate-and-fire neuron with subthreshold resonance. Phys Rev E, 67:051916.Bryant H.L., Segundo J.P. (1976) Spike <strong>in</strong>itiation by transmembrane current: a white-noiseanalysis. J. Physiol, 260:279-314.Butera, R. J., R<strong>in</strong>zel J., and Smith J. C. (1999) Models of Respiratory Rhythm Generation <strong>in</strong>the Pre-Botz<strong>in</strong>ger Complex. I. Burst<strong>in</strong>g Pacemaker Neurons. J. Neurophysiol. 82: 382-397.


References 443Canavier C. C. , Clark J. W. and Byrne J. H. (1991) Simulation of the burst<strong>in</strong>g activity ofneuron R15 <strong>in</strong> Aplysia: role of ionic currents, calcium balance, and modulatory transmitters.J Neurophysiol 66:2107–2124Cartwright M.L. and Littlewood J.E. (1945) On Nonl<strong>in</strong>ear Differential Equations of the SecondOrder: I. The Equation ÿ − k(1 − y 2 )ẏ + y = bλk cos(λt + α), k large. J. London Math Soc.20:180-189.del Castillo J. and Morales T. (1967) The Electrical and Mechanical Activity of the EsophagealCell of Ascaris lumbricoides. The Journal of General Physiology, 50:603–629.Chay, T. R., and D. L. Cook. (1988) Endogenous burst<strong>in</strong>g patterns <strong>in</strong> excitable cells. Math.Biosci. 90:139 -153.Chay T.R. and Keizer J. (1983) M<strong>in</strong>imal Model for Membrane Oscillations <strong>in</strong> the Pancreaticβ-cell. Biophys. J. 42:181–190Chow C.C. and Kopell N. (2000) Dynamics of Spik<strong>in</strong>g Neurons with Electrical Coupl<strong>in</strong>g. NeuralComputation, 12:1643–1678Clay J. R. (1998) Excitability of the Squid Giant Axon Revisited. J. Neurophysiol. 80: 903-913.Cohen A.H., Holmes P.J., and Rand R.H. (1982) The nature of the coupl<strong>in</strong>g between segmentaloscillators of the lamprey sp<strong>in</strong>al generator for locomotion: A mathematical model. Journalof Mathematical Biology, 13:345–369.Cole K.S., Guttman R., and Bezanilla F. (1970) Nerve excitation without threshold. PNAS,65:884–891.Coll<strong>in</strong>s J.J. and Stewart I. (1994) A group-theoretic approach to r<strong>in</strong>gs of coupled biologicaloscillators. Biol. Cybern. 71:95–103Coll<strong>in</strong>s, J. J. and Stewart, I. (1993) Coupled nonl<strong>in</strong>ear oscillators and the symmetries of animalgaits. J. Nonl<strong>in</strong>. Sci. 3:349–392.Connor J.A., Walter D. and McKown R. (1977) Modifications of the Hodgk<strong>in</strong>-Huxley AxonSuggested by Experimental Results from Crustacean Axons, Biophysical Journal 18:81–102.Connors B.W. and Gutnick M.J. (1990) Intr<strong>in</strong>sic fir<strong>in</strong>g patterns of diverse neocortical neurons,Trends <strong>in</strong> <strong>Neuroscience</strong>, 13:99–104.Daido H. (1996) Onset of cooperative entra<strong>in</strong>ment <strong>in</strong> limit-cycle oscillators with uniform all-toall<strong>in</strong>teractions: bifurcation of the order function. Physica D, 91:24–66.Del Negro C.A., Hsiao C.-F. Chandler S.H., and Garf<strong>in</strong>kel A. (1998) Evidence for a NovelBurst<strong>in</strong>g Mechanism <strong>in</strong> Rodent Trigem<strong>in</strong>al Neurons. Biophys Journal, 75:174–182.Denjoy A. (1932) Sur les courbes def<strong>in</strong>ies par les equations differentielles a la surface du tore.J. Math. Pures et Appl. 11:333-375.Destexhe A. , Contreras D., Sejnowski T. J., and Steriade M. (1994) A model of sp<strong>in</strong>dle rhythmicity<strong>in</strong> the isolated thalamic reticular nucleus. J Neurophysiol, 72: 803–818.Destexhe A. and Gaspard P. (1993) Burst<strong>in</strong>g oscillations from a homocl<strong>in</strong>ic tangency <strong>in</strong> a timedelay system. Phys. Lett. A. 173:386–391.Destexhe A., Rudolph M., Fellous J.M., and Sejnowski T.J. (2001) Fluctuat<strong>in</strong>g synaptic conductancesrecreate <strong>in</strong> vivo-like activity <strong>in</strong> neocortical neurons. <strong>Neuroscience</strong>, 107: 13–24.


444 ReferencesDickson C.T., Magistretti J., Shal<strong>in</strong>sky M.H., Fransen E., Hasselmo M.E., Alonso A. (2000)Properties and Role of I h <strong>in</strong> the Pac<strong>in</strong>g of Subthreshold Oscillations <strong>in</strong> Entorh<strong>in</strong>al CortexLayer II Neurons. Journal of Neurophysiology, 83:2562–2579.Doiron B., La<strong>in</strong>g C., Longt<strong>in</strong> A/, and Maler L. (2002) Ghostburst<strong>in</strong>g: a novel neuronal burstmechanism, J. Comput. Neurosci. 12:5–25.Dong C.-J. and Werbl<strong>in</strong> F.S. (1995) Inwardly Rectify<strong>in</strong>g Potassium Conductance Can Acceleratethe Hyperpolarization Response <strong>in</strong> Ret<strong>in</strong>al Horizontal Cells. Journal of Neurophysiology,74:2258–2265Eckhaus W. (1983) Relaxation Oscillations Includ<strong>in</strong>g a Standard Chase of French Ducks. LectureNotes <strong>in</strong> Math., 985:432–449.Ermentrout G. B. (1981) n : m Phase-Lock<strong>in</strong>g of Weakly Coupled Oscillators.Mathematical Biology, 12:327–342.Journal ofErmentrout G. B. (1985) Los<strong>in</strong>g Amplitude and Sav<strong>in</strong>g Phase. In Othmer H. G. (ed) Nonl<strong>in</strong>earOscillations <strong>in</strong> Biology and Chemistry. Spr<strong>in</strong>ger-Verlag.Ermentrout G. B. (1992) Stable Periodic Solutions to Discrete and Cont<strong>in</strong>uum Arrays of WeaklyCoupled Nonl<strong>in</strong>ear Oscillators. SIAM Journal on Applied Mathematics, 52:1665–1687.Ermentrout G. B. (1994) An Introduction to Neural Oscillators. In Neural Model<strong>in</strong>g and NeuralNetworks, F Ventriglia, (ed.), Pergamon Press, Oxford, pp.79–110.Ermentrout G. B. (1996) Type I Membranes, Phase Resett<strong>in</strong>g Curves, and Synchrony. NeuralComputation 8:979–1001.Ermentrout G. B. (1998) L<strong>in</strong>earization of F-I Curves by Adaptation.10:1721–1729.Neural ComputationErmentrout, G. B. (2003) <strong>Dynamical</strong> Consequences of Fast-Ris<strong>in</strong>g, Slow-Decay<strong>in</strong>g Synapses <strong>in</strong>Neuronal Networks Neural Computation 15:2483-2522Ermentrout G. B. (2002) Simulat<strong>in</strong>g, Analyz<strong>in</strong>g, and Animat<strong>in</strong>g <strong>Dynamical</strong> <strong>Systems</strong>: A Guideto XPPAUT for Researchers and Students (Software, Environments, Tools). SIAMErmentrout G. B. and Kopell N. (1984) Frequency Plateaus <strong>in</strong> a Cha<strong>in</strong> of Weakly CoupledOscillators, I. SIAM Journal on Applied Mathematics 15:215–237.Ermentrout G. B. and Kopell N. (1986a) Parabolic Burst<strong>in</strong>g <strong>in</strong> an Excitable System CoupledWith a Slow Oscillation. SIAM Journal on Applied Mathematics 46:233–253.Ermentrout G. B. and Kopell N. (1986b) Subcellular Oscillations and Burst<strong>in</strong>g. MathematicalBiosciences 78:265–291.Ermentrout G. B. and Kopell N. (1990) Oscillator death <strong>in</strong> systems of coupled neural oscillators.SIAM Journal on Applied Mathematics 50:125–146.Ermentrout G. B. and Kopell N. (1991) Multiple pulse <strong>in</strong>teractions and averag<strong>in</strong>g <strong>in</strong> systems ofcoupled neural oscillators. Journal of Mathematical Biology 29:195–217.Ermentrout G. B. and Kopell N. (1994) Learn<strong>in</strong>g of phase lags <strong>in</strong> coupled neural oscillators.Neural Computation 6:225–241.FitzHugh R. (1955) Mathematical Models of Threshold Phenomena <strong>in</strong> the Nerve Membrane,Bullet<strong>in</strong> of Mathematical Biophysics, 7:252-278.


References 445FitzHugh R. (1960) Threshold and Plateaus <strong>in</strong> the Hodgk<strong>in</strong>-Huxley Nerve Equations.Journal of General Physiology, 43:867–896.FitzHugh, R.A. (1961) Impulses and physiological states <strong>in</strong> theoretical models of nerve membrane.Biophys. J., 1:445–466.FitzHugh R. (1969) Mathematical Models of Excitation and Propagation <strong>in</strong> Nerve. In Schwan(ed.) Biological Eng<strong>in</strong>eer<strong>in</strong>g, McGraw-Hill Book Co., Inc., N.Y.FitzHugh R. (1976) Anodal excitation <strong>in</strong> the Hodgk<strong>in</strong>-Huxley nerve model. Biophys J. 16:209–226.Fourcaud-Trocme N., Hansel D., van Vreeswijk C., and Brunel N. (2003) How Spike GenerationMechanisms Determ<strong>in</strong>e the Neuronal Response to Fluctuat<strong>in</strong>g Inputs. The Journal of<strong>Neuroscience</strong>, 23(37):11628–11640.Frankel P. and Kiemel T. (1993) Relative Phase Behavior of Two Slowly Coupled Oscillators.SIAM J. Appl. Math., 53, pp. 1436–1446.Geiger J.R.P. and Jonas P. (2000) Dynamic Control of Presynaptic Ca Inflow by Fast-Inactivat<strong>in</strong>gK Channels <strong>in</strong> Hippocampal Mossy Fiber Boutons. Neuron, 28:927–939.Gerstner W. and Kistler W.M. (2002) Spik<strong>in</strong>g Neuron Models: S<strong>in</strong>gle Neurons, Populations,Plasticity. Cambridge University Press.George R. M. (1914) On Circulat<strong>in</strong>g Excitations on Heart Muscles and Their Possible Relationto Tachycardia and Fibrillation. Trans. R. Soc. Can. 4:43–53.Gibson J.R., Belerle<strong>in</strong> M. and Connors B.W. (1999) Two networks of electrically coupled <strong>in</strong>hibitoryneurons <strong>in</strong> neocortex, Nature, 402:75–79Glass L. and MacKey M.C. (1988) From Clocks to Chaos. Pr<strong>in</strong>ceton University Press.Goel P. and Ermentrout B. (2002) Synchrony, stability, and fir<strong>in</strong>g patterns <strong>in</strong> pulse-coupledoscillators. Physica D, 163:191–216.Golubitsky M., Josic K. and Kaper T.J. (2001) An unfold<strong>in</strong>g theory approach to burst<strong>in</strong>g <strong>in</strong>fast-slow systems. In: Global Analysis of <strong>Dynamical</strong> <strong>Systems</strong>: Festschrift dedicated to FlorisTakens on the occasion of his 60th birthday. (H.W. Broer, B. Krauskopf and G. Vegter, eds.)Institute of Physics Publ., 277–308Golubitsky M. and Stewart I. (2002) Patterns of oscillation <strong>in</strong> coupled cell systems. In: Geometry,Dynamics, and Mechanics: 60th Birthday Volume for J.E. Marsden. (P. Holmes, P.Newton and A. We<strong>in</strong>ste<strong>in</strong>, eds.) Spr<strong>in</strong>ger-Verlag, 243–286.Gray C.M. and McCormick D.A. (1996) Chatter<strong>in</strong>g cells: superficial pyramidal neurons contribut<strong>in</strong>gto the generation of synchronous oscillations <strong>in</strong> the visual cortex. Science. 274(5284): 109–113Grundfest H. (1971) Biophysics and Physiology of Excitable Membranes. Ed. WJ Adelman,Van Nostrand Re<strong>in</strong>hold Co, New York,Guckenheimer J, Harris-Warrick R, Peck J, Willms A. (1997) Bifurcation, Burst<strong>in</strong>g, and SpikeFrequency Adaptation. J Comput Neurosci. 4:257–277.Guckenheimer J. and Holmes D. (1983) Nonl<strong>in</strong>ear Oscillations, <strong>Dynamical</strong> <strong>Systems</strong>, and Bifurcationsof Vector Fields. Spr<strong>in</strong>ger-Verlag, New York.The


446 ReferencesGuckenheimer J. (1975) Isochrons and Phaseless Sets. Journal of Mathematical Biology, 1:259–273.Guevara M.R. and Glass L. (1982) Phase lock<strong>in</strong>g, periodic doubl<strong>in</strong>g bifurcations and chaos <strong>in</strong>a mathematical model of a periodically driven oscillator: a theory for the entra<strong>in</strong>ment ofbiological oscillators and the generation of cardiac dysrhythmias. Math. Biol. 14: 1-23.Guttman R., Lewis S., R<strong>in</strong>zel J. (1980) Control of repetitive fir<strong>in</strong>g <strong>in</strong> squid axon membrane asa model for a neuroneoscillator. J. Physiology. 305: 377-395.Hansel D., Mato G., and Meunier C. (1995) Synchrony <strong>in</strong> excitatory neural networks. NeuralComputations 7:307–335.Hansel D. and Mato G. (2003) Asynchronous States and the Emergence of Synchrony <strong>in</strong> LargeNetworks of Interact<strong>in</strong>g Excitatory and Inhibitory Neurons. Neural Computation, 15:1–56.Hansel D., Mato G., Meunier C., and Neltner L. (1998) On Numerical Simulations of Integrateand-FireNeural Networks. Neural Computation, 10:467–483.Harris-Warrick R.M. and Flamm R.E. (1987) Multiple mechanisms of burst<strong>in</strong>g <strong>in</strong> a conditionalburst<strong>in</strong>g neuron. J. Neurosci., 7:2113–2128Hast<strong>in</strong>gs J.W. and Sweeney B.M. (1958) A persistent diurnal rhythms of lum<strong>in</strong>escence <strong>in</strong>Gonyaulax polyedra. Biol. Bull. 115:440-458.Hausser M., Spruston N., and Stuart G.J. (2000) Diversity and Dynamics of Dendritic Signal<strong>in</strong>g.Science, 290:739–744Hausser M. and Mel B. (2003) Dendrites: bug or feature? Current Op<strong>in</strong>ion <strong>in</strong> Neurobiology,13:372–383Heyward P., Ennis M., Keller A., and Shipley M.T. (2001) Membrane Bistability <strong>in</strong> OlfactoryBulb Mitral Cells The Journal of <strong>Neuroscience</strong>, 21:5311–5320Hille B. (2001) Ion Channels of Excitable Membranes. (2nd edition) S<strong>in</strong>auer, Sunderland, MA.H<strong>in</strong>dmarsh J.L. and Rose R.M. (1982) A model of the nerve impulse us<strong>in</strong>g two first-orderdifferential equations. Nature 296:162–164H<strong>in</strong>es M.A. (1989) Program for simulation of nerve equations with branch<strong>in</strong>g geometries. Int JBiomed Comput 24:55–68.Hodgk<strong>in</strong> A.L. (1948) The local electric changes associated with repetitive action <strong>in</strong> a nonmedulatedaxon. Journal of Physiology 107:165–181.Hodgk<strong>in</strong> A.L. and Huxley A.F. (1952) A quantitative description of membrane current andapplication to conduction and excitation <strong>in</strong> nerve. Journal Physiol., 117:500–544.Holden L. and Erneux T. (1993a) Slow passage through a Hopf bifurcation: form oscillatory tosteady state solutions. SIAM Journal on Applied Mathematics 53:1045–1058Holden L. and Erneux T. (1993b) Understand<strong>in</strong>g burst<strong>in</strong>g oscillations as periodic slow passagesthrough bifurcation and limit po<strong>in</strong>ts. J. Math. Biol. 31:351–365Hopf E. (1942) Abzweigung e<strong>in</strong>er periodischen Losung von e<strong>in</strong>er stationaren Losung e<strong>in</strong>es Differetialsystems.Ber. Math.-Phys. Kl. Sachs, Aca. Wiss. Leipzig, 94:1–22.Hoppensteadt F.C. (1997) An Introduction to the Mathematics of Neurons. Second edition:Model<strong>in</strong>g <strong>in</strong> the Frequency Doma<strong>in</strong>. Cambridge Univ. Press, Cambridge, U. K.


References 447Hoppensteadt F.C. (2000) Analysis and simulations of chaotic systems.Spr<strong>in</strong>ger-Verlag, New York.Hoppensteadt F.C. and Izhikevich E.M. (1996) Synaptic Organizations and <strong>Dynamical</strong> Propertiesof Weakly Connected Neural Oscillators. II. Learn<strong>in</strong>g of Phase Information. BiologicalCybernetics, 75:129–135Hoppensteadt F.C. and Izhikevich E.M. (1996) Synaptic Organizations and <strong>Dynamical</strong> Propertiesof Weakly Connected Neural Oscillators: I. Analysis of Canonical Model. BiologicalCybernetics, 75:117–127Hoppensteadt F.C. and Izhikevich E.M. (1997) Weakly Connected Neural Networks. Spr<strong>in</strong>ger-Verlag, NYHoppensteadt F.C. and Keener J.P. (1982) Phase Lock<strong>in</strong>g of Biological Clocks.Biology, 15:339–349.J. MAth.Hoppensteadt F.C. and Pesk<strong>in</strong> C.S.(2002) Model<strong>in</strong>g and Simulation <strong>in</strong> Medic<strong>in</strong>e and the LifeSciences. Second edition. Spr<strong>in</strong>ger-Verlag, NYHughes S.W., Cope D.W., Toth T.L., Williams S.R., and Crunelli V. (1999) All thalamocorticalneurones possess a T-type Ca2+ ’w<strong>in</strong>dow’ current that enables the expression of bistabilitymediatedactivities. The Journal of Physiology, 517:805–815.Huguenard J.R. and McCormick D.A. (1992) Simulation of the Currents Involved <strong>in</strong> RhythmicOscillations <strong>in</strong> Thalamic Relay Neurons. Journal of Neurophysiology, 68:1373–1383Hutcheon, B., Miura, R.M., Puil, E., (1996). Subthreshold membrane resonance <strong>in</strong> neocorticalneurons. Journal of Neurophysiology 76:683–697.Izhikevich E.M. (2003) Simple Model of Spik<strong>in</strong>g Neurons IEEE Transactions on Neural Networks,14:1569- 1572.Izhikevich E.M. (2002) Resonance and Selective Communication Via Bursts <strong>in</strong> Neurons Hav<strong>in</strong>gSubthreshold Oscillations. Bio<strong>Systems</strong>, 67:95-102Izhikevich E.M. (2001) Resonate-and-Fire Neurons. Neural Networks, 14:883-894Izhikevich E.M. (2001) Synchronization of Elliptic Bursters SIAM Review, 43:315-344Izhikevich E.M. (2000) Neural Excitability, Spik<strong>in</strong>g, and Burst<strong>in</strong>g.Bifurcation and Chaos. 10:1171-1266.International Journal ofIzhikevich E.M. (2000) Phase Equations For Relaxation Oscillators SIAM Journal on AppliedMathematics, 60:1789-1805Izhikevich E.M. (1999) Class 1 Neural Excitability, Conventional Synapses, Weakly ConnectedNetworks, and Mathematical Foundations of Pulse-Coupled Models. IEEE Transactions OnNeural Networks,10:499-507Izhikevich E.M. (1999) Weakly Connected Quasiperiodic Oscillators, FM Interactions, and Multiplex<strong>in</strong>g<strong>in</strong> the Bra<strong>in</strong>. SIAM Journal on Applied Mathematics, 59:2193-2223Izhikevich E.M. (1998) Phase Models With Explicit Time Delays. Physical Review E, 58:905-908Izhikevich E.M., Desai N.S., Walcott E.C., Hoppensteadt F.C. (2003) Bursts as a unit of neural<strong>in</strong>formation: selective communication via resonance . Trends <strong>in</strong> <strong>Neuroscience</strong> 26:161-167Izhikevich E.M. and Hoppensteadt F.C. (2004) Classification of Burst<strong>in</strong>g Mapp<strong>in</strong>gs. Interna-


448 Referencestional Journal of Bifurcation and Chaos, 14:3847–3854Jahnsen H. and Ll<strong>in</strong>as R. (1984) Electrophysiological properties of gu<strong>in</strong>ea-pig thalamic neurons:an <strong>in</strong> vitro study. Journal of Physiology London, 349:205–226.Jensen M.S., Azouz R., and Yaari Y. (1994) Variant Fir<strong>in</strong>g Patterns <strong>in</strong> RAt HippocampalPyramidal Cells Modulated by Extracellular Potassium. Journal of Neurophysiology, 71:831–839Jian Z., X<strong>in</strong>g J.L., Yang G.S., and Hu S.J. (2004) A novel Burst<strong>in</strong>g Mechanism of Type ANeurons <strong>in</strong> Injured Dorsal Root Ganglia. NeuroSignals, 13:150-156.Johnson C.H. (1999) Forty Years of PRC — What Have We Learned?. Chronobiology International,16:711-743.Johnston D. and Wu S.M. (1995) Foundations of Cellular Neurophysiology. The MIT Press.Katriel G. (2005) Stability of Synchronized Oscillations <strong>in</strong> Networks of Phase-Oscillators. Descreteand Cont<strong>in</strong>uous <strong>Dynamical</strong> <strong>Systems</strong>-Series B. 5:353–364.Kay A.R., Sugimori M., and Ll<strong>in</strong>as R. (1998) K<strong>in</strong>etic and Stochastic Properties of a PersistentSodium Current <strong>in</strong> Mature Gu<strong>in</strong>ea Pig Cerebellar Purk<strong>in</strong>je Cells, J. Neurophysiology,80:1167–1179.Keener J. and Sneyd J. (1998) Mathematical Physiology. Spr<strong>in</strong>ger-Verlag, New YorkKepler T.B., Abbott L.F., and Marder E. (1992) Reduction of conductance based neuron models.Biological Cybernetics, 66:381–387.K<strong>in</strong>ard T.A., de Vries G., Sherman A., Sat<strong>in</strong> L.S. (1999) Modulation of the burst<strong>in</strong>g propertiesof s<strong>in</strong>gle mouse pancreatic beta-cells by artificial conductances. Biophys J. 76:1423-35Kl<strong>in</strong>k R. and Alonso A. (1993) Ionic Mechanisms for the Subthreshold Oscillations and DifferentialElectroresponsiveness of Medial Entorh<strong>in</strong>al Cortex Layer II Neurons. Journal ofNeurophysiology, 70:144–157.Koch C. (1999) Biophysics of Computation: Information Process<strong>in</strong>g <strong>in</strong> S<strong>in</strong>gle Neurons. OxfordUniversity Press.Kopell N. (1986) Coupled oscillators and locomotion by fish. In Othmer HG (Ed) Nonl<strong>in</strong>earOscillations <strong>in</strong> Biology and Chemistry. Lecture Notes <strong>in</strong> Biomathematics, Spr<strong>in</strong>ger-Verlag.Kopell N. (1995) Cha<strong>in</strong>s of Coupled Oscillators. In Arbib M.A. (Ed) Bra<strong>in</strong> Theory and NeuralNetworks, The MIT press, Cambridge, MA.Kopell N. and Ermentrout G.B. (1990) Phase transitions and other phenomena <strong>in</strong> cha<strong>in</strong>s ofcoupled oscillators. SIAM Journal on Applied Mathematics 50:1014–1052.Kopell N., Ermentrout G.B., Williams T.L. (1991) On Cha<strong>in</strong>s of Oscillators Forced at One End.SIAM Journal on Applied Mathematics 51:1397–1417.Kopell N. and Somers D. (1995) Anti-phase solutions <strong>in</strong> relaxation oscillators coupled throughexcitatory <strong>in</strong>teractions. J. Math. Biol. 33:261–280.Korngreen A. and Sakmann B. (2000) Voltage-gated K + channels <strong>in</strong> layer 5 neocortical pyramidalneurones from young rats: subtypes and gradients. Journal of Physiology, 525.3:621–639Kr<strong>in</strong>skii V.I. and Kokoz Yu.M. (1973) Analysis of Equations of Excitable Membranes - I. Reductionof the Hodgk<strong>in</strong>-Huxley Equations to a Second Order System. Biofizika, 18 3, 506–511


References 449Kuramoto, Y., (1975) <strong>in</strong> H. Araki (Ed) International Symposium on Mathematical Problems <strong>in</strong>Theoretical Physics, Lecture Notes <strong>in</strong> Physics, Vol. 39, Spr<strong>in</strong>ger, New York, p. 420-422.Kuramoto Y. (1984) Chemical Oscillations, Waves, and Turbulence. Spr<strong>in</strong>ger-Verlag, New York.Kuznetsov Yu. (1995) Elements of Applied Bifurcation Theory. Spr<strong>in</strong>ger-Verlag, New York.Lapicque, L (1907) Recherches quantitatives sur l’excitation electrique des nerfs traitee commeune polarization. J Physiol Pathol Gen 9:620-635.Latham PE, Richmond BJ, Nelson PG, Nirenberg S. Intr<strong>in</strong>sic dynamics <strong>in</strong> neuronal networks.I. Theory. J Neurophysiol 2000 Feb;83(2):808-27Levitan E.S., Kramer R.H., and Levitan I.B. (1987) Augmentation of burst<strong>in</strong>g pacemaker activityby egg-lay<strong>in</strong>g hormone <strong>in</strong> Aplysia neuron R15 is mediated by a cyclic AMP-dependent<strong>in</strong>crease <strong>in</strong> Ca2+ and K+ currents. Proc Natl Acad Sci U S A. 84 (17): 6307–6311.Levitan E.S. and Livitan I.B. (1988) A Cyclic GMP Analog Decreases the Currents Underly<strong>in</strong>gBurst<strong>in</strong>g Activity <strong>in</strong> the Aplysia Neuron R15. J. <strong>Neuroscience</strong>, 8:1162–1171Li, J., Bickford M.E., and Guido W. (2003) Dist<strong>in</strong>ct fir<strong>in</strong>g properties of higher order thalamicrelay neurons. Journal of Neurophysiology, 90: 291–299.Liénard A. (1928) Etude des oscillations entretenues, Rev. Gen. Elec. 23:901–912, 946–954Lisman J. (1997) Bursts as a unit of neural <strong>in</strong>formation: mak<strong>in</strong>g unreliable synapses reliable,Trends Neurosci. 20:38–43Lopat<strong>in</strong> A.N., Makh<strong>in</strong>a E.N., Nichols C.G. (1994) Potassium channel block by cytoplasmicpolyam<strong>in</strong>es as the mechanism of <strong>in</strong>tr<strong>in</strong>sic rectification. Nature 373:366–369Luk W.K. and Aihara K. (2000) Synchronization and sensitivity enhancement of the Hodgk<strong>in</strong>-Huxley neurons due to <strong>in</strong>hibitory <strong>in</strong>puts. Biological Cybernetics, 82:455–467.Magee J.C. (1998) Dendritic Hyperpolarization-Activated Currents Modify the Integrative Propertiesof Hippocampal CA1 Pyramidal Neurons. The Journal of <strong>Neuroscience</strong>, 18:7613–7624.Magee J.C. and Carruth M. (1999) Dendritic Voltage-Gated IOn Channels Regulate the ActionPotential Fir<strong>in</strong>g Mode of Hippocampal CA1 Pyramidal neurons. J. Neurophysiology,82:1895–1901.Magistretti J. and Alonso A. (1999) Biophysical Properties and Slow Voltage- dependent Inactivationof a Susta<strong>in</strong>ed Sodium Current <strong>in</strong> Entorh<strong>in</strong>al Cortex Layer-II Pr<strong>in</strong>cipal Neurons.J.Gen.Physiol., 114:491-509.Ma<strong>in</strong>en Z.F., Sejnowski T.J. (1995) Reliability of spike tim<strong>in</strong>g <strong>in</strong> neocortical neurons. Science,268(5216):1503–1506.Ma<strong>in</strong>en Z.F., Sejnowski T.J. (1996) Influence of dendritic structure on fir<strong>in</strong>g pattern <strong>in</strong> modelneocortical neurons. Nature. 382:363-366.Malk<strong>in</strong> I.G. (1949) Methods of Po<strong>in</strong>care and Liapunov <strong>in</strong> theory of non-l<strong>in</strong>ear oscillations. [<strong>in</strong>Russian: “Metodi Puankare i Liapunova v teorii nel<strong>in</strong>e<strong>in</strong>ix kolebanii” Gostexizdat, Moscow].Malk<strong>in</strong> I.G. (1956) Some Problems <strong>in</strong> Nonl<strong>in</strong>ear Oscillation Theory. [<strong>in</strong> Russian: “Nekotoryezadachi teorii nel<strong>in</strong>e<strong>in</strong>ix kolebanii” Gostexizdat, Moscow].Marder E. and Bucher D. (2001) Central pattern generators and the control of rhythmic movements.Current Biology, 11:986–996.


450 ReferencesMarkram H, Toledo-Rodriguez M, Wang Y, Gupta A, Silberberg G, and Wu C. (2004) Interneuronsof the neocortical <strong>in</strong>hibitory system. Nature Review <strong>Neuroscience</strong> 5:793–807Medvedev G. (2005) Reduction of a model of an excitable cell to a one-dimensional map. PhysicaD, 202:37–59.Melnikov V.K. (1963) On the stability of the center for time periodic perturbations. Trans.Moscow Math. Soc. 12:1-57.McCormick D.A. (2004) Membrane Properties and Neurotransmitter Actions, <strong>in</strong> Shepherd G.M.The Synaptic Organization of the Bra<strong>in</strong>. Oxford University Press, New York, Fifth Edition.McCormick D.A. and Pape H.-C. (1990) Properties of a Hyperpolarization-Activated CationCurrent and Its Role <strong>in</strong> Rhythmic Oscillation <strong>in</strong> Thalamic Relay Neurones. Journal of Physiology,431:291–318Mirollo R.E. and Strogatz S.H. (1990) Synchronization of pulse-coupled biological oscillators.SIAM J. Appl. Math. 50:1645–1662Mishchenko E.F., Kolesov Yu.S., Kolesov A.Yu., and Rozov N.K. (1994) Asymptotic Methods<strong>in</strong> S<strong>in</strong>gularly Perturbed <strong>Systems</strong>. Plenum Press, New York.Morris C. and Lecar H. (1981) Voltage Oscillations <strong>in</strong> the Barnacle Giant Muscle Fiber, BiophysicalJournal 35:193–213.Murray J.D. (1993). Mathematical Biology, Spr<strong>in</strong>ger-Verlag.Nagumo, J., S. Arimoto, and S. Yoshizawa (1962) An active pulse transmission l<strong>in</strong>e simulat<strong>in</strong>gnerve axon. Proc IRE. 50:2061–2070.Nejshtadt A. (1985) Asymptotic <strong>in</strong>vestigation of the loss of stability by an equilibrium as a pairof eigenvalues slowly cross the imag<strong>in</strong>ary axis. Usp. Mat. Nauk 40:190-191.Neu J.C. (1979) Coupled chemical oscillators. SIAM Journal of Applied Mathematics 37:307–315.Nisenbaum E. S., Xu Z.C., and Wilson C.J. (1994) Contribution of a slowly <strong>in</strong>activat<strong>in</strong>g potassiumcurrent to the transition to fir<strong>in</strong>g of neostriatal sp<strong>in</strong>y projection neurons. Journal ofNeurophysiology, 71:1174–1189Noble D. (1966) Applications of Hodgk<strong>in</strong>-Huxley equations to excitable tissues. Physiol Rev1966 46:1–50Nowak L.G., Azouz R., Sanchez-Vives M.V., Gray C.M., and McCormick D.A. (2003) ElectrophysiologicalClasses of Cat Primary Visual Cortical Neurons In Vivo as Revealed byQuantitative Analyses. J Neurophysiol 89: 1541-1566Parri H. R. and Crunelli V. (1999) Sodium Current <strong>in</strong> Rat and Cat Thalamocortical Neurons:Role of a Non-Inactivat<strong>in</strong>g Component <strong>in</strong> Tonic and Burst Fir<strong>in</strong>g, The Journal of <strong>Neuroscience</strong>,February 1, 1998, 18(3):854-867Pape, H.-C., and McCormick, D.A. (1995) Electrophysiological and pharmacological propertiesof <strong>in</strong>terneurons <strong>in</strong> the cat dorsal lateral geniculate nucleus. <strong>Neuroscience</strong>, 68: 1105–1125.Pernarowski M., Miura R.M., and Kevorkian J. (1992) Perturbation techniques for models ofburst<strong>in</strong>g electrical activity <strong>in</strong> pancreatic β-cells. SIAM Journal on Applied Mathematics52:1627–1650.


References 451Pernarowski M. (1994) Fast Subsystem Bifurcations <strong>in</strong> a Slowly Varied Liénard System Exhibit<strong>in</strong>gBurst<strong>in</strong>g. SIAM Journal on Applied Mathematics, 54:814–832.Pfeuty B., Mato G., Golomb D., Hansel D. (2003) Electrical Synapses and Synchrony: The Roleof Intr<strong>in</strong>sic Currents. J. <strong>Neuroscience</strong>, 23:6280–6294.P<strong>in</strong>sky P. and R<strong>in</strong>zel J. (1994) Intr<strong>in</strong>sic and network rhythmogenesis <strong>in</strong> a reduced Traub modelof CA3 neurons. J. Comput. Neurosci. 1:39–60.Pikovsky A., Rosenblum M., Kurths J. (2001) Synchronization: A Universal Concept <strong>in</strong> Nonl<strong>in</strong>earScience. CUP, Cambridge.Pirchio M., Turner J. P., Williams S. R., Asprod<strong>in</strong>i E. and Crunelli V. (1997) Postnatal Developmentof Membrane Properties and delta Oscillations <strong>in</strong> Thalamocortical Neurons of theCat Dorsal Lateral Geniculate Nucleus. J. Neurosci. 17 :5428–5444.Plant R.E. (1981) Bifurcation and Resonance <strong>in</strong> a Model for Burst<strong>in</strong>g Nerve Cells. Journal ofMathematical Biology, 11:15–32.Pontryag<strong>in</strong> L.S. and Rodyg<strong>in</strong> L.V. (1960) Periodic solution of a system of ord<strong>in</strong>ary differentialequations with a small parameter <strong>in</strong> the terms conta<strong>in</strong><strong>in</strong>g derivatives Sov. Math. Dokl. 1:611–614.Rall W. (1959) Branch<strong>in</strong>g dendritic trees and motoneuron membrane resistivity. Exp. Neurol.1:491–527.Reuben J. P., Werman R., and Grundfest H. (1961) The Ionic Mechanisms of Hyperpolariz<strong>in</strong>gResponses <strong>in</strong> Lobster Muscle Fibers. The Journal of General Physiology, 45:243–265Reyes A.D. and Fetz E.E. (1993) Two modes of <strong>in</strong>terspike <strong>in</strong>terval shorten<strong>in</strong>g by brief transientdepolarizations <strong>in</strong> cat neocortical neurons. Journal of Neurophysiology 69: 1661–1672.Richardson, M.J.E., Brunel, N. and Hakim, V. (2003) From Subthreshold to Fir<strong>in</strong>g-Rate Resonance.Journal of Neurophysiology 89:2538–2554R<strong>in</strong>zel J. and Ermentrout G.B. (1989) Analysis of Neural Excitability and Oscillations. In KochC., Segev I. (eds) Methods <strong>in</strong> Neuronal Model<strong>in</strong>g, The MIT Press, Cambridge, Mass.R<strong>in</strong>zel J. (1978) On repetitive activity <strong>in</strong> nerve. Fed. Proc. 37:2793-2802.R<strong>in</strong>zel J. (1985) Burst<strong>in</strong>g oscillations <strong>in</strong> an excitable membrane model. In: Sleeman BD, JarvisRJ, eds. Ord<strong>in</strong>ary and partial Differential Equations Proceed<strong>in</strong>gs of the 8th Dundee ConferenceLecture Notes <strong>in</strong> Mathematics 1151. Berl<strong>in</strong>: Spr<strong>in</strong>ger, 304–316.R<strong>in</strong>zel J. (1987) A formal classification of burst<strong>in</strong>g mechanisms <strong>in</strong> excitable systems. In: E.Teramoto, M. Yamaguti, eds. Mathematical Topics <strong>in</strong> Population Biology, Morphogenesis,and <strong>Neuroscience</strong>s, vol. 71 of Lecture Notes <strong>in</strong> Biomathematics, Spr<strong>in</strong>ger-Verlag, Berl<strong>in</strong>.R<strong>in</strong>zel J. and Lee Y.S. (1986) On different mechanisms for membrane potential burst<strong>in</strong>g. In OthmerH.G. (Ed) Nonl<strong>in</strong>ear Oscillations <strong>in</strong> Biology and Chemistry. Lecture Notes <strong>in</strong> Biomathematics,Spr<strong>in</strong>ger-Verlag.R<strong>in</strong>zel J. and Lee Y.S. (1987) Dissection of a model for neuronal parabolic burst<strong>in</strong>g. Journal ofMathematical Biology, 25:653–675.Robb<strong>in</strong>s J., Trouslard J., Marsh S.J., and Brown D.A. (1992) K<strong>in</strong>etic and PharmacologicalProperties of the M-current <strong>in</strong> Rodent Neuroblastoma × Glioma Hybrid Cells, Journal ofPhysiology, 451:159-185.


452 ReferencesRosenblum M.G. and Pikovsky A.S. (2001) Detect<strong>in</strong>g direction of coupl<strong>in</strong>g <strong>in</strong> <strong>in</strong>teract<strong>in</strong>g oscillators.Physical Review E, 64, p. 045202.Roy J. P., Clercq M. , Steriade M. , and Deschenes M. (1984) Electrophysiology of neurons oflateral thalamic nuclei <strong>in</strong> cat: mechanisms of long-last<strong>in</strong>g hyperpolarizations. J Neurophysiol,51:1220–1235.Rub<strong>in</strong> J. and Terman D. (2000) Analysis of clustered fir<strong>in</strong>g patterns <strong>in</strong> synaptically couplednetworks of oscillators. J. Math. Biology, 41:513–545.Rub<strong>in</strong> J. and Terman D. (2002) Geometric s<strong>in</strong>gular perturbation analysis of neuronal dynamics.Handbook of <strong>Dynamical</strong> systems, vol. 2: Toward Applications (B. Fiedler, G. Iooss, eds.)Elsevier, AmsterdamRush M. E. and R<strong>in</strong>zel J. (1995) The Potassium A-Current, Low Fir<strong>in</strong>g Rates and ReboundExcitation <strong>in</strong> Hodgk<strong>in</strong>-Huxley Models. Bullet<strong>in</strong> of Mathematical Biology, 57:899–929.Rush M. E. and R<strong>in</strong>zel J. (1994) Analysis of burst<strong>in</strong>g <strong>in</strong> a thalamic neuron model, BiologicalCybernetics 71:281–291.Samborskij S.N. (1985) Limit trajectories of s<strong>in</strong>gularly perturbed differential equations. Dokl.Akad. Nauk Ukr. SSR. Ser. A, 9:22–25.Sanabria E.R.G., Su H., and Yaari Y. (2001) Initiation of network bursts by Ca 2+ -dependent <strong>in</strong>tr<strong>in</strong>sicburst<strong>in</strong>g <strong>in</strong> the rat pilocarp<strong>in</strong>e model of temporal lobe epilepsy. Journal of Physiology,532:205–216Sharp A.A., O’Neil M.B., Abbott L.F., Marder E. (1993) Dynamic clamp: computer-generatedconductances <strong>in</strong> real neurons. J Neurophysiol, 69:992-995Shilnikov, A. L., Calabrese R. and Cymbalyuk, G. (2005) Mechanism of bi-stability: tonicspik<strong>in</strong>g and burst<strong>in</strong>g <strong>in</strong> a neuron model. Phys Review E, 71, 056214.Shilnikov A.L. and Cymbalyuk G (2005) Transition between Tonic Spik<strong>in</strong>g and Burst<strong>in</strong>g <strong>in</strong> aNeuron Model via the Blue-Sky Catastrophe. Physical Review Letters, 94, 048101.Shilnikov A.L. and Cymbalyuk G (2004) Homocl<strong>in</strong>ic bifurcations of periodic orbits en a routefrom tonic spik<strong>in</strong>g to burst<strong>in</strong>g <strong>in</strong> neuron models. Regular and Chaotic Dynamics, vol. 9,[2004-11-19]Shilnikov L.P., Shilnikov A.L., Turaev D., Chua L.O. (2001) Methods of qualitative theory <strong>in</strong>nonl<strong>in</strong>ear dynamics. Part II, World Scientific, S<strong>in</strong>gaporeShilnikov L.P., Shilnikov A.L., Turaev D., Chua L.O. (1998) Methods of qualitative theory <strong>in</strong>nonl<strong>in</strong>ear dynamics. Part I, World Scientific, S<strong>in</strong>gaporeShilnikov, A.L. and Shilnikov, L.P, (1995) Dangerous and safe stability boundaries of equilibriaand periodic orbits” <strong>in</strong> NDES’95, University College Dubl<strong>in</strong>, Ireland, 55-63.Shishkova M.A. (1973) Investigation of a system of differential equations with a small parameter<strong>in</strong> highest derivatives. Dokl. Adad. Nauk SSSR 209. No.3, 576–579. English transl. Sov.Math., Dokl. 14,483–487Shepherd G.M. (2004) The Synaptic Organization of the Bra<strong>in</strong>. Oxford University Press, NewYork, Fifth Edition.Sk<strong>in</strong>ner F.K., Kopell N., Marder E. (1994) Mechanisms for oscillation and frequency control <strong>in</strong>reciprocal <strong>in</strong>hibitory model neural networks. J. Comput. Neurosci., 1:69–87.


References 453Smolen P., Terman D., and R<strong>in</strong>zel J. (1993) Properties of a burst<strong>in</strong>g model with two slow<strong>in</strong>hibitory variables. SIAM Journal on Applied Mathematics 53:861–892.Somers D. and Kopell N. (1993) Rapid synchronization through fast threshold modulation.Biological Cybernetics, 68:393–407Somers D. and Kopell N. (1995) Waves and synchrony <strong>in</strong> networks of oscillators or relaxationand non-relaxation type. Physica D, 89:169–183Stanford, I. M., Traub R. D., and Jefferys J. G.R. (1998) Limbic gamma rhythms. II. Synapticand <strong>in</strong>tr<strong>in</strong>sic mechanisms underly<strong>in</strong>g spike doublets <strong>in</strong> oscillat<strong>in</strong>g subicular neurons. J.Neurophysiol. 80:162–171Ste<strong>in</strong> RB. Some models of neuronal variability. Biophys J 7: 37-68, 1967Steriade M. (2003) Neuronal Substrates of Sleep and Epilepsy. Cambridge University Press.Steriade M. (2004) Neocortical Cell Classes Are Flexible Entities. Nature Reviews <strong>Neuroscience</strong>,5:121–134.Strogatz S.H. (1994) Nonl<strong>in</strong>ear Dynamics and Chaos. Cambridge, Mass.Strogatz S. H. (2000) From Kuramoto to Crawford: explor<strong>in</strong>g the onset of synchronization <strong>in</strong>populations of coupled oscillators. Physica D, 143: 1–20.Stuart G., Spruston N., Hausser M. (1999) Dendrites. Oxford University Press.Su H, Alroy G, Kirson ED, Yaari Y. (2001) Extracellular calcium modulates persistent sodiumcurrent-dependent burst-fir<strong>in</strong>g <strong>in</strong> hippocampal pyramidal neurons. Journal of <strong>Neuroscience</strong>21:4173-4182.Szmolyan P. and Wechselberger M. (2001) Canards <strong>in</strong> R 3 . Journal of Differential Equations,177:419–453.Szmolyan P. and Wechselberger M. (2004) Relaxation oscillations <strong>in</strong> R 3 . Journal of DifferentialEquations, 200:69–104.Tateno T., Harsch A., and Rob<strong>in</strong>son H.P.C. (2004) Threshold fir<strong>in</strong>g frequency-current relationshipsof neurons <strong>in</strong> rat somatosensor cortex: type 1 and type 2 dynamics. Journal ofNeurophysiology 92: 2283–2294.Tennigkeit F., Ries C.R., Schwarz D.W.F., and Puil E. (1997) Isoflurane attenuates resonantresponses of auditory thalamic neurons. J. Neurophysiol. 78:591–596Terman D. (1991) Chaotic spikes aris<strong>in</strong>g from a model of burst<strong>in</strong>g <strong>in</strong> excitable membranes.SIAM J. Appl. Math. 51:1418–1450.Timofeev I., Grenier F., Bazhenov M., Sejnowski T.J. and Steriade M. (2000) Orig<strong>in</strong> of SlowCortical Oscillations <strong>in</strong> Deafferented Cortical Slabs Cerebral Cortex, Vol. 10, No. 12, 1185-1199Toledo-Rodriguez M., Blumenfeld B., Wu C., Luo J, Attali B., Goodman P., and Markram H.(2004) Correlation Maps Allow Neuronal Electrical Properties to be Predicted from S<strong>in</strong>gle-cellGene Expression Profiles <strong>in</strong> Rat Neocortex. Cerebral Cortex, 14:1310–1327.Traub R. D., Wong R. K., Miles R., and Michelson H. (1991) A model of a CA3 hippocampalpyramidal neuron <strong>in</strong>corporat<strong>in</strong>g voltage-clamp data on <strong>in</strong>tr<strong>in</strong>sic conductances. Journal ofNeurophysiology, 66:635–650


454 ReferencesTuckwell H.C. (1988) Introduction to Theoretical Neurobiology. Cambridge, UK, CambridgeUniversity Press.Uhlenbeck G.E. and Ornste<strong>in</strong> L.S. (1930) On the theory of the Brownian motion. Phys. Rev.36:823–841.van Vreeswijk C., Abbott L.F., Ermentrout G.B. (1994) When <strong>in</strong>hibition not excitation synchronizesneural fir<strong>in</strong>g. Journal of Computational <strong>Neuroscience</strong>, 1:313–321van Vreeswijk C. and Hansel D. (2001) Patterns of Synchrony <strong>in</strong> Neural Networks with SpikeAdaptation. Neural Comp. 13:959–992van Vreeswijk C. (2000) Analysis of the Asynchronous State <strong>in</strong> Networks of Strongly CoupledOscillators. Physical Review Letters, 84, 5110–5113.de Vries G. (1998) Multiple Bifurcations <strong>in</strong> a Polynomial Model of Burst<strong>in</strong>g Oscillations. Journalof Nonl<strong>in</strong>ear Science, 8:281–316.Wang X.-J. (1999) Fast burst fir<strong>in</strong>g and short-term synaptic plasticity: a model of neocorticalchatter<strong>in</strong>g neurons. <strong>Neuroscience</strong> 89:347-362.Wang X.-J. and R<strong>in</strong>zel J. (1992) Alternat<strong>in</strong>g and synchronous rhythms <strong>in</strong> reciprocally <strong>in</strong>hibitorymodel neurons. Neural Computation. 4:84–97.Wechselberger M. (2005) Existence and Bifurcation of Canards <strong>in</strong> R 3 <strong>in</strong> the Case of a FoldedNode. SIAM J. Appl. <strong>Dynamical</strong> <strong>Systems</strong>, 4:101–139.Wessel R., Kristan W.B., and Kle<strong>in</strong>feld D. (1999) Supral<strong>in</strong>ear Summation of Synaptic Inputs byan Invertebrate Neuron: Dendritic Ga<strong>in</strong> Is Mediated by an “Inward Rectifier” K + Current.The Journal of <strong>Neuroscience</strong>, 19:5875–5888.White J.A., Rub<strong>in</strong>ste<strong>in</strong> J. T., and Kay A. R. (2000) Channel noise <strong>in</strong> neurons.<strong>Neuroscience</strong>, 23:131–137.Trends <strong>in</strong>Williams J.T., North R.A., and Tokimasa T. (1988) Inward Rectification of Rest<strong>in</strong>g and Opiate-Activated Potassium Currents <strong>in</strong> Rat Locus Coeruleus Neurons. The Journal of <strong>Neuroscience</strong>,8:4299–4306Williams S.R. and Stuart G.J. (2003) Role of dendritic synapse location <strong>in</strong> the control of actionpotential. TINS, 26:147–154.Willms A.R., Baro D.J., Harris-Warrick R.M., and Guckenheimer J. (1999) An Improved ParameterEstimation Method for Hodgk<strong>in</strong>-Huxley Models, Journal of Computational <strong>Neuroscience</strong>,6:145–168.Wilson C.J. (1993) The generation of natural fir<strong>in</strong>g patterns <strong>in</strong> neostriatal neurons. In Progress<strong>in</strong> Bra<strong>in</strong> Research. (Arbuthnott G., Emson P.C. eds),277–297. Amsterdam. ElsevierWilson C. J. and Groves P. M. (1981) Spontaneous fir<strong>in</strong>g patterns of identified sp<strong>in</strong>y neurons<strong>in</strong> the rat neostriatum. Bra<strong>in</strong> Research, 220:67–80Wilson H.R. and Cowan J.D. (1972) Excitatory and <strong>in</strong>hibitory <strong>in</strong>teraction <strong>in</strong> localized populationsof model neurons. Biophys J 12:1–24.Wilson H.R. and Cowan J.D. (1973) A Mathematical theory of the functional dynamics ofcortical and thalamic nervous tissue. Kybernetik 13:55–80.Wilson H.R. (1999) Spikes, Decisions, and Actions: The dynamical foundations of neuroscience.


References 455Oxford University Press.W<strong>in</strong>free A. (1967) Biological Rhythms and the Behavior of Populations of Coupled Oscillators.Journal of Theoretical Biology, 16:15–42.W<strong>in</strong>free A. (1974) Patterns of Phase Compromise <strong>in</strong> Biological Cycles. Journal of Math. Biology1:73–95.W<strong>in</strong>free A. (2001) The Geometry of Biological Time. Spr<strong>in</strong>ger-Verlag, New York, second edition.Wolfram S. (2002) A New K<strong>in</strong>d of Science. Wolfram Media.Wu N., Hsiao C.-F., Chandler S. (2001) Membrane resonance and subthreshold membrane oscillations<strong>in</strong> mesencephalic V neurons: participants <strong>in</strong> burst generation. J. Neurosci 21:3729–3739.Young G. (1937) Psychometrika, 2:103.Yuan A., Dourado M., Butler A., Walton N., Wei A., Salkoff L. (2000) SLO-2, a K + channelwith an unusual Cl − dependence. Nature Neurscience, 3:771–779.Yuan A., Santi C.M., Wei A., Wang Z.W., Pollak K., Nonet M., Kaczmarek L., Crowder C.M.,Salkoff L. (2003) The sodium-activated potassium channel is encoded by a member of the Slogene family. Neuron. 37:765–773.Yue C. and Yaari Y. (2004) KCNQ/M Channels Control Spike Afterdepolarization and BurstGeneration <strong>in</strong> Hippocampal Neurons. J. <strong>Neuroscience</strong>, 24:4614–4624.Zhan, X. J., Cox C. L., R<strong>in</strong>zel J., and Sherman S. M. (1999) Current clamp and model<strong>in</strong>gstudies of low-threshold calcium spikes <strong>in</strong> cells of the cat’s lateral geniculate nucleus. J.Neurophysiol. 81:2360–2373.


456 References9


Chapter 10Synchronization (seewww.izhikevich.com)In this chapter we consider networks of tonically spik<strong>in</strong>g neurons. As any other k<strong>in</strong>d ofphysical, chemical, or biological oscillators, such neurons could synchronize and exhibitcollective behavior that is not <strong>in</strong>tr<strong>in</strong>sic to any <strong>in</strong>dividual neuron. For example, partialsynchrony <strong>in</strong> cortical networks is believed to generate various bra<strong>in</strong> oscillations, such asthe alpha and gamma EEG rhythms. Increased synchrony may result <strong>in</strong> pathologicaltypes of activity, such as epilepsy. Coord<strong>in</strong>ated synchrony is needed for locomotion andswim pattern generation <strong>in</strong> fish. There is an ongo<strong>in</strong>g debate on the role of synchrony<strong>in</strong> neural computation, see e.g., the special issue of Neuron (September 1999) devotedto the b<strong>in</strong>d<strong>in</strong>g problem.Depend<strong>in</strong>g on the circumstances, synchrony could be good or bad, and it is importantto know what factors contribute to synchrony and how to control it. This is thesubject of the present chapter – the most advanced chapter of the book. It provides anice application of the theory developed earlier and hopefully gives some <strong>in</strong>sight <strong>in</strong>towhy the previous chapters might be worth master<strong>in</strong>g.Our goal is to understand how the behavior of two coupled neurons depends on their<strong>in</strong>tr<strong>in</strong>sic dynamics. First, we <strong>in</strong>troduce the method of description of an oscillation byits phase. Then, we describe various methods of reduction of coupled oscillators tosimple phase models. The reduction method and the exact form of the phase modeldepends on the type of coupl<strong>in</strong>g, i.e., whether it is pulsed, weak, or slow, and on thetype of bifurcations of the limit cycle attractor generat<strong>in</strong>g tonic spik<strong>in</strong>g. F<strong>in</strong>ally, weshow how to use phase models to understand the collective dynamics of many coupledoscillators.10.1 Pulsed Coupl<strong>in</strong>gIn this section we consider oscillators of the formẋ = f(x) + Aδ(t − t s ) , x ∈ R m , (10.1)457


458 Synchronization (see www.izhikevich.com)membrane potential, V (mV)x( )x( ) x( )(c)phase of oscillation(a)200x(t)-20-40-60-800 2 4 6 T2T3Ttime (ms)T00 T 2T 3TK + activation gate, n(b)0.60.40.20V-nullcl<strong>in</strong>e-80 -60 -40 -20 0 20membrane potential, V (mV)(d)T/43T/85T/8T/23T/4 7T/8T/8n-nullcl<strong>in</strong>eT/43T/8T/80,T x 0limit cycle attractorT/20,T5T/8 7T/83T/4Figure 10.1: Def<strong>in</strong>ition of a phase of oscillation, ϑ, <strong>in</strong> the I Na + I K -model with parametersas <strong>in</strong> Fig. 4.1a and I = 10.hav<strong>in</strong>g exponentially stable limit cycles and experienc<strong>in</strong>g pulsed stimulation at timest s that <strong>in</strong>stantaneously <strong>in</strong>creases the state variable by constant A. The Dirac deltafunction δ(t) is a mathematical shorthand notation for resett<strong>in</strong>g x by A. The strengthof pulsed stimulation, A, is not assumed to be small. Most of the results of this sectioncan also be applied to the case when the action of <strong>in</strong>put pulse is not <strong>in</strong>stantaneous,but smeared over an <strong>in</strong>terval of time, typically shorter than the period of oscillation.10.1.1 Phase of oscillationMany types of physical, chemical, and biological oscillators share an astonish<strong>in</strong>g feature:They can be described by a s<strong>in</strong>gle phase variable ϑ. In the context of tonic spik<strong>in</strong>g,the phase is just the time s<strong>in</strong>ce the last spike, as <strong>in</strong> Fig. 10.1a.In general, the notion of the phase is related to the notion of parametrization ofa limit cycle attractor, as <strong>in</strong> Fig. 10.1b. Take a po<strong>in</strong>t x 0 on the attractor and plotthe trajectory x(t) with x(0) = x 0 . Then, the phase of x(t) is just ϑ = t. As t<strong>in</strong>creases past the period T , then 2T , etc., the phase variable ϑ wraps around the<strong>in</strong>terval [0, T ], jump<strong>in</strong>g from T to 0; see Fig. 10.1c. Glu<strong>in</strong>g together the po<strong>in</strong>ts 0 andT , as <strong>in</strong> Fig. 10.1d, we can treat the <strong>in</strong>terval [0, T ] as a circle, denoted as S 1 , withcircumference T . The parametrization is the mapp<strong>in</strong>g of S 1 <strong>in</strong> Fig. 10.1d <strong>in</strong>to the


Synchronization (see www.izhikevich.com) 4590.60.5V-nullcl<strong>in</strong>eK + activation gate, n K + activation gate, n0.40.30.20.100.60.50.40.30.20.1n-nullcl<strong>in</strong>ey 0y(t 1 )x(t 1 )x 0isochrony(t 3 )y(t 2 )x(t 3 )x(t 2 )0-80 -70 -60 -50 -40 -30 -20 -10 0 10 20membrane potential, V (mV)Figure 10.2: Top: An isochron, or a stable manifold, of a po<strong>in</strong>t x 0 on the limit cycleattractor is the set of all <strong>in</strong>itial conditions y 0 such that y(t) → x(t) as t → +∞.Bottom: Isochrons of the limit cycle attractor <strong>in</strong> Fig. 10.1 correspond<strong>in</strong>g to 40 evenlydistributed phases nT/40, n = 1, . . . , 40.phase space R 2 <strong>in</strong> Fig. 10.1b given by ϑ ↦→ x(ϑ).We could put the <strong>in</strong>itial po<strong>in</strong>t x 0 correspond<strong>in</strong>g to the zero phase anywhere else onthe limit cycle, and not necessarily at the peak of the spike. The choice of the <strong>in</strong>itialpo<strong>in</strong>t <strong>in</strong>troduces an ambiguity <strong>in</strong> parameteriz<strong>in</strong>g the phase of oscillation. Differentparametrizations, however, are equivalent up to a constant phase shift, i.e., translation<strong>in</strong> time. In the rest of the chapter, ϑ always denotes the phase of oscillation, theparameter T denotes the period of oscillation, and ϑ = 0 corresponds to the peak ofthe spike unless stated otherwise. If the system has two or more co-exist<strong>in</strong>g limit cycleattractors, then a separate phase variable needs to be def<strong>in</strong>ed for each attractor.10.1.2 IsochronsThe phase of oscillation can also be <strong>in</strong>troduced outside the limit cycle. Consider, forexample, po<strong>in</strong>t y 0 <strong>in</strong> Fig. 10.2, top. S<strong>in</strong>ce the trajectory y(t) is not on a limit cycle,


460 Synchronization (see www.izhikevich.com)it is not periodic. However, it approaches the cycle as t → +∞. Hence, there is somepo<strong>in</strong>t x 0 on the limit cycle, not necessarily the closest to y 0 , such thaty(t) → x(t) as t → +∞ . (10.2)Now take the phase of the non-periodic solution y(t) to be the phase of its periodicproxy x(t).Alternatively, we can consider a po<strong>in</strong>t on the limit cycle x 0 and f<strong>in</strong>d all the otherpo<strong>in</strong>ts y 0 that satisfy (10.2). The set of all such po<strong>in</strong>ts is called the stable manifoldof x 0 . S<strong>in</strong>ce any solution start<strong>in</strong>g on the stable manifold has an asymptotic behavior<strong>in</strong>dist<strong>in</strong>guishable from that of x(t), its phase is the same as that of x(t). For thisreason, the manifold represents solutions hav<strong>in</strong>g equal phases, and it is often referredto as be<strong>in</strong>g the isochron of x 0 (iso: equal and chronos: time), a notion go<strong>in</strong>g back toBernoulli and Leibniz.Every po<strong>in</strong>t on the plane <strong>in</strong> Fig. 10.2, except the unstable equilibrium, gives rise toa trajectory that approaches the limit cycle. Therefore, every po<strong>in</strong>t has some phase.Let ϑ(x) denote the phase of the po<strong>in</strong>t x. Then, isochrons are level contours of thefunction ϑ(x), s<strong>in</strong>ce the function is constant on each isochron.The entire plane is foliated by isochrons. We depict only 40 representative ones<strong>in</strong> Fig. 10.2. In this chapter we consider neighborhoods of exponentially stable limitcycles, where the foliation is cont<strong>in</strong>uous and <strong>in</strong>variant (Guckenheimer 1975):• Cont<strong>in</strong>uity: The function ϑ(x) is cont<strong>in</strong>uous so that nearby po<strong>in</strong>ts have nearbyphases.• Invariance: If ϑ(x(0)) = ϑ(y(0)), then ϑ(x(t)) = ϑ(y(t)) for all t. Isochrons aremapped to isochrons by the flow of the vector-field.Fig. 10.3 shows the geometry of isochrons of various oscillators. The Andronov-Hopfoscillator <strong>in</strong> the figure is often called a radial isochron clock for the obvious reason.It is simple enough to be solved explicitly, see Ex. 1. In general, f<strong>in</strong>d<strong>in</strong>g isochrons isa daunt<strong>in</strong>g mathematical task. In Ex. 3 we present a MATLAB program that f<strong>in</strong>dsisochrons numerically.10.1.3 PRCConsider a periodically spik<strong>in</strong>g neuron (10.1) receiv<strong>in</strong>g a s<strong>in</strong>gle brief pulse of currentthat <strong>in</strong>creases the membrane potential by A = 1 mV, as <strong>in</strong> Fig. 10.4, left. Such aperturbation may not elicit an immediate spike, but it can change the tim<strong>in</strong>g, i.e., thephase, of the follow<strong>in</strong>g spikes. For example, the perturbed trajectory (solid l<strong>in</strong>e <strong>in</strong>Fig. 10.4, left) fires earlier than the free-runn<strong>in</strong>g unperturbed trajectory (dashed l<strong>in</strong>e).That is, right after the perturbation, the phase, ϑ new , is greater that the old phase,ϑ. The magnitude of the phase shift of the spike tra<strong>in</strong> depends on the exact tim<strong>in</strong>g ofthe stimulus relative to the phase of oscillation ϑ. Stimulat<strong>in</strong>g the neuron at different


Synchronization (see www.izhikevich.com) 4611.5Andronov-Hopf oscillator2van der Pol oscillator10.51.510.5Im z0y0-0.5-1-0.5-1-1.5-1.5-1.5 -1 -0.5 0 0.5 1 1.5Re z-2-1.5 -1 -0.5 0 0.5 1 1.5xI Na +I K -model (Class 1) I Na +I K -model (Class 2)0.60.50.4n 0.3n0.80.70.60.50.40.20.30.10.200.10-80 -60 -40 -20 0 20V-80 -60 -40 -20 0VFigure 10.3: Isochrons of various oscillators. Andronov-Hopf oscillator: ż = (1 + i)z −z|z| 2 , z ∈ C. van der Pol oscillator: ẋ = x − x 3 − y, ẏ = x. The I Na + I K -model withparameters as <strong>in</strong> Fig. 4.1a and I = 10 (Class 1) and I = 35 (Class 2). Only isochronscorrespond<strong>in</strong>g to phases nT/20, n = 1, . . . , 20, are shown.membrane potential, V (mV)phase of stimulation,PRC= new20 0 T/2 T0new-20-40-60-80I(t)0 10 20 30 40 50 60 70 80time (ms)phase resett<strong>in</strong>g, newPRCT/4T/800 T/2phase of stimulation,TFigure 10.4: Phase response of the I Na + I K -model with parameters as <strong>in</strong> Fig. 4.1a andI = 4.7. The dashed voltage trace is the free-runn<strong>in</strong>g trajectory.


462 Synchronization (see www.izhikevich.com)K + activation gate, n0.60.50.40.30.20.10pulsepulsepulsex pulsepulsepulsePRC newypulsenew= +PRC( )pulse-80 -70 -60 -50 -40 -30 -20 -10 0 10 20membrane potential, V (mV)Figure 10.5: The geometrical relationship between isochrons and the phase responsecurve (PRC) of the I Na + I K -oscillator <strong>in</strong> Fig. 10.1.phases, we can measure the phase response curve (also called phase resett<strong>in</strong>g curvePRC, or spike time response curve STRC)PRC (ϑ) = {ϑ new − ϑ} mod T (shift = new phase – old phase) ,depicted <strong>in</strong> Fig. 10.4, right. Positive (negative) values of the function correspond tophase advances (delays) <strong>in</strong> the sense that they advance (delay) the tim<strong>in</strong>g of the nextspike.In contrast to the common folklore, the function PRC (ϑ) can be measured for anarbitrary stimulus, not necessarily weak or brief. The only caveat is that to measurethe new phase of oscillation we need to wait long enough for transients to subside. Thisbecomes a limit<strong>in</strong>g factor when PRCs are used to study synchronization of oscillatorsto periodic pulses, as we do <strong>in</strong> Sect. 10.1.5.There is a simple geometrical relationship between the structure of isochrons of anoscillator and its PRC, illustrated <strong>in</strong> Fig. 10.5, see also Ex. 6. Let us stimulate theoscillator at phase ϑ by a pulse, which moves the trajectory from po<strong>in</strong>t x ly<strong>in</strong>g onthe <strong>in</strong>tersection of isochron ϑ and the limit cycle attractor to a po<strong>in</strong>t y ly<strong>in</strong>g on someisochron ϑ new . From the def<strong>in</strong>ition of PRC, it follows that ϑ new = ϑ+PRC (ϑ).In general, one uses simulations to determ<strong>in</strong>e PRCs, as we do <strong>in</strong> Fig. 10.4. Us<strong>in</strong>g theMATLAB program presented <strong>in</strong> Ex. 5, we can determ<strong>in</strong>e PRCs of all four oscillators <strong>in</strong>Fig. 10.3 and plot them <strong>in</strong> Fig. 10.6. It is a good exercise to expla<strong>in</strong> the shape of eachPRC <strong>in</strong> the figure, or at least its sign, us<strong>in</strong>g the geometry of isochrons of correspond<strong>in</strong>goscillators. In Sect. 10.2.4 we discuss pitfalls of us<strong>in</strong>g the straightforward method <strong>in</strong>Fig. 10.4 to measure PRCs <strong>in</strong> biological neurons, and we present a better technique.Notice that PRC of the I Na + I K -model <strong>in</strong> Fig. 10.6 is ma<strong>in</strong>ly positive <strong>in</strong> the Class 1regime, i.e., when the oscillations appear via saddle-node on <strong>in</strong>variant circle bifurcation,


Synchronization (see www.izhikevich.com) 463Andronov-Hopf oscillatorvan der Pol oscillator0.2Re z(t)0.2x(t)PRC PRC0-0.2 PRC 1PRC 2PRC 2-0.2 PRC1PRC 1 PRC 20 2 4 T0 2 4 6stimulus phase,stimulus phase,I Na +I K -model (Class 1) I Na +I K -model (Class 2)0.5V(t)V(t)-0.5 PRC 2000.2-0.2PRC 1PRCPRC0T0 2 4 6stimulus phase,T0 1 2 3stimulus phase,TFigure 10.6: Examples of phase response curves (PRC) of the oscillators <strong>in</strong> Fig. 10.3.PRC 1 (ϑ): Horizontal pulses (along the first variable) with amplitudes 0.2, 0.2, 2, 0.2 forAndnronov-Hopf, van der Pol, Class 1 and Class 2 oscillators, respectively. PRC 2 (ϑ):Vertical pulses (along the second variable) with amplitudes 0.2, 0.2, 0.02, 0.002, respectively.An example of oscillation is plotted as a dotted curve <strong>in</strong> each subplot (not toscale).but it changes sign <strong>in</strong> the Class 2 regime, correspond<strong>in</strong>g <strong>in</strong> this case to the supercriticalAndronov-Hopf bifurcation. In Sect. 10.4 we f<strong>in</strong>d PRCs analytically <strong>in</strong> the case of weakcoupl<strong>in</strong>g, and show that PRC of a Class 1 oscillator has the shape s<strong>in</strong> 2 ϑ (period T = π)or 1 − cos ϑ (period T = 2π), whereas that of Class 2 oscillator has the shape s<strong>in</strong> ϑ(period T = 2π). We show <strong>in</strong> Sect. 10.1.7 how the synchronization properties of anoscillator depend on the shape of its PRC.10.1.4 Type 0 and 1 phase responseInstead of phase resett<strong>in</strong>g curves, many researchers <strong>in</strong> the field of circadian rhythmsconsider phase transition curves (W<strong>in</strong>free 1980)S<strong>in</strong>ceϑ new = PTC (ϑ old ) .PTC (ϑ) = {ϑ + PRC (ϑ)} mod T ,the two approaches are equivalent. PRCs are convenient when the phase shifts aresmall, so that they can be magnified and clearly seen. PTCs are convenient when the


464 Synchronization (see www.izhikevich.com)Type 1 (weak) resett<strong>in</strong>gType 0 (strong) resett<strong>in</strong>gphase resett<strong>in</strong>g0PRC( ) PRC( )phase resett<strong>in</strong>g00stimulus phase,20stimulus phase,2phase transition2PTC( )={ +PRC( )} modphase transition0PTC( )={ +PRC( )} mod00 2stimulus phase,0stimulus phase,2Figure 10.7: Types of phase resett<strong>in</strong>g of the Andronov-Hopf oscillator <strong>in</strong> Fig. 10.3.phase shifts are large and comparable with the period of oscillation. We present PTCs<strong>in</strong> this section solely for the sake of review, and we use PRCs throughout the rest ofthe chapter.In Fig. 10.7, top, we depict phase portraits of the Andronov-Hopf oscillator hav<strong>in</strong>gradial isochrons and receiv<strong>in</strong>g pulses of magnitude A = 0.5 (left) and A = 1.5 (right).Notice the drastic difference between the correspond<strong>in</strong>g PRCs or PTCs. W<strong>in</strong>free (1980)dist<strong>in</strong>guishes two cases:• Type 1 (weak) resett<strong>in</strong>g results <strong>in</strong> cont<strong>in</strong>uous PRCs and PTCs with mean slope1.• Type 0 (strong) resett<strong>in</strong>g results <strong>in</strong> discont<strong>in</strong>uous PRCs and PTCs with meanslope 0.(Do not confuse these classes with Class 1, 2, and 3 excitability.) The discont<strong>in</strong>uity


Synchronization (see www.izhikevich.com) 4652PTC642021.51stimulus 0.5amplitude, A0642stimulus phase, θstimulus amplitude, A1.81.61.41.210.80.60.40.20.36 0.360.731.101.471.830.732.200.362.572.943.315.88blackhole3.67 4.045.510 2 4 6stimulus phase, θ4.414.785.155.88 5.885.51Figure 10.8: Time crystal (left) and its contour plot (right). Shown is the PTC (ϑ, A)of the Andronov-Hopf oscillator (see Ex. 4).phase of oscillationn nPRC nnT spulse n pulse n+1timen+1 n PRC n T sFigure 10.9: Calculation ofthe Po<strong>in</strong>care phase map.of Type 0 PRC <strong>in</strong> Fig. 10.7 is a topological property that cannot be removed byreallocat<strong>in</strong>g the <strong>in</strong>itial po<strong>in</strong>t x 0 that corresponds to zero phase. As an exercise, provethat the discont<strong>in</strong>uity stems from the fact that the shifted image of the limit cycle(dashed circle) goes beyond the central equilibrium at which the phase is not def<strong>in</strong>ed.If we vary not only the phase ϑ of the applied stimulus, but also its amplitude A,then we obta<strong>in</strong> parameterized PRC and PTC. In Fig. 10.8 we plot PTC (ϑ, A) of theAndronov-Hopf oscillator (the correspond<strong>in</strong>g PRC is derived <strong>in</strong> Ex. 4). The surfaceis called time crystal and it can take quite amaz<strong>in</strong>g shapes (W<strong>in</strong>free 1980). Thecontour plot of PTC (ϑ, A) <strong>in</strong> the figure conta<strong>in</strong>s the s<strong>in</strong>gularity po<strong>in</strong>t (black hole) thatcorresponds to the phaseless equilibrium of the Andronov-Hopf oscillator. Stimulationat phase ϑ = π with magnitude A = 1 pushes the trajectory <strong>in</strong>to the equilibrium andstalls the oscillation.


466 Synchronization (see www.izhikevich.com)(a)membrane potential, V(c)phase, n101 2 3 4 5 6 7?T stime, tT345 6 71K + gat<strong>in</strong>g variable, b(b)0.60.40.2(d)Tmod Tn+10314-7-80 -60 -40 -20 0 20membrane potential, V102Po<strong>in</strong>care phase map315,6,74stablefixedpo<strong>in</strong>t021 2 3 4 5 6 7pulse number, n020 10nTFigure 10.10: Description of synchronization of I Na + I K -oscillator <strong>in</strong> Fig. 10.4 us<strong>in</strong>gPo<strong>in</strong>care phase map.10.1.5 Po<strong>in</strong>care phase mapThe phase resett<strong>in</strong>g curve (PRC) describes the response of an oscillator to a s<strong>in</strong>glepulse, but it can be used to study its response to a periodic pulse tra<strong>in</strong> us<strong>in</strong>g thefollow<strong>in</strong>g “stroboscopic” approach. Let ϑ n denote the phase of oscillation at the timethe nth <strong>in</strong>put pulse arrives. Such a pulse resets the phase by PRC (ϑ n ), so that the newphase right after the pulse is ϑ n +PRC (ϑ n ); see Fig. 10.9. Let T s denote the period ofpulsed stimulation. Then the phase of oscillation before the next, (n + 1)th pulse, isϑ n +PRC (ϑ n ) + T s . Thus, we have a stroboscopic mapp<strong>in</strong>g of a circle to itselfϑ n+1 = (ϑ n + PRC (ϑ n ) + T s ) mod T (10.3)called Po<strong>in</strong>care phase map (two pulse-coupled oscillators are considered <strong>in</strong> Ex. 11).Know<strong>in</strong>g the <strong>in</strong>itial phase of oscillation ϑ 1 at the first pulse, we can determ<strong>in</strong>e ϑ 2 , thenϑ 3 , etc. The sequence {ϑ n } with n = 1, 2, . . . , is called the orbit of the map, and it isquite easy to f<strong>in</strong>d numerically.Let us illustrate this concept us<strong>in</strong>g the I Na + I K -oscillator with PRC shown <strong>in</strong>Fig. 10.4. Its free-runn<strong>in</strong>g period is T ≈ 21.37 ms, and the period of stimulation<strong>in</strong> Fig. 10.10a is T s = 18.37, which results <strong>in</strong> the Po<strong>in</strong>care phase map depicted <strong>in</strong>Fig. 10.10d. The cobweb <strong>in</strong> the figure is the orbit go<strong>in</strong>g from ϑ 1 to ϑ 2 to ϑ 3 , etc. Notice


Synchronization (see www.izhikevich.com) 467that the phase ϑ 3 cannot be measured directly from the voltage trace <strong>in</strong> Fig. 10.10abecause pulse 2 changes the phase, so it is not the time s<strong>in</strong>ce the last spike when pulse3 arrives. The Po<strong>in</strong>care phase map (10.3) takes <strong>in</strong>to account such multiple pulses.The orbit approaches a po<strong>in</strong>t (called fixed po<strong>in</strong>t, see below) that corresponds to asynchronized or phase-locked state.A word of caution is <strong>in</strong> order. Recall that PRCs are measured on the limit cycleattractor. However, each pulse displaces the trajectory away from the attractor, as<strong>in</strong> Fig. 10.5. To use the PRC formalism to describe the effect of the next pulse, theoscillator must be given enough time to relax back to the limit cycle attractor. Thus,if the period of stimulation T s is too small, or the attraction to the limit cycle is tooslow, or the stimulus amplitude is too large, then the Po<strong>in</strong>care phase map may be notan appropriate tool to describe the phase dynamics.10.1.6 Fixed po<strong>in</strong>tsTo understand the structure of orbits of the Po<strong>in</strong>care phase map (10.3), or any othermapϑ n+1 = f(ϑ n ) , (10.4)we need to f<strong>in</strong>d its fixed po<strong>in</strong>tsϑ = f(ϑ)(ϑ is a fixed po<strong>in</strong>t),which are analogues of equilibria of cont<strong>in</strong>uous dynamical systems. Geometrically, afixed po<strong>in</strong>t is the <strong>in</strong>tersection of the graph of f(ϑ) with the diagonal l<strong>in</strong>e ϑ n+1 = ϑ n ;see Fig. 10.10d or Fig. 10.11. At such a po<strong>in</strong>t, the orbit ϑ n+1 = f(ϑ n ) = ϑ n is fixed.A fixed po<strong>in</strong>t ϑ is asymptotically stable if it attracts all nearby orbits, i.e., if ϑ 1 is <strong>in</strong>a sufficiently small neighborhood of ϑ, then ϑ n → ϑ as n → ∞, as <strong>in</strong> Fig. 10.11, left.The fixed po<strong>in</strong>t is unstable if any small neighborhood of the po<strong>in</strong>t conta<strong>in</strong>s an orbitdiverg<strong>in</strong>g from it, as <strong>in</strong> Fig. 10.11, right.The stability of the fixed po<strong>in</strong>t is determ<strong>in</strong>ed by the slopem = f ′ (ϑ)of the graph of f at the po<strong>in</strong>t, which is called Floquet multiplier of the mapp<strong>in</strong>g. It playsthe same role as the eigenvalue λ of an equilibrium of a cont<strong>in</strong>uous dynamical system.Mnemonically, the relationship between them is µ = e λ , to that the fixed po<strong>in</strong>t is stablewhen |m| < 1 (λ < 0) and unstable when |m| > 1 (λ > 0). Fixed po<strong>in</strong>ts bifurcatewhen |m| = 1 (λ is zero or purely imag<strong>in</strong>ary). They lose stability via flip bifurcation (adiscrete analogue of Andronov-Hopf bifurcation) when m = −1 and disappear via foldbifurcation (a discrete analogue of saddle-node bifurcation) when m = 1. The formerplays an important role <strong>in</strong> the period-doubl<strong>in</strong>g phenomenon illustrated <strong>in</strong> Fig. 10.14,bottom trace. The latter plays an important role <strong>in</strong> the cycle slipp<strong>in</strong>g phenomenonillustrated <strong>in</strong> Fig. 10.16.


468 Synchronization (see www.izhikevich.com)stable fixed po<strong>in</strong>tsunstable fixed po<strong>in</strong>tsnfn+1fn+1nnnnnn+1fnn+1fnnnnnFigure 10.11: The stability of fixed po<strong>in</strong>ts of the mapp<strong>in</strong>g (10.4) depends on the slopeof the function f.10.1.7 SynchronizationWe say that two periodic pulse tra<strong>in</strong>s are synchronous, when the pulses occur at thesame times or with a constant phase shift, as <strong>in</strong> Fig. 10.12a. Each subplot <strong>in</strong> the figureconta<strong>in</strong>s an <strong>in</strong>put pulse tra<strong>in</strong> (bottom) and an output spike tra<strong>in</strong> (top), assum<strong>in</strong>gthat spikes are fired at zero cross<strong>in</strong>gs of the phase variable, as <strong>in</strong> Fig. 10.1. Sucha synchronized state corresponds to a stable fixed po<strong>in</strong>t of the Po<strong>in</strong>care phase map(10.3). The <strong>in</strong>-phase, anti-phase, or out-of-phase synchronization corresponds to thephase shift ϑ = 0, ϑ = T/2, or ϑ equal some other value, respectively. Many scientistrefer to the <strong>in</strong>-phase synchronization as just “synchronization”, and use the adjectivesanti-phase or out-of-phase to denote the other types of synchronization.When the period of stimulation, T s , is near the free-runn<strong>in</strong>g period of tonic spik<strong>in</strong>g,T , then the fixed po<strong>in</strong>t of (10.3) satisfiesPRC (ϑ) = T − T s ,i.e., it is the <strong>in</strong>tersection of the PRC and the horizontal l<strong>in</strong>e, as <strong>in</strong> Fig. 10.13. Thus,synchronization occurs with such a phase shift ϑ that compensates the <strong>in</strong>put periodmismatch T − T s . The maxima and the m<strong>in</strong>ima of the PRC determ<strong>in</strong>e the toleranceof the oscillator to the mismatch. As an exercise, check that stable fixed po<strong>in</strong>ts lie onthe side of the graph with the slope−2 < PRC ′ (ϑ) < 0(stability region)


Synchronization (see www.izhikevich.com) 469(a)synchronization<strong>in</strong>-phase anti-phase out-of-phase(b)phase-lock<strong>in</strong>g1:2 2:13:2Figure 10.12: Examples of fundamental types of synchronization of spik<strong>in</strong>g activity toperiodic pulsed <strong>in</strong>puts (synchronization is 1:1 phase-lock<strong>in</strong>g).T-T sunstablestabilityregionPRC00stablephaseshiftstimulus phase,stabilityregionTFigure 10.13: Fixed po<strong>in</strong>ts of the Po<strong>in</strong>carephase map (10.1.5).marked by the bold curves <strong>in</strong> Fig. 10.13.Now consider the Class 1 and Class 2 I Na + I K -oscillators shown <strong>in</strong> Fig. 10.6. ThePRC <strong>in</strong> Class 1 regime is mostly positive, imply<strong>in</strong>g that such an oscillator can easilysynchronize with faster <strong>in</strong>puts (T − T s > 0) but cannot synchronize with slower <strong>in</strong>puts.Indeed, the oscillator can only advance its phase to catch up faster pulse tra<strong>in</strong>s, but itcannot delay the phase to wait for the slower <strong>in</strong>put. Synchronization with the <strong>in</strong>puthav<strong>in</strong>g T s ≈ T is only marg<strong>in</strong>al. In contrast, Class 2 I Na + I K -oscillator does not havethis problem because its PRC has well-def<strong>in</strong>ed positive and negative regions.10.1.8 Phase lock<strong>in</strong>gp:q-phase-lock<strong>in</strong>g occurs when the oscillator fires p spikes for every q <strong>in</strong>put pulses,such as the 3:2-phase-lock<strong>in</strong>g <strong>in</strong> Fig. 10.12b or 2:2 phase-lock<strong>in</strong>g <strong>in</strong> Fig. 10.14, whichtypically occurs when pT ≈ qT s . The <strong>in</strong>tegers p and q need not be relatively prime <strong>in</strong>the case of pulsed-coupled oscillators. Synchronization, i.e., 1:1 phase-lock<strong>in</strong>g, as wellas p:1 phase-lock<strong>in</strong>g corresponds to a fixed po<strong>in</strong>t of the Po<strong>in</strong>care phase map (10.3) withp fired spikes per one <strong>in</strong>put pulse. Indeed, the map tells only the phase of the oscillatorat each pulse, but does not tell the number of oscillations made between the pulses.Each p:q-locked solution corresponds to a stable periodic orbit of the Po<strong>in</strong>carephase map with the period q (so that ϑ n = ϑ n+q for any n). Such orbits <strong>in</strong> maps (10.4)


470 Synchronization (see www.izhikevich.com)synchronization (1:1)po<strong>in</strong>care phase mapphase-lockedphase-lock<strong>in</strong>g (2:2)n+1=f nsync0nFigure 10.14: Co-existence of synchronized and phase-locked solutions corresponds toco-existence of stable fixed po<strong>in</strong>t and a stable periodic orbit of the Po<strong>in</strong>care phase map.correspond to stable equilibria <strong>in</strong> the iterates ϑ k+1 = f q (ϑ k ), where f q = f ◦ f ◦ · · · ◦ fis the composition of f with itself q times. Geometrically, study<strong>in</strong>g such maps is likeconsider<strong>in</strong>g every q-th <strong>in</strong>put pulse <strong>in</strong> Fig. 10.12b and ignor<strong>in</strong>g all the <strong>in</strong>termediatepulses.S<strong>in</strong>ce maps can have co-existence of stable fixed po<strong>in</strong>ts and periodic orbits, varioussynchronized and phase-lock<strong>in</strong>g states can co-exist <strong>in</strong> response to the same <strong>in</strong>put pulsetra<strong>in</strong>, as <strong>in</strong> Fig. 10.14. The oscillator converges to one of the states depend<strong>in</strong>g on the<strong>in</strong>itial phase of oscillation, but could be switched between the states by a transient<strong>in</strong>put.10.1.9 Arnold tonguesTo synchronize an oscillator, the <strong>in</strong>put pulse tra<strong>in</strong> must have period T s sufficiently nearthe oscillator’s free-runn<strong>in</strong>g period T so that the graph of PRC and the horizontal l<strong>in</strong>e<strong>in</strong> Fig. 10.13 <strong>in</strong>tersect. The amplitude of the function |PRC (ϑ, A)| decreases as thestrength of the pulse A decreases, because weaker pulses produce weaker phase shifts.Hence the region of existence of a synchronized state shr<strong>in</strong>ks as A → 0, and it looks likea horn or a tongue on the (T s , A)-plane depicted <strong>in</strong> Fig. 10.15, called Arnold tongue.Each p:q-phase-locked state has its own region of existence (p:q-tongue <strong>in</strong> the figure),also shr<strong>in</strong>k<strong>in</strong>g to a po<strong>in</strong>t pT/q on the T s -axis. The larger the order of lock<strong>in</strong>g, p + q,the narrower the tongue, and the more difficult it is to observe such a phase-lockedstate numerically, let alone experimentally.The tongues can overlap, lead<strong>in</strong>g to the co-existence of phase-locked states, as <strong>in</strong>Fig. 10.14. If A is sufficiently large, the Po<strong>in</strong>care phase map (10.3) becomes non<strong>in</strong>vertible,i.e., it has a region of negative slope, and there is a possibility of chaoticdynamics (Glass and MacKey 1988).In Fig. 10.16 we illustrate the cycle slipp<strong>in</strong>g phenomenon that occurs when the <strong>in</strong>putperiod T s drifts away from the 1:1 Arnold tongue. The fixed po<strong>in</strong>t of the Po<strong>in</strong>carephase map correspond<strong>in</strong>g to the synchronized state undergoes a fold bifurcation anddisappears. Similarly to the case of saddle-node on <strong>in</strong>variant circle bifurcation, the


Synchronization (see www.izhikevich.com) 471amplitude of stimulation, A1:61:4synchronization1:2 3:21:11:3 2:32:15:20T/4 T/3 T/2 3T/4 T 5T/4 3T/2 7T/4 2T 9T/4 5T/2period of stimulation, TsFigure 10.15: Arnold tongues are regions of existence of various phase-locked states onthe “period-strength” plane.cycle slipp<strong>in</strong>gT/2Po<strong>in</strong>care phase mapn+1=f( n)0?ghost ofattractor-T/2-T/2 0 n T/2Figure 10.16: Cycle slipp<strong>in</strong>g phenomenon at the edge of the Arnold tongue correspond<strong>in</strong>gto a synchronized state.fold fixed po<strong>in</strong>t becomes a ghost attractor that traps orbits and keeps them near thesynchronized state for a long period of time. Eventually the orbit escapes, the synchronizedstate is briefly lost, and then the orbit returns to the ghost attractor to betrapped aga<strong>in</strong>. Such an <strong>in</strong>termittently synchronized orbit typically corresponds to ap:q-phase-locked state with high order of lock<strong>in</strong>g p + q.10.2 Weak Coupl<strong>in</strong>gIn this section we consider dynamical systems of the formẋ = f(x) + εp(t) , (10.5)describ<strong>in</strong>g periodic oscillators, ẋ = f(x), forced by a time-depended <strong>in</strong>put εp(t), e.g.,from other oscillators <strong>in</strong> a network. The positive parameter ε measures the overallstrength of the <strong>in</strong>put, and it is assumed to be sufficiently small, denoted as ε ≪ 1. Wedo not assume ε → 0 here. In fact, most of the results <strong>in</strong> this section can be cast <strong>in</strong>


472 Synchronization (see www.izhikevich.com)Figure 10.17: Arthur W<strong>in</strong>free, 2001.the form “there is an ε 0 such that for all ε < ε 0 , the follow<strong>in</strong>g holds...” (Hoppensteadtand Izhikevich 1997), with ε 0 depend<strong>in</strong>g on the function f(x) <strong>in</strong> (10.5) and sometimestak<strong>in</strong>g not so small values, e.g., ε 0 = 1.Notice that if ε = 0 <strong>in</strong> (10.5), then we can transform ẋ = f(x) to ˙ϑ = 1 us<strong>in</strong>gthe theory presented <strong>in</strong> the previous section. What happens when we apply the sametransformation to (10.5) with ε ≠ 0? In this section we present three different butequivalent approaches that transform (10.5) <strong>in</strong>to the phase model˙ϑ = 1 + ε PRC (ϑ)p(t) + o(ε) .Here, the Landau’s “little oh” function o(ε) denotes the error terms smaller than ε sothat o(ε)/ε → 0 if ε → 0. For the sake of clarity of notation, we omit o(ε) throughoutthe book, and implicitly assume that all equalities are valid up to the terms of ordero(ε).S<strong>in</strong>ce we do not impose restrictions on the form of p(t), the three methods arereadily applicable to the casep(t) = ∑ sg s (x(t), x s (t)) ,where the set {x s (t)} denotes oscillators <strong>in</strong> the network connected to x, and p(t) is thepostsynaptic current.10.2.1 W<strong>in</strong>free’s approachA sufficiently small neighborhood of the limit cycle attractor of the unperturbed(ε = 0) oscillator (10.5), magnified <strong>in</strong> Fig. 10.18, has nearly coll<strong>in</strong>ear uniformly spacedisochrons. Coll<strong>in</strong>earity implies that a po<strong>in</strong>t x on the limit cycle <strong>in</strong> Fig. 10.18 has thesame phase-resett<strong>in</strong>g as any other po<strong>in</strong>t y on the isochron of x near the cycle. Uniformdensity of isochrons implies that the phase resett<strong>in</strong>g scales l<strong>in</strong>early with the strengthof the pulse, i.e., a half-pulse at the po<strong>in</strong>t z <strong>in</strong> Fig. 10.18 produces a half-resett<strong>in</strong>g ofthe phase.


Synchronization (see www.izhikevich.com) 473n0.007xyz0.60.50.4n 0.30.20.005limit cycle attractor0.10-62 -61 V -60 -59-80 -60 -40 -20 0 20VFigure 10.18: Magnification of isochrons <strong>in</strong> a small neighborhood of the limit cycle ofthe I Na + I K -model <strong>in</strong> Fig. 10.3. Isochron time step: 0.025 ms on the left, 0.35 ms onthe right.p(t 2 )hp(t)hp(t 2 )t 0 t 1 t 2 t 3 t 4 t 5time, tFigure 10.19: A cont<strong>in</strong>uous function p(t) is replaced by an equivalent tra<strong>in</strong> of pulses ofvariable amplitudes.L<strong>in</strong>ear scal<strong>in</strong>g of PRC with respect to the strength of the pulse motivates thesubstitutionPRC (ϑ, A) ≈ Z(ϑ)A ,where Z(ϑ) = ∂ PRC(ϑ, A)/∂A at A = 0 is the l<strong>in</strong>ear response or sensitivity function(W<strong>in</strong>free 1967) describ<strong>in</strong>g the slight alteration of rate, or of <strong>in</strong>stantaneous frequency ofoscillation, accompany<strong>in</strong>g application of a small stimulus. Some call it the <strong>in</strong>f<strong>in</strong>itesimalPRC.Now suppose ε ≠ 0 but sufficiently small, so that the trajectory of the weaklyperturbed oscillator (10.5) rema<strong>in</strong>s near the limit cycle attractor all the time. Letus replace the cont<strong>in</strong>uous <strong>in</strong>put function εp(t) by the equivalent tra<strong>in</strong> of pulses ofstrength A = εp(t n )h, where h is a small <strong>in</strong>terpulse <strong>in</strong>terval (denoted as T s <strong>in</strong> theprevious section), and t n = nh is the tim<strong>in</strong>g of the n-th pulse, see Fig. 10.19. We


474 Synchronization (see www.izhikevich.com)Figure 10.20: Yoshiki Kuramoto <strong>in</strong>1988, while he was visit<strong>in</strong>g Jim Murray’s<strong>in</strong>stitute at University of Oxford.rewrite the correspond<strong>in</strong>g Po<strong>in</strong>care phase map (10.3)PRC{ }} {ϑ(t n+1 ) = {ϑ(t n ) + Z(ϑ(t n )) εp(t n )h +h} mod T} {{ }A<strong>in</strong> the formϑ(t n + h) − ϑ(t n )hwhich is a discrete version of= Z(ϑ(t n ))εp(t n ) + 1 ,˙ϑ = 1 + εZ(ϑ) · p(t) , (10.6)<strong>in</strong> the limit h → 0.To be consistent with all the examples <strong>in</strong> the previous section, we implicitly assumehere that p(t) perturbs only the first, voltage-like variable x 1 of the state vector x =(x 1 , . . . , x m ) ∈ R m and Z(ϑ) is the correspond<strong>in</strong>g sensitivity function. However, thephase model (10.6) is also valid for an arbitrary <strong>in</strong>put p(t) = (p 1 (t), . . . , p m (t)). Indeed,let Z i describe the l<strong>in</strong>ear response to perturbations of the ith state variable x i , andZ(ϑ) = (Z 1 (ϑ), . . . , Z m (ϑ)) denote the correspond<strong>in</strong>g l<strong>in</strong>ear response vector-function.Then, the comb<strong>in</strong>ed phase shift Z 1 p 1 + · · · + Z m p m is just the dot product Z · p <strong>in</strong>(10.6).10.2.2 Kuramoto’s approachConsider the unperturbed (ε = 0) oscillator (10.5), and let the function ϑ(x) denotethe phases of po<strong>in</strong>ts near its limit cycle attractor. Obviously, isochrons are the levelcontours of ϑ(x) s<strong>in</strong>ce the function is constant on each isochron. Differentiat<strong>in</strong>g thefunction us<strong>in</strong>g the cha<strong>in</strong> rule yieldsdϑ(x)dt= grad ϑ · dxdt= grad ϑ · f(x) ,


Synchronization (see www.izhikevich.com) 475xgrad (x)isochronf(x)limit cycleFigure 10.21: Geometrical <strong>in</strong>terpretationof the vector grad ϑ.where grad ϑ = (ϑ x1 (x), . . . , ϑ xm (x)) is the gradient of ϑ(x) with respect to the statevector x = (x 1 , . . . , x m ) ∈ R m . However,dϑ(x)dtnear the limit cycle, because isochrons are mapped to isochrons by the flow of thevector-field f(x). Therefore, we get a useful equality= 1grad ϑ · f(x) = 1 . (10.7)Figure 10.21 shows a geometrical <strong>in</strong>terpretation of grad ϑ(x): it is the vector based atpo<strong>in</strong>t x, normal to the isochron of x and with the length equal to the number densityof isochrons at x. Its length can also be found from (10.7).Kuramoto (1984) applied the cha<strong>in</strong> rule to the perturbed system (10.5)dϑ(x)dt= grad ϑ · dxdtand, us<strong>in</strong>g (10.7), obta<strong>in</strong>ed the phase model= grad ϑ · {f(x) + εp(t)} = grad ϑ · f(x) + ε grad ϑ · p(t) ,˙ϑ = 1 + ε grad ϑ · p(t) , (10.8)which has the same form as (10.6). Subtract<strong>in</strong>g (10.8) from (10.6) yields (Z(ϑ) −grad ϑ) · p(t) = 0. S<strong>in</strong>ce this is valid for any p(t), we conclude that Z(ϑ) = grad ϑ;see also Ex. 6. Thus, Kuramoto’s phase model (10.8) is <strong>in</strong>deed equivalent to W<strong>in</strong>free’smodel (10.8).10.2.3 Malk<strong>in</strong>’s approachYet another equivalent method of reduction of weakly perturbed oscillators to theirphase models follows from Malk<strong>in</strong> theorem (1949,1956), which we state <strong>in</strong> the simplestform below. The most abstract form and its proof is provided by Hoppensteadt andIzhikevich (1997).Malk<strong>in</strong>’s theorem. Suppose the unperturbed (ε = 0) oscillator <strong>in</strong> (10.5) has anexponentially stable limit cycle of period T . Then its phase is described by the equation˙ϑ = 1 + εQ(ϑ) · p(t) , (10.9)


476 Synchronization (see www.izhikevich.com)Figure 10.22: Ioel Gil’evich Malk<strong>in</strong> (IoзlьGilьeviq Malk<strong>in</strong>, 1907-1958).where the T -periodic function Q is the solution to the l<strong>in</strong>ear “adjo<strong>in</strong>t” equation˙Q = −{Df(x(t))} ⊤ Q , with Q(0) · f(x(0)) = 1 , (10.10)where Df(x(t)) ⊤ is the transposed Jacobian of f (matrix of partial derivatives) atthe po<strong>in</strong>t x(t) on the limit cycle, and the normalization condition can be replaced byQ(t) · f(x(t)) = 1 for any and hence all t (prove it). Here Q · f is the dot product oftwo vectors, which is the same as Q ⊤ f.Though this theorem looks less <strong>in</strong>tuitive than the methods of W<strong>in</strong>free and Kuramoto,it is actually more useful because (10.10) can be solved numerically quiteeasily. Apply<strong>in</strong>g the MATLAB procedure <strong>in</strong> Ex. 12 to the four oscillators <strong>in</strong> Fig. 10.3,we plot their functions Q <strong>in</strong> Fig. 10.23. It is not a co<strong>in</strong>cidence that each componentof Q looks like PRC along the first or the second state variable, respectively, shown <strong>in</strong>Fig. 10.6. Subtract<strong>in</strong>g (10.9) from (10.8) or from (10.6), we conclude thatZ(ϑ) = grad ϑ(x) = Q(ϑ) ,(see also Ex. 7), so that we can determ<strong>in</strong>e the l<strong>in</strong>ear response function of the phasemodel us<strong>in</strong>g any of the three alternative methods: via PRCs, via isochrons, or solv<strong>in</strong>gthe adjo<strong>in</strong>t equation (10.10). This justifies why many refer to the function as just PRC,implicitly assum<strong>in</strong>g that it is measured to the <strong>in</strong>f<strong>in</strong>itesimal stimuli and then normalizedby the stimulus amplitude.10.2.4 Measur<strong>in</strong>g PRCs experimentallyIn Fig. 10.24 we exploit the relationship (10.9) and measure the <strong>in</strong>f<strong>in</strong>itesimal PRCsof a layer 5 pyramidal neuron of mouse visual cortex. First, we stimulate the neuronwith 40 pA dc-current to elicit periodic spik<strong>in</strong>g. Initially, the fir<strong>in</strong>g period starts at50 ms, and then relaxes to the averaged value of 110 ms (Fig. 10.24a). The standardmethod of f<strong>in</strong>d<strong>in</strong>g PRCs consists <strong>in</strong> stimulat<strong>in</strong>g the neuron by brief pulses of currentat different phases of the cycle and measur<strong>in</strong>g the <strong>in</strong>duced phase shift, which couldbe approximated by the difference between two successive periods of oscillation. Themethod works f<strong>in</strong>e <strong>in</strong> models, see Ex. 5, but should be used with caution <strong>in</strong> real neuronsbecause their fir<strong>in</strong>g is too noisy, as we demonstrate <strong>in</strong> Fig. 10.24b. Thus, one needs


Synchronization (see www.izhikevich.com) 4771Andronov-Hopf oscillator1Q 1 ( )Q( )Q 2 ( )Q 2Q 1 phase,1Q 2Q 1van der Pol oscillatorQQ( )1 ( )1Q 2 ( )0000Q 1 Q 1-1-1-1-1 0 1 0 2 4 6-1-2 0 2 0 2phase,4 6Q 2 Q 21000.2 Q 1 ( )800Q( ) 0.01Q 2 ( )Q( )I Na +I K -model (Class 1) I Na +I K -model (Class 2)-1006005-20400-30-0.22000-40-0.1 0 0.1 0.2-0.40.01Q 2 ( )0 2 4 6phase,0-200-5 0 5-5Q 1 ( )0 1 2 3phase,Figure 10.23: Solutions Q = (Q 1 , Q 2 ) to adjo<strong>in</strong>t problem (10.10) for oscillators <strong>in</strong>Fig. 10.3.to apply hundreds if not thousands of pulses and then average the result<strong>in</strong>g phasedeviations (Reyes and Fetz 1993).Start<strong>in</strong>g with time 10s we <strong>in</strong>ject a relatively weak noisy current εp(t) that cont<strong>in</strong>uouslyperturbs the membrane potential (Fig. 10.24c) and hence the phase of oscillation(the choice of p(t) is important; its Fourier spectrum must span a range of frequenciesthat depends on the frequency of fir<strong>in</strong>g of the neuron). Know<strong>in</strong>g εp(t), the momentsof fir<strong>in</strong>g of the neuron, i.e., zero cross<strong>in</strong>gs ϑ(t) = 0, and the relationship˙ϑ = 1 + PRC (ϑ)εp(t) ,we solve the <strong>in</strong>verse problem for the <strong>in</strong>f<strong>in</strong>itesimal PRC (ϑ) and plot the solution <strong>in</strong>Fig. 10.24d. As one expects, the PRC is mostly positive, maximal just before the spikeand almost zero dur<strong>in</strong>g the spike. It would resemble the PRC <strong>in</strong> Fig. 10.23 (Q 1 (ϑ) <strong>in</strong>Class 1) if not for the dip <strong>in</strong> the middle, for which we have no explanation (probably itis due to overfitt<strong>in</strong>g). The advantage of this method is that it is more immune to noise,because <strong>in</strong>tr<strong>in</strong>sic fluctuations are spread over the entire p(t) and not concentrated atthe moments of pulses, unless of course p(t) consists of random pulses, <strong>in</strong> which casethis method is equivalent to the standard one. The drawback is that we need to solvethe equation above, which we do <strong>in</strong> Ex. 13 us<strong>in</strong>g an optimization technique.


478 Synchronization (see www.izhikevich.com)<strong>in</strong>terspike periods, Ti (ms)15010050membranepotentia(a)<strong>in</strong>jected current, 40 pAp(t)0 5 time (sec) 10 15 -T/20 2 4 6time (sec)8 100.3(c)(d) pyramidal neuron8T i T i+1-30 mV010 mV10 pAp(t) -0.10phase of oscillation,Tperiod difference, Ti+1-TiT/2PRC( ), (pA -1 )0(b)averaged spike1(e)Andronov-Hopf (f)1I Na +I K -model (Class 1)0.2023012 3-10 2 4 6phase of oscillation,0 2 4 6phase of oscillation,Figure 10.24: Measur<strong>in</strong>g the <strong>in</strong>f<strong>in</strong>itesimal PRC experimentally <strong>in</strong> a layer 5 pyramidalneuron of mouse visual cortex. (a) Interspike periods <strong>in</strong> response to the <strong>in</strong>jection ofdc-current. (b) Differences between successive periods. (c) Spik<strong>in</strong>g 1 second before andafter the noisy current p(t) is <strong>in</strong>jected. (d) Inf<strong>in</strong>itesimal PRC of the neuron (cont<strong>in</strong>uouscurve) obta<strong>in</strong>ed from 40 cycles and the MATLAB program <strong>in</strong> Ex. 13 (first 8 Fourierterms). Averaged voltage trace dur<strong>in</strong>g the spike (dotted curve) is plotted for reference.The same procedure applied to (e) the Andronov-Hopf oscillator and (f) the I Na,p +I K -model. Numbers <strong>in</strong> boxes show the number of Fourier terms used to fit the curve,theoretical curves (functions Q 1 (ϑ) from Fig. 10.23) are dashed.


Synchronization (see www.izhikevich.com) 479membranepotential (mV)phasephasedeviationphasedifference200-20-40-60-80T0T0TV 1 V 22122 1100 5 10 15 20 25 30 35 40 45 50time (ms)Figure 10.25: The relationship between membrane potential oscillation of two neurons,V 1 (solid) and V 2 (dashed), their phases, phase deviations, and phase difference. Shownare simulation of two I Na + I K -models with parameters as <strong>in</strong> Fig. 10.3 and coupledsymmetrically via gap junctions 0.1(V j − V i ) (see Sect. 2.3.4).10.2.5 Phase model for coupled oscillatorsNow consider n weakly coupled oscillators of the formp i (t){ }} {n∑ẋ i = f i (x i ) + ε g ij (x i , x j ) , x i ∈ R m , (10.11)j=1and assume that the oscillators, when uncoupled (ε = 0), have equal free-runn<strong>in</strong>gperiods T 1 = · · · = T n = T . Apply<strong>in</strong>g any of the three methods above to such a weaklyperturbed system, we obta<strong>in</strong> the correspond<strong>in</strong>g phase model˙ϑ i = 1 + ε Q i (ϑ i ) ·p i (t){ }} {n∑g ij (x i (ϑ i ), x j (ϑ j )) , (10.12)j=1where each x i (ϑ i ) is the po<strong>in</strong>t on the limit cycle hav<strong>in</strong>g phase ϑ i . Note that (10.11) isdef<strong>in</strong>ed <strong>in</strong> R nm , whereas the phase model (10.12) is def<strong>in</strong>ed on the n-torus, denoted asT n .To study collective properties of the network, such as synchronization, it is convenientto represent each ϑ i (t) asϑ i (t) = t + ϕ i , (10.13)


480 Synchronization (see www.izhikevich.com)Andronov-Hopf oscillatorvan der Pol oscillator0.5H ij ( )0.5H ij ( )00-0.5-0.5G( )G( )0 2 4 T0 2 4 6T20I Na +I K -model (Class 1) I Na +I K -model (Class 2)H ij ( )4H ij ( )0-2G( )-4G( )0 2 4 6phase difference,T0 1 2 3phase difference,TFigure 10.26: Solid curves: Functions H ij (χ) def<strong>in</strong>ed by (10.16) with the <strong>in</strong>putg(x i , x j ) = (x j1 − x i1 , 0) correspond<strong>in</strong>g to electrical synapse via gap-junction. Dashedcurves: Functions G(χ) = H ji (−χ) − H ij (χ). Parameters as <strong>in</strong> Fig. 10.3.with the first term captur<strong>in</strong>g fast free-runn<strong>in</strong>g natural oscillation ˙ϑ i = 1, and thesecond term captur<strong>in</strong>g slow network-<strong>in</strong>duced build-up of phase deviation from thenatural oscillation. The relationship between x i (t), ϑ i (t) and ϕ i (t) is illustrated <strong>in</strong>Fig. 10.25.Substitut<strong>in</strong>g (10.13) <strong>in</strong>to (10.12) results <strong>in</strong>˙ϕ i = ε Q i (t + ϕ i ) ·n∑g ij (x i (t + ϕ i ), x j (t + ϕ j )) . (10.14)j=1Notice that the right hand-side is of order ε, reflect<strong>in</strong>g the slow dynamics of phasedeviations ϕ i seen <strong>in</strong> Fig. 10.25. Thus, it conta<strong>in</strong>s two time scales: fast oscillations(variable t) and slow phase modulation of phase (variables ϕ). The classical methodof averag<strong>in</strong>g, reviewed by Hoppensteadt and Izhikevich (1997, Chap. 9) consists <strong>in</strong> anear-identity change of variables that transforms the system <strong>in</strong>to the formwhereH ij (ϕ j − ϕ i ) = 1 T˙ϕ i = εω i + ε∫ T0n∑H ij (ϕ j − ϕ i ) , (10.15)j≠iQ i (t) · g ij (x i (t), x j (t + ϕ j − ϕ i )) dt , (10.16)and each ω i = H ii (ϕ i − ϕ i ) = H ii (0) describes a constant frequency deviation fromthe free-runn<strong>in</strong>g oscillation. Figure 10.26 depicts the functions H ij correspond<strong>in</strong>g togap-junction (i.e., electrical; see Sect. 2.3.4) coupl<strong>in</strong>g of oscillators <strong>in</strong> Fig. 10.3. Prove


Synchronization (see www.izhikevich.com) 481(a) (b) identify (c)2identify2211 1Figure 10.27: Torus knot of type (2,3) (a) and its representation on the square (b).The knot produces frequency lock<strong>in</strong>g and phase lock<strong>in</strong>g. (c) Torus knot that does notproduce phase lock<strong>in</strong>g.that H(χ) = Q(χ) · A/T <strong>in</strong> the case of pulse-coupl<strong>in</strong>g (10.1), so that H(χ) is justre-scaled PRC.A special case of (10.15) is when H is replaced by its first Fourier term, s<strong>in</strong>. Theresult<strong>in</strong>g system written <strong>in</strong> the slow time τ = εtϕ ′ i = ω i +n∑c ij s<strong>in</strong>(ϕ j − ϕ i + ψ ij )j=1is called the Kuramoto phase model (Kuramoto 1975). Here, the frequency deviationsω i are <strong>in</strong>terpreted as <strong>in</strong>tr<strong>in</strong>sic frequencies of oscillators. The strengths of connectionsc ij are often assumed to be equal to K/n for some constant K, so that the model canbe studied <strong>in</strong> the limit n → ∞. The phase deviations ψ ij are often neglected for thesake of simplicity.To summarize, we transformed the weakly coupled system (10.11) <strong>in</strong>to the phasemodel (10.15) with H given by (10.16) and each Q be<strong>in</strong>g the solution to the adjo<strong>in</strong>tproblem (10.10). This constitutes the Malk<strong>in</strong> theorem for weakly coupled oscillators(Hoppensteadt and Izhikevich 1997, Theorem 9.2).10.3 SynchronizationConsider two coupled phase variables (10.12) <strong>in</strong> a general form˙ϑ 1 = h 1 (ϑ 1 , ϑ 2 ) ,˙ϑ 2 = h 2 (ϑ 1 , ϑ 2 ) ,with some positive functions h 1 and h 2 . S<strong>in</strong>ce each phase variable is def<strong>in</strong>ed on thecircle S 1 , the state space of this system is the 2-torus T 2 = S 1 ×S 1 depicted <strong>in</strong> Fig. 10.27,with ϑ 1 and ϑ 2 be<strong>in</strong>g the longitude and the latitude, respectively. The torus can berepresented as a square with vertical and horizontal sides identified, so that a solutiondisappear<strong>in</strong>g at the right side of the square appears at the left side.


482 Synchronization (see www.izhikevich.com)frequency lock<strong>in</strong>g<strong>in</strong>-phaseentra<strong>in</strong>ment(1:1 frequency lock<strong>in</strong>g)synchronizationphase lock<strong>in</strong>ganti-phaseFigure 10.28: Various degrees of lock<strong>in</strong>g of oscillators.The coupled oscillators above are said to be frequency locked when there is a periodictrajectory on the 2-torus, which is called a torus knot. It is said to be of type (p, q)if ϑ 1 makes p rotations while ϑ 2 makes q rotations, and p and q are relatively prime<strong>in</strong>tegers, i.e., do not have a common divisor greater than 1. Torus knots of type (p, q)produce p:q frequency lock<strong>in</strong>g, e.g., 2:3 frequency lock<strong>in</strong>g <strong>in</strong> Fig. 10.27. A 1:1 frequencylock<strong>in</strong>g is called entra<strong>in</strong>ment. There could be many periodic orbits on the torus, withstable orbits between unstable ones. S<strong>in</strong>ce the orbits on the 2-torus cannot <strong>in</strong>tersect,they all are knots of the same type, result<strong>in</strong>g <strong>in</strong> the same p:q frequency lock<strong>in</strong>g.Let us follow a trajectory on the torus and count the number of rotations of thephase variables. The limit of the ratio of rotations as t → ∞ is <strong>in</strong>dependent on thetrajectory we follow, and it is called the rotation number of the torus flow. It is rationalif and only if there is a (p, q) periodic orbit, <strong>in</strong> which case the rotation number is p/q.An irrational rotation number implies there are no periodic orbits, and it corresponds toa quasi-periodic or multifrequency torus flow. Oscillators exhibit phase drift<strong>in</strong>g <strong>in</strong> thiscase. Denjoy (1932) proved that such coupled oscillators are topologically equivalentto the uncoupled system ˙ϑ 1 = r, ˙ϑ2 = 1 with irrational r.Suppose the oscillators are frequency locked; that is, there is a p : q limit cycleattractor on the torus. We say that the oscillators are p : q phase locked ifqϑ 1 (t) − pϑ 2 (t) = conston the cycle. The value of the constant determ<strong>in</strong>es whether the lock<strong>in</strong>g is <strong>in</strong>-phase(const= 0), anti-phase (const= T/2, half-period), or out-of-phase. Frequency lock<strong>in</strong>gdoes not necessarily imply phase lock<strong>in</strong>g: The (2, 3) torus knot <strong>in</strong> Fig. 10.27b correspondsto phase lock<strong>in</strong>g, whereas that <strong>in</strong> Fig. 10.27c does not. Frequency lock<strong>in</strong>gwithout phase lock<strong>in</strong>g is called phase trapp<strong>in</strong>g. F<strong>in</strong>ally, synchronization is a 1:1 phaselock<strong>in</strong>g. The phase difference ϑ 2 − ϑ 1 is also called phase lag or phase lead. Therelationships between all these def<strong>in</strong>itions are shown <strong>in</strong> Fig. 10.28.Frequency lock<strong>in</strong>g, phase lock<strong>in</strong>g, entra<strong>in</strong>ment, and synchronization of a networkof n > 2 oscillators is the same as pair-wise lock<strong>in</strong>g, entra<strong>in</strong>ment, and synchronization


Synchronization (see www.izhikevich.com) 483Figure 10.29: A major part of computational neuroscience concerns coupled oscillators.of the oscillators compris<strong>in</strong>g the network. In addition, a network can exhibit partialsynchronization, when only a subset of oscillators is synchronized.Synchronization of oscillators with nearly identical frequencies is described by thephase model (10.15). Existence of one equilibrium of (10.15) implies the existence ofthe entire circular family of equilibria, s<strong>in</strong>ce translation of all ϕ i by a constant phaseshift does not change the phase differences ϕ j − ϕ i and hence the form of (10.15).This family corresponds to a limit cycle of (10.11), on which all oscillators, x i (t + ϕ i ),have equal frequencies and constant phase shifts, i.e., they are synchronized, possiblyout-of-phase.10.3.1 Two oscillatorsConsider (10.11) with n = 2, describ<strong>in</strong>g two coupled oscillators, as <strong>in</strong> Fig. 10.29. Letus <strong>in</strong>troduce the “slow” time τ = εt and rewrite the correspond<strong>in</strong>g phase model (10.15)<strong>in</strong> the formϕ ′ 1 = ω 1 + H 1 (ϕ 2 − ϕ 1 ) ,ϕ ′ 2 = ω 2 + H 2 (ϕ 1 − ϕ 2 ) ,where ′ = d/dτ is the derivative with respect to slow time. Let χ = ϕ 2 − ϕ 1 denotethe phase difference between the oscillators. Then the two-dimensional system abovebecomes one-dimensionalχ ′ = ω + G(χ) , (10.17)


484 Synchronization (see www.izhikevich.com)(a)max G0(b)max G0G( )m<strong>in</strong> GG( )0 1 2 3phase difference,T0 1 2 3phase difference,TFigure 10.30: Geometrical <strong>in</strong>terpretation of equilibria of the phase model (10.17) forgap-junction-coupled Class 2 I Na + I K -oscillators (see Fig. 10.26).whereω = ω 2 − ω 1 and G(χ) = H 2 (−χ) − H 1 (χ) ,is the frequency mismatch and the anti-symmetric part of the coupl<strong>in</strong>g, respectively,illustrated <strong>in</strong> Fig. 10.26, dashed curves. A stable equilibrium of (10.17) corresponds toa stable limit cycle of the phase model.All equilibria of (10.17) are solutions to G(χ) = −ω, and they are <strong>in</strong>tersections ofthe horizontal l<strong>in</strong>e −ω with the graph of G, as illustrated <strong>in</strong> Fig. 10.30a. They arestable if the slope of the graph is negative at the <strong>in</strong>tersection. If the oscillators areidentical, then G(χ) = H(−χ) − H(χ) is an odd function (i.e., G(−χ) = −G(χ)),and χ = 0 and χ = T/2 are always equilibria, possibly unstable, correspond<strong>in</strong>g to the<strong>in</strong>-phase and anti-phase synchronized solutions. The stability condition of the <strong>in</strong>-phasesynchronized state isG ′ (0) = −2H ′ (0) < 0(stability of <strong>in</strong>-phase synchronization)The <strong>in</strong>-phase synchronization of electrically (gap-junction) coupled oscillators <strong>in</strong> Fig. 10.26is stable because the slope of G (dashed curves) is negative at χ = 0. Simulation of twocoupled I Na + I K -oscillators <strong>in</strong> Fig. 10.25 confirms that. Coupled oscillators <strong>in</strong> Class 2regime also have a stable anti-phase solution, s<strong>in</strong>ce G ′ < 0 at χ = T/2 <strong>in</strong> Fig. 10.30a.The max and m<strong>in</strong> values of the function G determ<strong>in</strong>e the tolerance of the networkto the frequency mismatch ω, s<strong>in</strong>ce there are no equilibria outside this range. Geometrically,as ω <strong>in</strong>creases (the second oscillator speeds up), the horizontal l<strong>in</strong>e −ω <strong>in</strong>Fig. 10.30a slides down, and the phase difference χ = ϕ 2 − ϕ 1 <strong>in</strong>creases, compensat<strong>in</strong>gfor the frequency mismatch ω. When ω > − m<strong>in</strong> G, the second oscillator becomestoo fast, and the synchronized state is lost via saddle-node on <strong>in</strong>variant circle bifurcation;see Fig. 10.30b. This bifurcation corresponds to the annihilation of stable andunstable limit cycles of the weakly coupled network, and the result<strong>in</strong>g activity is calleddrift<strong>in</strong>g, cycle slipp<strong>in</strong>g, or phase walk-through. The variable χ slowly passes the ghostof the saddle-node po<strong>in</strong>t, where G(χ) ≈ 0, then <strong>in</strong>creases past T , appears at 0, andapproaches the ghost aga<strong>in</strong>, thereby slipp<strong>in</strong>g a cycle and walk<strong>in</strong>g through all the phasevalues [0, T ]. The frequency of such slipp<strong>in</strong>g scales as √ ω + m<strong>in</strong> G; see Sect. 6.1.2.In Fig. 10.31 we contrast synchronization properties of weakly coupled oscillatorsof relaxation and non-relaxation type. The function G(χ) of the former has a negative


Synchronization (see www.izhikevich.com) 485non-relaxation oscillatorG( )G( )relaxation oscillator000 phase difference, T0 phase difference, TFigure 10.31: Functions G(χ) for weakly coupled oscillators of non-relaxation (smooth)and relaxation type. The frequency mismatch ω creates a phase difference <strong>in</strong> thesmooth case, but not <strong>in</strong> the relaxation case.discont<strong>in</strong>uity at χ = 0; see Sect. 10.4.4 below. An immediate consequence is thatthe <strong>in</strong>-phase synchronization is rapid and persistent <strong>in</strong> the presence of the frequencymismatch ω. Indeed, if G is smooth, then χ slows down while it approaches theequilibrium χ = 0. As a result, complete synchronization is an asymptotic processthat requires an <strong>in</strong>f<strong>in</strong>ite period of time to atta<strong>in</strong>. In contrast, when G is discont<strong>in</strong>uousat 0, the variable χ does not slow down and it takes a f<strong>in</strong>ite period of time to lock.Chang<strong>in</strong>g the frequency mismatch ω shifts the root of −ω = G(χ) <strong>in</strong> the cont<strong>in</strong>uouscase, but not <strong>in</strong> the discont<strong>in</strong>uous case. Hence, the <strong>in</strong>-phase synchronized state χ = 0of coupled relaxation oscillators exists and it is stable <strong>in</strong> a wide range of ω.10.3.2 Cha<strong>in</strong>sUnderstand<strong>in</strong>g synchronization properties of two coupled oscillators helps one <strong>in</strong> study<strong>in</strong>gthe dynamics of cha<strong>in</strong>s of n > 2 oscillatorsϕ ′ i = ω i + H + (ϕ i+1 − ϕ i ) + H − (ϕ i−1 − ϕ i ) , (10.18)where the functions H + and H − describe the coupl<strong>in</strong>g <strong>in</strong> the ascend<strong>in</strong>g and descend<strong>in</strong>gdirections of the cha<strong>in</strong>, as <strong>in</strong> Fig. 10.32. Any phase-locked solution of (10.18) has theform ϕ i (τ) = ω 0 τ + φ i , where ω 0 is the common frequency of oscillation and φ i areconstants. These satisfy n conditionsω 0 = ω 1 + H + (φ 2 − φ 1 ) ,ω 0 = ω i + H + (φ i+1 − φ i ) + H − (φ i−1 − φ i ) , i = 2, . . . , n − 1 ,ω 0 = ω n + H − (φ n−1 − φ n ) .A solution with φ 1 < φ 2 < · · · < φ n or with φ 1 > φ 2 > · · · > φ n (as <strong>in</strong> Fig. 10.32)is called a travel<strong>in</strong>g wave. Indeed, the oscillators oscillate with a common frequencyω 0 but with different phases that <strong>in</strong>crease or decrease monotonically along the cha<strong>in</strong>.Such a behavior is believed to correspond to central pattern generation (CPG) <strong>in</strong>crayfish, undulatory locomotion <strong>in</strong> lamprey and dogfish, and peristalsis <strong>in</strong> vascular and<strong>in</strong>test<strong>in</strong>al smooth muscles. Below we consider two fundamentally different mechanismsof generation of travel<strong>in</strong>g waves.


486 Synchronization (see www.izhikevich.com)H - H - H - H - H -H+ H+ H+ H+ H+ H+ H+nH - H -travel<strong>in</strong>g waveFigure 10.32: Travel<strong>in</strong>g wave solutions <strong>in</strong> cha<strong>in</strong>s of oscillators (10.18) describe undulatorylocomotion and central pattern generation.Frequency differencesSuppose the connections <strong>in</strong> (10.18) look qualitatively similar to those <strong>in</strong> Fig. 10.26, <strong>in</strong>particular, H + (0) = H − (0) = 0. If the frequencies are all equal, then the <strong>in</strong>-phasesynchronized solution ϕ 1 = · · · = ϕ n exists and is stable. A travel<strong>in</strong>g wave exists whenthe frequencies are not all equal.Let us seek the conditions for the existence of a travel<strong>in</strong>g wave with a constantphase shift, say χ = φ i+1 − φ i , along the cha<strong>in</strong>. Subtract<strong>in</strong>g each equation from thesecond one, we f<strong>in</strong>d that0 = ω 2 − ω 1 + H − (−χ) , 0 = ω 2 − ω i , 0 = ω 2 − ω n + H + (χ) ,and ω 0 = ω 1 +ω n −2ω 2 . In particular, if ω 1 ≤ ω 2 = · · · = ω n−1 ≤ ω n , which correspondsto the first oscillator be<strong>in</strong>g tuned up and the last oscillator be<strong>in</strong>g tuned down, thenχ < 0 and the travel<strong>in</strong>g wave moves up, as <strong>in</strong> Fig. 10.32, i.e., from the fastest to theslowest oscillator. Interest<strong>in</strong>gly, such an ascend<strong>in</strong>g wave exists even when H − = 0, i.e.,even when the coupl<strong>in</strong>g is only <strong>in</strong> the opposite, descend<strong>in</strong>g direction.When there is a l<strong>in</strong>ear gradient of frequencies (ω 1 > ω 2 > · · · > ω n or vice versa),as <strong>in</strong> the cases of the smooth muscle of <strong>in</strong>test<strong>in</strong>es or leech CPG for swimm<strong>in</strong>g, onemay still observe a travel<strong>in</strong>g wave but with a non-constant phase difference along thecha<strong>in</strong>. When the gradient is large enough, the synchronized solution correspond<strong>in</strong>g toa s<strong>in</strong>gle travel<strong>in</strong>g wave disappears, and frequency plateaus may appear (Ermentroutand Kopell 1984). That is, solutions occur <strong>in</strong> which the first k < n oscillators arephase locked and the last n − k oscillators are phase locked as well, but the two poolsoscillate with different frequencies. There may be many frequency plateaus.Coupl<strong>in</strong>g functionsA travel<strong>in</strong>g wave solution may exist even when all the frequencies are equal, if eitherH + (0) ≠ 0 or H − (0) ≠ 0. As an example, consider the case of descend<strong>in</strong>g coupl<strong>in</strong>g(H − = 0)ϕ ′ i = ω + H + (ϕ i+1 − ϕ i ) , i = 1, . . . , n − 1 .


Synchronization (see www.izhikevich.com) 487From ϕ ′ n = ω we f<strong>in</strong>d that ω 0 = ω, i.e., the common frequency is the frequency ofthe free oscillation of the last, uncoupled oscillator. The phase lag along the cha<strong>in</strong>,χ = ϕ i+1 − ϕ i , satisfies n − 1 identical conditions 0 = H + (χ). Thus, the travel<strong>in</strong>g wavewith a constant phase shift exists when H + has a zero cross<strong>in</strong>g with positive slope, <strong>in</strong>contrast to Fig. 10.26. The sign of χ, and not the direction of coupl<strong>in</strong>g, determ<strong>in</strong>es thedirection of wave propagation.10.3.3 NetworksNow let us consider weakly connected networks (10.11) with arbitrary, possibly allto all coupl<strong>in</strong>g. To study synchronized states of the network, we need to determ<strong>in</strong>ewhether the correspond<strong>in</strong>g phase model (10.15) has equilibria and exam<strong>in</strong>e their stabilityproperties. A vector φ = (φ 1 , . . . , φ n ) is an equilibrium of (10.15) whenn∑0 = ω i + H ij (φ j − φ i ) for all i . (10.19)j≠1It is stable when all eigenvalues of the l<strong>in</strong>earization matrix (Jacobian) at φ have negativereal parts, except one zero eigenvalue correspond<strong>in</strong>g to the eigenvector along thecircular family of equilibria (φ plus a phase shift is a solution of (10.19) too s<strong>in</strong>ce thephase differences φ j − φ i are not affected).In general, determ<strong>in</strong><strong>in</strong>g the stability of equilibria is a difficult problem. Ermentrout(1992) found a simple sufficient condition. Namely, if• a ij = H ′ ij(φ j − φ i ) ≥ 0, and• the directed graph def<strong>in</strong>ed by the matrix a = (a ij ) is connected, (i.e., each oscillatoris <strong>in</strong>fluenced, possibly <strong>in</strong>directly, by every other oscillator),then the equilibrium φ is neutrally stable, and the correspond<strong>in</strong>g limit cycle x(t + φ)of (10.11) is asymptotically stable.Another sufficient condition was found by Hoppensteadt and Izhikevich (1997). Itstates that if system (10.15) satisfies• ω 1 = · · · = ω n = ω(identical frequencies), and• H ij (−χ) = −H ji (χ) (pair-wise odd coupl<strong>in</strong>g)for all i and j, then the network dynamics converge to a limit cycle. On the cycle, alloscillators have equal frequencies 1 + εω and constant phase deviations.The proof follows from the observation that (10.15) is a gradient system <strong>in</strong> therotat<strong>in</strong>g coord<strong>in</strong>ates ϕ = ωτ + φ, with the energy functionE(φ) = 1 n∑ n∑∫ χR ij (φ j − φ i ) , where R ij (χ) = H ij (s) ds .2i=1j=1One can check that dE(φ)/dτ = − ∑ (φ ′ i) 2 ≤ 0 along the trajectories of (10.15), withequality only at equilibria.0


488 Synchronization (see www.izhikevich.com)rre iFigure 10.33: The Kuramoto synchronization <strong>in</strong>dex(10.21) describes the degree of coherence <strong>in</strong>the network (10.20).10.3.4 Mean-field approximationsSynchronization of the phase model (10.15) with randomly distributed frequency deviationsω i can be analyzed <strong>in</strong> the limit n → ∞, often called thermodynamic limit byphysicists. We illustrate the theory us<strong>in</strong>g the special case, H(χ) = s<strong>in</strong> χ (Kuramoto1975)ϕ ′ i = ω i + K n∑s<strong>in</strong>(ϕ j − ϕ i ) , ϕ i ∈ [0, 2π] , (10.20)nj=1where K > 0 is the coupl<strong>in</strong>g strength and the factor 1/n ensures that the modelbehaves well as n → ∞. The complex-valued sum of all phases,re iψ = 1 nn∑e iϕ j(Kuramoto synchronization <strong>in</strong>dex), (10.21)j=1describes the degree of synchronization <strong>in</strong> the network. The parameter r is oftencalled order parameter by physicists. Apparently, the <strong>in</strong>-phase synchronized stateϕ 1 = · · · = ϕ n corresponds to r = 1, with ψ be<strong>in</strong>g the population phase. In contrast,the <strong>in</strong>coherent state with all ϕ i hav<strong>in</strong>g different values randomly distributed on theunit circle, corresponds to r ≈ 0. (The case r ≈ 0 can also correspond to two or moreclusters of synchronized neuron, oscillat<strong>in</strong>g anti-phase or out-of-phase and cancel<strong>in</strong>geach other). Intermediate values of r correspond to a partially synchronized or coherentstate, depicted <strong>in</strong> Fig. 10.33. Some phases are synchronized form<strong>in</strong>g a cluster, whileothers roam around the circle.Multiply<strong>in</strong>g both sides of (10.21) by e −iϕ iand consider<strong>in</strong>g only the imag<strong>in</strong>ary parts,we can rewrite (10.20) <strong>in</strong> the equivalent formϕ ′ i = ω i + Kr s<strong>in</strong>(ψ − ϕ i ) ,which emphasizes the mean-filed character of <strong>in</strong>teractions between the oscillators: Theyare all pulled <strong>in</strong>to the synchronized cluster (ϕ i → ψ) with the effective strength proportionalto the cluster size r. This pull is offset by the random frequency deviationsω i , which pull away from the cluster.


Synchronization (see www.izhikevich.com) 489Let us assume that the frequencies ω i are distributed randomly around 0 with asymmetric probability density function g(ω), e.g., Gaussian. Kuramoto (1975) hasshown that <strong>in</strong> the limit n → ∞, the cluster size r obeys the self-consistency equationr = rK∫ +π/2−π/2g(rK s<strong>in</strong> ϕ) cos 2 ϕ dϕ (10.22)derived <strong>in</strong> Ex. 21. Notice that r = 0, correspond<strong>in</strong>g to the <strong>in</strong>coherent state, is alwaysa solution of this equation. When the coupl<strong>in</strong>g strength K is greater than a certa<strong>in</strong>critical value,K c = 2πg(0) ,an additional, nontrivial solution r > 0 appears, which corresponds to a partiallysynchronized state. It scales as r = √ 16(K − K c )/(−g ′′ (0)πK 4 c ), as the reader canprove himself by expand<strong>in</strong>g g <strong>in</strong> a Taylor series. Thus, the stronger the coupl<strong>in</strong>g Krelative to the random distribution of frequencies, the more oscillators synchronize <strong>in</strong>toa coherent cluster. The issue of stability of <strong>in</strong>coherent and partially synchronized statesis discussed by Strogatz (2000).10.4 ExamplesBelow we consider simple examples of oscillators to illustrate the theory developed <strong>in</strong>this chapter. Our goal is to understand which details of oscillators are important <strong>in</strong>shap<strong>in</strong>g the PRC, the form of the function H <strong>in</strong> the phase deviation model, and hencethe existence and stability of synchronized states.10.4.1 Phase oscillatorsLet us consider the simplest possible k<strong>in</strong>d of a non-l<strong>in</strong>ear oscillator, known as the phaseoscillator:ẋ = f(x) + εp(t) , x ∈ S 1 , (10.23)where f(x) > 0 is a periodic function, for example, f(x) = a + s<strong>in</strong> x with a > 1.Notice that this k<strong>in</strong>d of oscillator is quite different from the two- or high-dimensionalconductance-based models with limit cycle attractors that we considered <strong>in</strong> the previouschapters. Here, the state variable x is one-dimensional def<strong>in</strong>ed on a circle S 1 , that is,it may be <strong>in</strong>terpreted as a measure of distance along a limit cycle attractor of a multidimensionalsystem.Consider the unperturbed (ε = 0) phase oscillator ẋ = f(x), and let x(t) be itssolution with some period T > 0. Follow<strong>in</strong>g Kuramoto’s idea, we substitute x(ϑ) <strong>in</strong>to(10.23) and use the cha<strong>in</strong> rule,f(x(ϑ)) + εp(t) = {x(ϑ)} ′ = x ′ (ϑ) ϑ ′ = f(x(ϑ))ϑ ′ ,


490 Synchronization (see www.izhikevich.com)to get the new phase equation˙ϑ = 1 + εp(t)/f(x(ϑ)) , (10.24)which is equivalent to (10.23) for any, not necessarily small, ε.We can also get (10.24) us<strong>in</strong>g any of the three methods of reduction of oscillatorsto phase models:• Malk<strong>in</strong>’s method is the easiest one: We do not even have to solve the onedimensionaladjo<strong>in</strong>t equation (10.10) hav<strong>in</strong>g the form ˙Q = −f ′ (x(t)) Q, becausewe can get the solution Q(t) = 1/f(x(t)) directly from the normalization conditionQ(t)f(x(t)) = 1.• Kuramoto’s method relies on the function ϑ(x), which we can f<strong>in</strong>d implicitly.S<strong>in</strong>ce the phase at a po<strong>in</strong>t x(t) on the limit cycle is just t, x(ϑ) is the <strong>in</strong>verse ofϑ(x). Us<strong>in</strong>g the rule for differentiat<strong>in</strong>g of <strong>in</strong>verse functions, ϑ ′ (x) = 1/x ′ (ϑ), wef<strong>in</strong>d grad ϑ = 1/f(x(ϑ)).• W<strong>in</strong>free’s method relies on PRC (ϑ), which we f<strong>in</strong>d us<strong>in</strong>g the follow<strong>in</strong>g procedure:A pulsed perturbation at phase ϑ moves the solution from x(ϑ) to x(ϑ) + A,which is x(ϑ + PRC (ϑ, A)) ≈ x(ϑ) + x ′ (ϑ)PRC (ϑ, A) when A is small. Hence,PRC (ϑ, A) ≈ A/x ′ (ϑ) = A/f(x(ϑ)), and the l<strong>in</strong>ear response is Z(ϑ) = 1/f(x(ϑ))when A → 0.Two coupled identical oscillatorsẋ 1 = f(x 1 ) + εg(x 2 )ẋ 2 = f(x 2 ) + εg(x 1 )can be reduced to the phase model (10.17) with G(χ) = H(−χ) − H(χ), whereH(χ) = 1 T∫ T0Q(t) g(x(t + χ)) dt = 1 T∫ T0g(x(t + χ))f(x(t))dtThe condition for exponential stability of the <strong>in</strong>-phase synchronized state, χ = 0, canbe expressed <strong>in</strong> the follow<strong>in</strong>g three equivalent forms∫ T0g ′ (x(t)) dt > 0as we prove <strong>in</strong> Ex. 24.or10.4.2 SNIC oscillators∫S 1 g ′ (x)f(x) dx > 0 or ∫S 1f ′ (x)g(x) dx > 0 , (10.25)f 2 (x)Let us go through all the steps of derivation of the phase equation us<strong>in</strong>g a neuronmodel exhibit<strong>in</strong>g low-frequency periodic spik<strong>in</strong>g. Such a model is near the saddle-node


Synchronization (see www.izhikevich.com) 491- +spike- + - +spike spikex(t)spikex(t)=-cot tspike-1 1000x'=-1+x2 x'=0+x2 x'=1+x2Figure 10.34: Phase portraits and typical oscillations of the quadratic <strong>in</strong>tegrate-and-fireneuron ẋ = I + x 2 with x ∈ R ∪ {±∞}. Parameter: I = −1, 0, +1.x'=1+x2A0xnormalized PRC (PRC( ,A)/A)100A=1A=2A=3A=0.1phase of oscillation,TFigure 10.35: The dependence of PRC of the quadratic <strong>in</strong>tegrate-and-fire model on thestrength of the pulse A.on <strong>in</strong>variant circle (SNIC) bifurcation studied <strong>in</strong> Sect. 6.1.2. Appropriate rescal<strong>in</strong>g ofthe membrane potential and time converts the model <strong>in</strong>to the normal formx ′ = 1 + x 2 , x ∈ R .Because of the quadratic term, x escapes to the <strong>in</strong>f<strong>in</strong>ity <strong>in</strong> a f<strong>in</strong>ite time, produc<strong>in</strong>ga spike depicted <strong>in</strong> Fig. 10.34. If we identify −∞ and +∞, then x exhibits periodicspik<strong>in</strong>g of <strong>in</strong>f<strong>in</strong>ite amplitude. Such a spik<strong>in</strong>g model is called quadratic <strong>in</strong>tegrate-and-fire(QIF) neuron; see also Sect. 8.1.3 for some generalizations of the model.Strong pulseThe solution of this system start<strong>in</strong>g at the spike, i.e., at x(0) = ±∞, isx(t) = − cot t ,as the reader can check by differentiat<strong>in</strong>g. It is a periodic function with T = π,hence, we can <strong>in</strong>troduce the phase of oscillation via the relation x = − cot ϑ. Thecorrespond<strong>in</strong>g PRC can be found explicitly (see Ex. 9) and it has the formPRC (ϑ, A) = π/2 + atan (A − cot ϑ) − ϑ ,


492 Synchronization (see www.izhikevich.com)depicted <strong>in</strong> Fig. 10.35, where A is the magnitude of the pulse. Notice that PRC tiltsto the left as A <strong>in</strong>creases. Indeed, the density of isochrons, denoted by black po<strong>in</strong>ts onthe x-axis <strong>in</strong> the figure, is maximal at the ghost of the saddle-node po<strong>in</strong>t x = 0 wherethe parabola 1 + x 2 has the knee. This corresponds to the <strong>in</strong>flection po<strong>in</strong>t of the graphof x(t) <strong>in</strong> Fig. 10.34 where the dynamics of x(t) is the slowest. The effect of a pulseis maximal just before the ghost because x can jump over the ghost and skip the slowregion. The stronger the pulse, the earlier it should arrive, hence the tilt.Weak coupl<strong>in</strong>gThe PRC behaves as A s<strong>in</strong> 2 ϑ, with ϑ ∈ [0, π], when A is small, as the reader can see<strong>in</strong> Fig. 10.35 or prove himself by differentiat<strong>in</strong>g the function PRC (ϑ, A) with respectto A. Therefore, Z(ϑ) = s<strong>in</strong> 2 ϑ, and we can use W<strong>in</strong>free’a approach to transform theweakly perturbed quadratic <strong>in</strong>tegrate-and-fire (QIF) oscillator<strong>in</strong>to its phase modelx ′ = 1 + x 2 + εp(t)x ′ = 1 + ε(s<strong>in</strong> 2 ϑ)p(t) , ϑ ∈ [0, π] .The results of the previous section, Q(ϑ) = 1/f(x(ϑ)) = 1/(1+cot 2 ϑ) = s<strong>in</strong> 2 ϑ, confirmthe phase model. In fact, any neuronal model C ˙V = I − I ∞ (V ) near saddle-node on<strong>in</strong>variant circle bifurcation po<strong>in</strong>t (I sn , V sn ) has <strong>in</strong>f<strong>in</strong>itesimal PRCPRC (ϑ) =CI − I sns<strong>in</strong> 2 ϑ , ϑ ∈ [0, π] ,as the reader can prove as an exercise. The function s<strong>in</strong> 2 ϑ co<strong>in</strong>cides with the familiar1 − cos θ when θ = 2ϑ has period 2π (notice the font difference).In Fig. 10.36a we compare the function with numerically obta<strong>in</strong>ed PRCs for theI Na + I K -model <strong>in</strong> Class 1 regime. S<strong>in</strong>ce the ghost of the saddle-node po<strong>in</strong>t, reveal<strong>in</strong>gitself as an <strong>in</strong>flection of the voltage trace <strong>in</strong> Fig. 10.36b, moves to the right as I <strong>in</strong>creasesaway from the bifurcation value I = 4.51, so does the peak of the PRC.Figure 10.36a emphasizes the common features of all systems undergo<strong>in</strong>g saddlenodeon <strong>in</strong>variant circle bifurcation: They are <strong>in</strong>sensitive to the <strong>in</strong>puts arriv<strong>in</strong>g dur<strong>in</strong>gthe spike, s<strong>in</strong>ce PRC ≈ 0 when ϑ ≈ 0, T . The oscillators are most sensitive to the <strong>in</strong>putwhen they are just enter<strong>in</strong>g the ghost of the rest<strong>in</strong>g state, where PRC is maximal.The location of the maximum tilts to the left as the strength of the <strong>in</strong>put <strong>in</strong>creases,and may tilt to the right as the distance to the bifurcation <strong>in</strong>creases. F<strong>in</strong>ally, PRCsare non-negative, so positive (negative) <strong>in</strong>puts can only advance (delay) the phase ofoscillation.


Synchronization (see www.izhikevich.com) 493(a)(b)20normalized PRC (Q1( ))10.80.60.40.20s<strong>in</strong> 2I=4.52I=4.55I=4.6I=4.7I=5I=10membrane potential, (mV)0-20-40-60<strong>in</strong>flectionpo<strong>in</strong>ts0phase of oscillation,T-800phase of oscillation,TFigure 10.36: (a) Numerically found PRCs of the I Na + I K -oscillator <strong>in</strong> Class 1 regime(parameters as <strong>in</strong> Fig. 4.1a) and various I us<strong>in</strong>g the MATLAB program <strong>in</strong> Ex. 12.(b) Correspond<strong>in</strong>g voltage traces show that the <strong>in</strong>flection po<strong>in</strong>t (slowest <strong>in</strong>crease) of Vmoves right as I <strong>in</strong>creases.Gap junctionsNow consider two oscillators coupled via gap junctions, discussed <strong>in</strong> Sect. 2.3.4,x ′ 1 = 1 + x 2 1 + ε(x 2 − x 1 ) ,x ′ 2 = 1 + x 2 2 + ε(x 1 − x 2 ) .Let us determ<strong>in</strong>e the stability of the <strong>in</strong>-phase synchronized state x 1 = x 2 . The correspond<strong>in</strong>gphase model (10.12) has the formϑ ′ 1 = 1 + ε(s<strong>in</strong> 2 ϑ 1 )(cot ϑ 1 − cot ϑ 2 ) ,ϑ ′ 2 = 1 + ε(s<strong>in</strong> 2 ϑ 2 )(cot ϑ 2 − cot ϑ 1 ) .The function (10.16) can be found analytically:H(χ) = 1 π∫ π0s<strong>in</strong> 2 t (cot t − cot(t + χ)) dt = 1 s<strong>in</strong> 2χ ,2so that the model <strong>in</strong> the phase deviation coord<strong>in</strong>ates, ϑ(t) = t + ϕ, has the formϕ ′ 1 = (ε/2) s<strong>in</strong>{2(ϕ 2 − ϕ 1 )} ,ϕ ′ 2 = (ε/2) s<strong>in</strong>{2(ϕ 1 − ϕ 2 )} .The phase difference, χ = ϕ 2 − ϕ 1 , satisfies the equation (compare with Fig. 10.26)χ ′ = −ε s<strong>in</strong> 2χ ,and, apparently, the <strong>in</strong>-phase synchronized state, χ = 0, is always stable while theanti-phase state χ = π/2 is not.


494 Synchronization (see www.izhikevich.com)Weak pulsesNow consider two weakly pulse-coupled oscillatorsx ′ 1 = 1 + x 2 1 + ε 1 δ(t − t 2 ) ,x ′ 2 = 1 + x 2 2 + ε 2 δ(t − t 1 ) ,where t 1 and t 2 are the moments of fir<strong>in</strong>g (x(t) = ∞) of the first and the secondoscillator, respectively, and ε 1 and ε 2 are the strengths of synaptic connections. Thecorrespond<strong>in</strong>g phase model (10.12) has the formS<strong>in</strong>ceH(χ) = 1 πϑ ′ 1 = 1 + ε 1 (s<strong>in</strong> 2 ϑ 1 )δ(t − t 2 )ϑ ′ 2 = 1 + ε 2 (s<strong>in</strong> 2 ϑ 2 )δ(t − t 1 ) .∫ π0s<strong>in</strong> 2 t δ(t + χ) dt = 1 π s<strong>in</strong>2 χ ,the correspond<strong>in</strong>g phase deviation model (10.15) isϕ ′ 1 = ε 1π s<strong>in</strong>2 (ϕ 2 − ϕ 1 ) ,ϕ ′ 2 = ε 2π s<strong>in</strong>2 (ϕ 1 − ϕ 2 ) .The phase difference χ = ϕ 2 − ϕ 1 satisfies the equationχ ′ = ε 2 − ε 1πs<strong>in</strong> 2 χ,which becomes χ ′ = 0 when the coupl<strong>in</strong>g is symmetric. In this case, the oscillatorspreserve (on average) the <strong>in</strong>itial phase difference. When ε 1 ≠ ε 2 , the <strong>in</strong>-phase synchronizedstate χ = 0 is only neutrally stable. Interest<strong>in</strong>gly, it becomes exponentiallyunstable <strong>in</strong> a network of three or more pulse-coupled Class 1 oscillators; see Ex. 23.Weak pulses with delaysThe synchronization properties of weakly pulse-coupled oscillators could change significantlywhen explicit axonal conduction delays are <strong>in</strong>troduced. As an example, considerthe systemx ′ 1 = 1 + x 2 1 + εδ(t − t 2 − d) ,x ′ 2 = 1 + x 2 2 + εδ(t − t 1 − d) ,where d ≥ 0 is the delay. Ex. 18 shows that delays <strong>in</strong>troduce simple phase shifts, sothat the phase model has the formϕ ′ 1 = ε π s<strong>in</strong>2 (ϕ 2 − ϕ 1 − d) ,ϕ ′ 2 = ε π s<strong>in</strong>2 (ϕ 1 − ϕ 2 − d) ,


Synchronization (see www.izhikevich.com) 4951p(t)s<strong>in</strong>2ts<strong>in</strong> 2 t00phase of oscillationFigure 10.37: Synaptic transmission function p(t) typically has an asymmetric shapewith fast rise and slow decay.The phase difference χ = ϕ 2 − ϕ 1 satisfiesχ ′ = ε π(s<strong>in</strong> 2 (χ + d) − s<strong>in</strong> 2 (χ − d) ) . =ε s<strong>in</strong> 2dπs<strong>in</strong> 2χ .The stability of synchronized states is determ<strong>in</strong>ed by the sign of the function s<strong>in</strong> 2d.The <strong>in</strong>-phase state χ = 0 is unstable when s<strong>in</strong> 2d > 0, i.e., when the delay is shorterthan the half-period π/2, stable when the delay is longer than half-period but shorterthan one period π, unstable for even longer delays, etc. The stability of the anti-phasestate χ = π/2 is reversed, i.e., it is stable for short delays, unstable for longer delays,then stable aga<strong>in</strong> for even longer delays, etc. F<strong>in</strong>ally, when the pulses are <strong>in</strong>hibitory(ε < 0), the (<strong>in</strong>)stability character is flipped so that the <strong>in</strong>-phase state becomes stablefor short delays.Weak synapsesNow suppose that each pulse is not a delta function, but it is smeared <strong>in</strong> time, i.e., ithas a time course p(t − t i ) with p(0) = p(π) = 0. That is, the synaptic transmissionstarts right after the spike of the pre-synaptic neuron and ends before the onset ofthe next spike. The function p has a typical unimodal shape with fast rise and slowdecay depicted <strong>in</strong> Fig. 10.37. The discussion below is equally applicable to the caseof p(t, x) = g(t)(E − x) with g > 0 be<strong>in</strong>g the synaptic conductance with the shape <strong>in</strong>the figure and E be<strong>in</strong>g the synaptic reverse potential, positive (negative) for excitatory(<strong>in</strong>hibitory) synapses.Two weakly synaptically coupled SNIC (Class 1) oscillatorsx ′ 1 = 1 + x 2 1 + εp(t − t 2 ) ,x ′ 2 = 1 + x 2 2 + εp(t − t 1 )can be converted <strong>in</strong>to a general phase model with the connection function (10.16) <strong>in</strong>the formH(χ) = 1 π∫ π0s<strong>in</strong> 2 t p(t + χ) dt ,


496 Synchronization (see www.izhikevich.com)nI Na +I K -modelV(t)Vquadratic <strong>in</strong>tegrate-and-fire modelx(t)=-coth 2 (t-T)resetspikex reset-1 0 +1xresetresetresetFigure 10.38: Top: Periodic spik<strong>in</strong>g of the I Na +I K -neuron near saddle-node homocl<strong>in</strong>icorbit bifurcation (parameters as <strong>in</strong> Fig. 4.1a with τ(V ) = 0.167 and I = 4.49). Bottom:Spik<strong>in</strong>g <strong>in</strong> the correspond<strong>in</strong>g quadratic <strong>in</strong>tegrate-and-fire model.and it can be computed explicitly for some simple p(t).The <strong>in</strong>-phase synchronized solution, χ = 0, is stable whenH ′ (0) = 1 π∫ π0s<strong>in</strong> 2 t p ′ (t) dt > 0 .S<strong>in</strong>ce the function s<strong>in</strong> 2 t depicted <strong>in</strong> Fig. 10.37 is small at the ends of the <strong>in</strong>terval andlarge <strong>in</strong> the middle, the <strong>in</strong>tegral is dom<strong>in</strong>ated by the sign of p ′ <strong>in</strong> the middle. Fast ris<strong>in</strong>gand slowly decay<strong>in</strong>g excitatory (p > 0) synaptic transmission has p ′ < 0 <strong>in</strong> the middle,as <strong>in</strong> the figure, so the <strong>in</strong>tegral is negative and the <strong>in</strong>-phase solution is unstable. Incontrast, fast ris<strong>in</strong>g slowly decay<strong>in</strong>g <strong>in</strong>hibitory (p < 0) synaptic transmission has p ′ > 0<strong>in</strong> the middle, so the <strong>in</strong>tegral is positive and the <strong>in</strong>-phase solution is stable. Anotherway to see this is to <strong>in</strong>tegrate the equation by parts, reduce it to − ∫ p(t) s<strong>in</strong> 2t dt,and notice that p(t) is concentrated <strong>in</strong> the first (left) half of the period where s<strong>in</strong> 2tis positive. Hence, positive (excitatory) p results <strong>in</strong> H ′ (0) < 0, and vice versa. Bothapproaches confirm the theoretical results <strong>in</strong>dependently obta<strong>in</strong>ed by van Vreeswijk etal. (1994) and Hansel et al. (1995) that <strong>in</strong>hibition, not excitation synchronizes Class1 (SNIC) oscillators. The relationship is <strong>in</strong>verted for the anti-phase solution χ = π/2(prove it), and no relationships is known for other types of oscillators.10.4.3 Homocl<strong>in</strong>ic oscillatorsBesides the SNIC bifurcation considered above, low frequency oscillations may also be<strong>in</strong>dicative of the proximity of the system to a saddle homocl<strong>in</strong>ic orbit bifurcation, as <strong>in</strong>


Synchronization (see www.izhikevich.com) 497(a)normalized PRC (Q1( ))0spikedownstroke00s<strong>in</strong>h 2 ( -T)-n'( )s<strong>in</strong>h 2 ( -T)numerical0.1T0 0.1T phase of oscillation,T(b)nBphase advancephase delayAVFigure 10.39: (a) Numerically found PRCs of the I Na + I K -oscillator near saddle-nodehomocl<strong>in</strong>ic orbit bifurcation (as <strong>in</strong> Fig. 10.38) us<strong>in</strong>g the MATLAB program <strong>in</strong> Ex. 12.Magnification shows the divergence from the theoretical curve s<strong>in</strong>h 2 (ϑ − T ) dur<strong>in</strong>g thespike. (b) A pulsed <strong>in</strong>put dur<strong>in</strong>g the downstroke of the spike can produce a significantphase delay (pulse A) or advance (pulse B) not captured by the quadratic <strong>in</strong>tegrateand-firemodel.Fig. 10.38, top. The spik<strong>in</strong>g trajectory <strong>in</strong> the figure quickly approaches a small shadedneighborhood of the saddle along the stable direction, and then slowly diverges fromthe saddle along the unstable direction, thereby result<strong>in</strong>g <strong>in</strong> a large period oscillation.As it is often the case <strong>in</strong> neuronal models, the saddle equilibrium is near a stable nodeequilibrium correspond<strong>in</strong>g to the rest<strong>in</strong>g state, and the system is near the co-dimension2 saddle-node homocl<strong>in</strong>ic orbit bifurcation studied <strong>in</strong> Sect. 6.3.6. As a result, thereis a drastic difference between the attraction and divergence rates to the saddle, sothat the dynamics <strong>in</strong> the shaded neighborhood of the saddle-node <strong>in</strong> the figure can bereduced to the one-dimensional V -equation, which <strong>in</strong> return can be transformed <strong>in</strong>tothe “quadratic <strong>in</strong>tegrate-and-fire” formx ′ = −1 + x 2 , if x = +∞, then x ← x reset ,with solutions depicted <strong>in</strong> Fig. 10.38, bottom. The saddle and the node correspond tox = ±1, respectively. One can check by differentiat<strong>in</strong>g that the solution of the modelwith x(0) = x reset > 1, is x(t) = − coth(t − T ), where coth(s) = (e s + e −s )/(e s − e −s )is the hyperbolic cotangent, and T = acoth (x reset ) is the period of oscillation, whichbecomes <strong>in</strong>f<strong>in</strong>ite as x reset → 1.Us<strong>in</strong>g the results of Sect. 10.4.1, we f<strong>in</strong>d the functionQ(ϑ) = 1/(−1 + coth 2 (ϑ − T )) = s<strong>in</strong>h 2 (ϑ − T )whose graph is shown <strong>in</strong> Fig. 10.39a. For comparison, we plotted the numerically foundPRC for the I Na + I K -oscillator to illustrate the disagreement between the theoreticaland numerical curves <strong>in</strong> the region ϑ < 0.1T correspond<strong>in</strong>g to the downstroke ofthe spike. Such a disagreement is somewhat expected, s<strong>in</strong>ce the quadratic <strong>in</strong>tegrateand-firemodel ignores spike downstroke. If a pulse arriv<strong>in</strong>g dur<strong>in</strong>g the downstroke


498 Synchronization (see www.izhikevich.com)displaces the trajectory to the exterior of the limit cycle, as <strong>in</strong> Fig. 10.39b, pulse A,then the trajectory becomes closer to the saddle equilibrium when it reenters a smallneighborhood of the saddle, thereby lead<strong>in</strong>g to a significant phase delay. Displacementsto the <strong>in</strong>terior of the cycle, as <strong>in</strong> Fig. 10.39b, pulse B, push away from the saddleand lead to phase advances. The direction and the magnitude of displacements isdeterm<strong>in</strong>ed by the derivative of the slow variable n ′ along the limit cycle.The region of disagreement between theoretical and numerical PRCs becomes <strong>in</strong>f<strong>in</strong>itesimalrelative to T → ∞ near the bifurcation. Theoretical PRC can be used tostudy anti-phase and out-of-phase synchronization of pulse-coupled oscillators, but not<strong>in</strong>-phase synchronization, because the region of breakdown is the only important region<strong>in</strong> this case. F<strong>in</strong>ally notice that as T → ∞, the spik<strong>in</strong>g limit cycle fails to be exponentiallystable, and the theory of weakly coupled oscillators is no longer applicable toit.Though the PRC <strong>in</strong> Fig. 10.39 is quite different from the one correspond<strong>in</strong>g to SNICoscillators <strong>in</strong> Fig. 10.36, there is an <strong>in</strong>terest<strong>in</strong>g similarity between these two cases: Bothcan be reduced to quadratic <strong>in</strong>tegrate-and-fire neurons, both have cotangent-shapedperiodic spik<strong>in</strong>g solutions and s<strong>in</strong>e-squared-shape PRCs, except they are “regular” <strong>in</strong>the SNIC case and hyperbolic <strong>in</strong> the homocl<strong>in</strong>ic case; see also Ex. 26.10.4.4 Relaxation oscillators and FTMConsider two relaxation oscillators hav<strong>in</strong>g weak fast → fast connectionsµẋ i = f(x i , y i ) + εp i (x i , x k ) ,ẏ i = g(x i , y i ) ,(10.26)where i = 1, 2 and k = 2, 1. This system can be converted to a phase model <strong>in</strong>the relaxation limit ε ≪ µ → 0 (Izhikevich 2000). The connection functions H i (χ)have a positive discont<strong>in</strong>uity at χ = 0, which occurs because the x-coord<strong>in</strong>ate of therelaxation limit cycle is discont<strong>in</strong>uous at the jump po<strong>in</strong>ts. Hence, the phase differencefunction G(χ) = H 2 (−χ) − H 1 (χ) has a negative discont<strong>in</strong>uity at χ = 0 depicted <strong>in</strong>Fig. 10.31. This reflects the profound difference between behaviors of weakly coupledoscillators of relaxation and non-relaxation type, discussed <strong>in</strong> Sect. 10.3.1: The <strong>in</strong>phasesynchronized solution, χ = 0, <strong>in</strong> the relaxation limit µ → 0 is stable, persistent<strong>in</strong> the presence of frequency mismatch ω, and it has a rapid rate of convergence.The reduction to a phase model breaks down when ε ≫ µ → 0, that is, when theconnections are relatively strong. One can still analyze such oscillators <strong>in</strong> the specialcase considered below.Fast threshold modulationConsider (10.26) and suppose that p 1 = p 2 = p is a piece-wise constant function: p = 1when the pre-synaptic oscillator, x k , is on the right branch of the cubic x-nullcl<strong>in</strong>ecorrespond<strong>in</strong>g to an active state, and p = 0 otherwise; see Fig. 10.40a. Somers and


Synchronization (see www.izhikevich.com) 499(a)p(x)=1(c)y 1 (t)y 2 (t)p(x)=0(b)(d)f(x,f(x,f(x,y)=0y)+=0f(x,y)=0y)+=0dd'ab'ab'Figure 10.40: Fast threshold modulation of relaxation oscillation. (a) The Heavisideor sigmoidal coupl<strong>in</strong>g function p(x) is constant while x is on the left or right branch ofthe x-nullcl<strong>in</strong>e. (b) In the relaxation limit µ = 0, the synchronized limit cycle consistsof the left branch of the nullcl<strong>in</strong>e f(x, y) = 0 and the right branch of the nullcl<strong>in</strong>ef(x, y) + ε = 0. When oscillator 1 is ahead of oscillator 2 (c), the phase differencebetween them decreases after the jump (d).Kopell (1993, 1995) referred to such a coupl<strong>in</strong>g <strong>in</strong> the relaxation limit µ → 0 as fastthreshold modulation (FTM), and found a simple criterion of stability of synchronizedstate that works even for strong coupl<strong>in</strong>g.S<strong>in</strong>ce the oscillators are identical, the <strong>in</strong>-phase synchronized state exists, dur<strong>in</strong>gwhich the variables x 1 and x 2 follow the left branch of the x-nullcl<strong>in</strong>e def<strong>in</strong>ed byf(x, y) = 0, see Fig. 10.40b, until they reach the jump<strong>in</strong>g po<strong>in</strong>t a. Dur<strong>in</strong>g the <strong>in</strong>stantaneousjump, they turn on the mutual coupl<strong>in</strong>g ε, and land at some po<strong>in</strong>t b ′ onthe perturbed x-nullcl<strong>in</strong>e def<strong>in</strong>ed by f(x, y) + ε = 0. They follow the new nullcl<strong>in</strong>euntil the right (upper) knee, and then jump back to the left branch.To determ<strong>in</strong>e the stability of the <strong>in</strong>-phase synchronization, we consider the casewhen oscillator 1 is slightly ahead of oscillator 2, as <strong>in</strong> Fig. 10.40c. We assume thatthe phase difference between the oscillators is so small, or alternatively, the strengthof coupl<strong>in</strong>g is so large, that when oscillator 1 jumps and turns on its <strong>in</strong>put to oscillator2, the latter, be<strong>in</strong>g at po<strong>in</strong>t d <strong>in</strong> Fig. 10.40d, is below the left knee of the perturbedx-nullcl<strong>in</strong>e f(x, y) + ε = 0 and therefore jumps too. As a result, both oscillators jumpto the perturbed x-nullcl<strong>in</strong>e and reverse their order. Although the apparent distancebetween the oscillators, measured by the difference of their y-coord<strong>in</strong>ates, is preserveddur<strong>in</strong>g such a jump, the phase difference between them is usually not.The phase difference between two po<strong>in</strong>ts on a limit cycle is the time needed totravel from one po<strong>in</strong>t to the other. Let τ 0 (d) be the time needed to slide from po<strong>in</strong>td to po<strong>in</strong>t a along the x-nullcl<strong>in</strong>e <strong>in</strong> Fig. 10.40d, i.e., the phase difference just before


500 Synchronization (see www.izhikevich.com)the jump. Let τ 1 (d) be the time needed to slide from po<strong>in</strong>t b ′ to po<strong>in</strong>t d ′ , i.e., thephase difference after the jump. The phase difference between the oscillators dur<strong>in</strong>gthe jump changes by the factor C(d) = τ 1 (d)/τ 0 (d) called the compression function.The difference decreases when the compression function C(d) < 1 uniformly for all dnear the left knee a. This condition has a simple geometrical mean<strong>in</strong>g: The rate ofchange of y(t) is slower before the jump than after it, so that y(t) has a “scalloped”shape, as <strong>in</strong> Fig. 10.40c. As an exercise, prove that C(d) → |g(a)|/|g(b ′ )| as d → a.If the compression function at the right (upper) knee is also less than 1, then the<strong>in</strong>-phase synchronization is stable. Indeed, the phase difference does not change whilethe oscillators slide along the nullcl<strong>in</strong>es, and it decreases geometrically with each jump.In fact, it suffices to require that the product of compression factors at the two kneesbe less than 1, so that any expansion at one knee is compensated by even strongercompression at the other knee.10.4.5 Burst<strong>in</strong>g oscillatorsLet us consider burst<strong>in</strong>g neurons coupled weakly through their fast variables:ẋ i = f(x i , y i ) + εp(x i , x j ) , (10.27)ẏ i = µg(x i , y i ) , (10.28)i = 1, 2 and j = 2, 1. S<strong>in</strong>ce burst<strong>in</strong>g <strong>in</strong>volves two time scales, fast spik<strong>in</strong>g and slowtransition between spik<strong>in</strong>g and rest<strong>in</strong>g, there are two synchronization regimes: spikesynchronization and burst synchronization, illustrated <strong>in</strong> Fig. 9.51 and discussed <strong>in</strong>Sect. 9.4. Below we outl<strong>in</strong>e some useful ideas and methods of study<strong>in</strong>g both regimes.Our exposition is not complete, but it rather lays the foundation for a more detailedresearch program.Spike synchronizationTo study synchronization of <strong>in</strong>dividual spikes with<strong>in</strong> the burst, we let µ = 0 to freezethe slow subsystem (10.28) and consider the fast subsystem (10.27) describ<strong>in</strong>g weaklycoupled oscillators. When y i ≈ y j , the fast variables oscillate with approximately equalperiods, so (10.27) can be reduced to the phase model˙ϕ i = εH(ϕ j − ϕ i , y i ) ,where y i = const parameterize the form of the connection function. For example,dur<strong>in</strong>g the ”circle/Hopf” burst, the function is transformed from H(χ) = s<strong>in</strong> 2 χ or1 − cos χ at the beg<strong>in</strong>n<strong>in</strong>g of the burst (saddle-node on <strong>in</strong>variant circle bifurcation)to H(χ) = s<strong>in</strong> χ at the end of the burst (supercritical Andronov-Hopf bifurcation).Chang<strong>in</strong>g y i slowly, one can study when spike synchronization appears and when itdisappears dur<strong>in</strong>g the burst. When the slow variables y i have different values, fastvariables typically oscillate with different frequencies, so one needs to look at low-orderresonances (see next section) to study the possibility of spike synchronization.


Synchronization (see www.izhikevich.com) 501rest<strong>in</strong>ga 2 'slow variable, yb 2a 1slownullcl<strong>in</strong>e, g=0spik<strong>in</strong>gX equiv (y, e)b 1 'C 2 = |g(a 2 ')||g(b 2 )|C 1 = |g(a 1 )||g(b 1 ')|y(t)X equiv (y, 0)fast variable, xFigure 10.41: Reduction of the I Na,p +I K +I K(M) -burster to a relaxation oscillator. Theslow variable exhibits “scalloped” oscillations needed for stability of <strong>in</strong>-phase burstsynchronization. C 1 and C 2 are compression functions at the two jumps.Burst synchronizationIn Chap. 9 we presented two methods, averag<strong>in</strong>g and equivalent voltage, that removefast oscillations and reduce bursters to slow relaxation oscillators. Burst synchronizationis then reduced to synchronization of such oscillators, and it can be studied us<strong>in</strong>gphase reduction or fast threshold modulation (FTM) approaches.To apply FTM, we assume that the coupl<strong>in</strong>g <strong>in</strong> (10.27) is piece-wise constant, thatis, p(x i , x j ) = 0 when the presynaptic burster x j is rest<strong>in</strong>g, and p(x i , x j ) = 1 (or anyfunction of x i ) when the presynaptic burster is spik<strong>in</strong>g. We also assume that the slowsubsystem (10.28) is one-dimensional so that we can use the equivalent voltage method(Sect. 9.2.4) and reduce the coupled system to0 = X equiv (y i , εp) − x i ,y ′ i = g(x i , y i ) .When the burster is of the hysteresis loop type, i.e., the rest<strong>in</strong>g and spik<strong>in</strong>g statescoexist, the function x = X equiv (y, εp) often, but not always, has a Z-shape on theslow/fast plane, as <strong>in</strong> Fig. 9.16, so that the system corresponds to a relaxation oscillatorwith nullcl<strong>in</strong>es as <strong>in</strong> Fig. 10.41. Fast threshold modulation occurs via the constant εp,which shifts the fast nullcl<strong>in</strong>e up or down. The compression criterion for stability ofthe <strong>in</strong>-phase burst synchronization, presented <strong>in</strong> the previous section, has a simplegeometrical illustration <strong>in</strong> the figure. The slow nullcl<strong>in</strong>e has to be sufficiently closeto the jump<strong>in</strong>g po<strong>in</strong>ts so that y(t) slows down before each jump and produces the“scalloped” shape curve. Many hysteresis loop fast/slow bursters do generate suchshape. In particular, “fold/*” bursters exhibit robust <strong>in</strong>-phase burst synchronizationwhen they are near the bifurcation from quiescence to burst<strong>in</strong>g, s<strong>in</strong>ce the slow nullcl<strong>in</strong>eis so close to the left knee that the compression dur<strong>in</strong>g the rest<strong>in</strong>g → spik<strong>in</strong>g jump (C 1<strong>in</strong> Fig. 10.41) dom<strong>in</strong>ates the expansion, if any, dur<strong>in</strong>g the spik<strong>in</strong>g → rest<strong>in</strong>g jump.


502 Synchronization (see www.izhikevich.com)Review of Important Concepts• Oscillations are described by their phase variables ϑ rotat<strong>in</strong>g on a circleS 1 . We def<strong>in</strong>e ϑ as the time s<strong>in</strong>ce the last spike.• The phase response curve, PRC (ϑ), describes the magnitude of the phaseshift of an oscillator caused by a strong pulsed <strong>in</strong>put arriv<strong>in</strong>g at phase ϑ.• PRC depends on the bifurcations of spik<strong>in</strong>g limit cycle, and it def<strong>in</strong>essynchronization properties of an oscillator.• Two oscillators are synchronized <strong>in</strong>-phase, anti-phase, or out-of-phase,when their phase difference, ϑ 2 − ϑ 1 , equals 0, half-period, or some othervalue, respectively; see Fig. 10.42.• Synchronized states of pulse-coupled oscillators are fixed po<strong>in</strong>ts of thecorrespond<strong>in</strong>g Po<strong>in</strong>care phase map.• Weakly coupled oscillatorscan be reduced to phase modelsẋ i = f(x i ) + ε ∑ g ij (x j )˙ϑ i = 1 + ε Q(ϑ i ) ∑ g ij (x j (ϑ j )) ,where Q(ϑ) is the <strong>in</strong>f<strong>in</strong>itesimal PRC def<strong>in</strong>ed by (10.10).• Weak coupl<strong>in</strong>g <strong>in</strong>duces a slow phase deviation of the natural oscillation,ϑ i (t) = t + ϕ i , described by the averaged model(˙ϕ i = ε ω i + ∑ )H ij (ϕ j − ϕ i ) ,where the ω i denote the frequency deviations, andH ij (ϕ j − ϕ i ) = 1 T∫ Tdescribe the <strong>in</strong>teractions between the phases.0Q(t) g ij (x j (t + ϕ j − ϕ i )) dt• Synchronization of two coupled oscillators correspond to equilibria of theone-dimensional system˙χ = ε(ω + G(χ)) , χ = ϕ 2 − ϕ 1 ,where G(χ) = H 21 (−χ) − H 12 (χ) describes how the phase difference χcompensates for the frequency mismatch ω = ω 2 − ω 1 .


Synchronization (see www.izhikevich.com) 503<strong>in</strong>-phase anti-phase out-of-phaseBibliographical NotesFigure 10.42: Different types of synchronization.Surpris<strong>in</strong>gly, this chapter turned out to be quite different from Chap. 9 (Weakly ConnectedOscillators) of the book Weakly Connected Neural Networks by Hoppensteadtand Izhikevich (1997) or from the book Synchronization: A universal concept <strong>in</strong> nonl<strong>in</strong>earsciences by Pikovsky, Rosenblum and Kurths. All three texts, though devotedto the same subject, do not repeat, but rather complement each other. The latter providesan excellent historical overview of synchronization, start<strong>in</strong>g with the work of thefamous Dutch mathematician, astronomer, and physicist Christiaan Huygens (1629–1695), who was the first one to describe synchronization of a couple of pendulum clockshang<strong>in</strong>g from a common support, which was <strong>in</strong>cidentally anti-phase. While provid<strong>in</strong>gmany examples of synchronization of biological, chemical, and physical systems, thebook by Pikovsky et al. also discusses the def<strong>in</strong>ition of a phase and synchronization ofnon-periodic, e.g., chaotic, oscillators, a topic not covered here. A major part of Spik<strong>in</strong>gNeuron Models by Gerstner and Kistler (2002) is devoted to synchronization ofspik<strong>in</strong>g neurons, with the emphasis on the <strong>in</strong>tegrate-and-fire model and spike-responsemethod.The formalism of phase response curve (PRC) was <strong>in</strong>troduced by Hast<strong>in</strong>gs andSweeney (1958), and it has been used extensively <strong>in</strong> the context of resett<strong>in</strong>g the circadianrhythms. Forty Years of PRC — What Have We Learned? by Johnson (1999)gives an historical overview of this idea and some recent developments. John Guckenheimer(1975) used the theory of normally hyperbolic <strong>in</strong>variant manifolds to provide amathematical foundation for the existence of isochrons, and their geometrical properties.An encyclopedic exposition of isochrons and phase resett<strong>in</strong>gs <strong>in</strong> nature, as well asnumerous anecdotes, can be found <strong>in</strong> Arthur W<strong>in</strong>free’s remarkable book The Geometryof Biological Time (first edition 1980, second edition 2001). In particular, W<strong>in</strong>free describesthe work of George R. M<strong>in</strong>es (1914) who was do<strong>in</strong>g phase resett<strong>in</strong>g experimentsby shock<strong>in</strong>g rabbits at various phases of their heartbeat. He found the phase and shockthat could stop a rabbit’s heart (black hole <strong>in</strong> Fig. 10.8), and then applied it to himself.He died.Glass and MacKey (1988) provide a detailed exposition of circle phase maps. Althoughthe structure of phase-lock<strong>in</strong>g regions <strong>in</strong> Fig. 10.15 was discovered by Cartwrightand Littlewood (1945), it is better known at present as Arnold tongues (Arnold 1965).


504 Synchronization (see www.izhikevich.com)Figure 10.43: Frank Hoppensteadt,the author’s adviser and mentor, circa1989.Guevara and Glass (1982) found this structure analytically for the Andronov-Hopf oscillator<strong>in</strong> Fig. 10.3 (radial isochron clock). Hoppensteadt (1997, 2000) provides manyexamples of oscillatory systems aris<strong>in</strong>g <strong>in</strong> biology and neuroscience; see also Hoppensteadtand Pesk<strong>in</strong> (2002).Malk<strong>in</strong>’s method of reduction of coupled oscillators to phase equations has beenknown, at least to Russian scientists, s<strong>in</strong>ce the early 1950s (Malk<strong>in</strong> 1949, 1956, Blechman1971). For example, Melnikov (1963) applied Malk<strong>in</strong>’s theorem to a homocl<strong>in</strong>icorbit of <strong>in</strong>f<strong>in</strong>ite period to obta<strong>in</strong> the transversality condition for the saddle homocl<strong>in</strong>icorbit bifurcation (Kuznetsov 1995).Malk<strong>in</strong>’s method was rediscovered <strong>in</strong> the West by J. Neu (1979), and hoorayedby W<strong>in</strong>free (1980) who f<strong>in</strong>ally saw a mathematical justification for his usage of phasevariables. S<strong>in</strong>ce then, the field of phase equations was largely dom<strong>in</strong>ated by BardErmentrout and Nancy Kopell (Ermentrout 1981, 1985, 1992, Ermentrout and Kopell1986a,b, 1990, 1991, 1994, Kopell and Ermentrout 1990, Kopell 1986, Kopell et al.1991). In particular, they developed the theory of travel<strong>in</strong>g wave solutions <strong>in</strong> cha<strong>in</strong>sof oscillators, build<strong>in</strong>g on the sem<strong>in</strong>al paper by Cohen et al. (1982). Incidentally, theone-page proof of Malk<strong>in</strong>’s theorem provided by Hoppensteadt and Izhikevich (1997,Sect. 9.6) is based on Ermentrout and Kopell idea of us<strong>in</strong>g the Fredholm alternative;Malk<strong>in</strong>’s and Neu’s proofs are quite long, mostly because they re-prove the alternative.There are only a handful of examples where the Malk<strong>in</strong> adjo<strong>in</strong>t problem can besolved analytically, i.e., without resort to simulations. The SNIC, homocl<strong>in</strong>ic andAndronov-Hopf cases are the most important and were considered <strong>in</strong> detail <strong>in</strong> thischapter. Brown et al. (2004) also derive PRCs for oscillators with homocl<strong>in</strong>ic orbitsto pure saddles (see Ex. 25) and for Baut<strong>in</strong> oscillators.Throughout this chapter we def<strong>in</strong>e the phase ϑ or ϕ on the <strong>in</strong>terval [0, T ], where Tis the period of free oscillation, and do not normalize it to be on the <strong>in</strong>terval [0, 2π].


Synchronization (see www.izhikevich.com) 505Figure 10.44: Nancy Kopell <strong>in</strong> her Boston University office <strong>in</strong> 1990 (photograph providedby Dr. Kopell).As a result, we avoided the annoy<strong>in</strong>g terms 2π/T and 2π/Ω <strong>in</strong> the formulae. The onlydrawback is that some of the results may have an “unfamiliar look”, such as s<strong>in</strong> 2 ϑ withϑ ∈ [0, π] for the PRC of Class 1 neurons, as opposed to 1 − cos ϑ with ϑ ∈ [0, 2π] usedby many authors before.Hansel, Mato, and Meunier (1995) were the first to notice that the shape of PRCdeterm<strong>in</strong>es the synchronization properties of synaptically coupled oscillators. Ermentrout(1996) related this result to the classification of oscillators and proved that PRCsof all Class 1 oscillators have the form of 1−cos ϑ, though the proof can be found <strong>in</strong> hisearlier papers with Kopell (Ermentrout and Kopell 1986a,b). Reyes and Fetz (1993)measured PRC of a cat neocortical neuron and largely confirmed the theoretical predictions.The experimental method <strong>in</strong> Sect. 10.2.4 is related to that of Rosenblum andPikovsky (2001). It needs to be developed further, e.g., by <strong>in</strong>corporat<strong>in</strong>g the measurementuncerta<strong>in</strong>ty (error bars). In fact, most experimentally obta<strong>in</strong>ed PRCs, <strong>in</strong>clud<strong>in</strong>gthe one <strong>in</strong> Fig. 10.24, are so noisy that noth<strong>in</strong>g useful could be derived from them.This issue is the subject of active research.Our treatment of the FTM theory follows closely that of Somers and Kopell (1993,1995). Anti-phase synchronization of relaxation oscillators is analyzed us<strong>in</strong>g phasemodels by Izhikevich (2000) and FTM theory by Kopell and Somers (1995). Ermentrout(1994) and Izhikevich (1998) considered weakly coupled oscillators with axonalconduction delays and showed that delays result <strong>in</strong> mere phase shifts (see Ex. 18).Frankel and Kiemel (1993) observed that slow coupl<strong>in</strong>g can be reduced to weak coupl<strong>in</strong>g.Izhikevich and Hoppensteadt (2003) used Malk<strong>in</strong>’s theorem to extend the resultsto slowly coupled networks, and to derive useful formulae for the coupl<strong>in</strong>g functions andcoefficients. Ermentrout (2003) showed that the result could be generalized to synapseshav<strong>in</strong>g fast-ris<strong>in</strong>g and slow-decay<strong>in</strong>g conductances. Goel and Ermentrout (2002) and


506 Synchronization (see www.izhikevich.com)Katriel (2005) obta<strong>in</strong>ed many <strong>in</strong>terest<strong>in</strong>g results on <strong>in</strong>-phase synchronization of identicalphase oscillators.Interactions between resonant oscillators were considered by Ermentrout (1981),Hoppensteadt and Izhikevich (1997) and Izhikevich (1999) <strong>in</strong> the context of quasiperiodic(multi-frequency) oscillations. Baesens et al. (1991) undertook heroic taskof study<strong>in</strong>g resonances and toroidal chaos <strong>in</strong> a system of three coupled phase oscillators.Mean-field approaches to the Kuramoto model are reviewed by Strogatz (2000)and Acebron et al. (2005). Daido (1996) extended the theory to the general coupl<strong>in</strong>gfunction H(χ). van Hemmen and Wresz<strong>in</strong>ski (1993) were the first to f<strong>in</strong>d the Lyapunovfunction for the Kuramoto model, which was generalized (<strong>in</strong>dependently) byHoppensteadt and Izhikevich (1997) to arbitrary coupl<strong>in</strong>g function H(χ).Ermentrout (1985), Aronson et al. (1990) and Hoppensteadt and Izhikevich (1996,1997) studied weakly coupled Andronov-Hopf oscillators, and discussed the phenomenaof self-ignition (coupl<strong>in</strong>g-<strong>in</strong>duced oscillations) and oscillator death (coupl<strong>in</strong>g-<strong>in</strong>ducedcessation of oscillation). Coll<strong>in</strong>s and Stewart (1993, 1994) and Golubitsky and Stewart(2002) applied group theory to study synchronization of coupled oscillators <strong>in</strong> networkswith symmetries.In this chapter we consider either strong pulsed coupl<strong>in</strong>g or weak cont<strong>in</strong>uous coupl<strong>in</strong>g.These limitations are severe, but they allow us to derive model-<strong>in</strong>dependentresults. Study<strong>in</strong>g synchronization <strong>in</strong> networks of strongly coupled neurons is an activearea of research, however, most of such studies fall <strong>in</strong>to two categories: (1) simulationsand (2) <strong>in</strong>tegrate-and-fire networks. In both cases, the results are model-depended. Ifthe reader wants to pursue this l<strong>in</strong>e of research, he or she will def<strong>in</strong>itely need to readthe follow<strong>in</strong>g papers: Mirollo and Strogatz (1990), van Vreeswijk et al. (1994), Chowand Kopell (2000), Rub<strong>in</strong> and Terman (2000, 2002), Bressloff and Coombes (2000),van Vreeswijk (2000), van Vreeswijk and Hansel (2001), Pfeuty et al. (2003), Hanseland Mato (2003).Exercises1. F<strong>in</strong>d the isochrons of the Andronov-Hopf oscillatorż = (1 + i)z − z|z| 2 , z ∈ C ,<strong>in</strong> Fig. 10.3.2. Prove that the isochrons of the Andronov-Hopf oscillator <strong>in</strong> the formż = (1 + i)z + (−1 + di)z|z| 2 , z ∈ C ,are the curvesz(s) = s (−1+di) e χi , s > 0 ,where χ is the phase of the isochron; see Fig. 10.45.


Synchronization (see www.izhikevich.com) 507d=-2d=+2s (-1+di)1100-1-1s (-1+di)-1 0 1-1 0 1Figure 10.45: Isochrons of the Andronov-Hopf oscillator; see Ex. 2.cos s<strong>in</strong> cos s<strong>in</strong>Az(t)Figure 10.46: Pulsed stimulation of theAndronov-Hopf oscillator <strong>in</strong> Fig. 10.3;see Ex. 4.3. [MATLAB] To determ<strong>in</strong>e isochrons of an oscillator ẋ = F (x), one can start withmany <strong>in</strong>itial po<strong>in</strong>ts near the limit cycle and <strong>in</strong>tegrate the system backwards, i.e.ẋ = −F (x). The images of the po<strong>in</strong>ts at any time t lie on the same isochron.Write a MATLAB program that implements this algorithm.4. Prove that the phase response curve of the Andronov-Hopf oscillator <strong>in</strong> Fig. 10.3is{ −ψ when 0 ≤ ϑ ≤ π,PRC (ϑ, A) =(10.29)+ψ when π ≤ ϑ ≤ 2π,whereψ = arcos1 + A cos ϑ√1 + 2A cos ϑ + A2and A is the magnitude of the horizontal displacement of z(t); see Fig. 10.46.5. [MATLAB] Write a program that stimulates an oscillator at different phases anddeterm<strong>in</strong>es its phase response curve (PRC).6. Show that Z(ϑ) = grad ϑ, so that W<strong>in</strong>free’s phase model (10.6) is equivalent toKuramoto’s phase model (10.8).7. Show that Z(ϑ) = Q(ϑ), so that W<strong>in</strong>free’s phase model (10.6) is equivalent toMalk<strong>in</strong>’s phase model (10.9).


508 Synchronization (see www.izhikevich.com)8. Show that the PRC of the leaky <strong>in</strong>tegrate-and-fire neuron (Sect. 8.1.1)˙v = b − v ,if v ≥ 1 (threshold), then v ← 0 (reset)with b > 1 has the formPRC (ϑ) = m<strong>in</strong> {ln(b/(b exp(−ϑ) − A)), T } − ϑ ,where T = ln(b/(b − 1)) is the period of free oscillations and A is the amplitudeof the pulse.9. Prove that the quadratic <strong>in</strong>tegrate-and-fire neuron˙v = 1 + v 2 ,if v = +∞ (peak of spike), then v ← −∞ (reset)has PTC (ϑ) = π/2 + atan (A − cot ϑ).10. F<strong>in</strong>d PRC of the quadratic <strong>in</strong>tegrate-and-fire neuron (Sect. 8.1.3)˙v = b + v 2 ,if v ≥ 1 (peak of spike), then v ← v reset (reset)with b > 0.11. Consider two mutually pulsed-coupled oscillators with periods T 1 ≈ T 2 and type1 phase transition curves PTC 1 and PTC 2 , respectively. Show that the lock<strong>in</strong>gbehavior of the system can be described by the Po<strong>in</strong>care phase mapχ n+1 = T 1 − PTC 1 (T 2 − PTC 2 (χ n )) ,where χ n is the phase difference between the oscillators, i.e., the phase of oscillator2 when oscillator 1 fires a spike.12. [MATLAB] Write a program that solves the adjo<strong>in</strong>t equation (10.10) numerically(h<strong>in</strong>t: <strong>in</strong>tegrate the equation backwards to achieve stability).13. [MATLAB] Write a program that f<strong>in</strong>ds the <strong>in</strong>f<strong>in</strong>itesimal PRC us<strong>in</strong>g the relationship˙ϑ = 1 + PRC (ϑ) εp(t) ,the moments of fir<strong>in</strong>gs of a neuron (zero cross<strong>in</strong>gs of ϑ(t)), and the <strong>in</strong>jectedcurrent εp(t); see Sect. 10.2.4 and Fig. 10.24.14. Use the approaches of W<strong>in</strong>free, Kuramoto, and Malk<strong>in</strong> to transform the <strong>in</strong>tegrateand-fireneuron ˙v = b − v + εp(t) <strong>in</strong> Ex. 8 to its phase modelwith T = ln(b/(b − 1)).˙ϑ = 1 + ε ( e ϑ /b ) p(t) ,


Synchronization (see www.izhikevich.com) 50915. Use the approaches of W<strong>in</strong>free, Kuramoto, and Malk<strong>in</strong> to transform the quadratic<strong>in</strong>tegrate-and-fire neuron ˙v = 1 + v 2 + εp(t) <strong>in</strong> Ex. 9 to its phase modelwith T = π.˙ϑ = 1 + ε (s<strong>in</strong> 2 ϑ) p(t) ,16. Use the approaches of W<strong>in</strong>free, Kuramoto, and Malk<strong>in</strong> to transform the Andronov-Hopf oscillator ż = (1 + i)z − z|z| 2 + εp(t) with real p(t) to its phase modelwith T = 2π.˙ϑ = 1 + ε (− s<strong>in</strong> ϑ)p(t) ,17. (PRC for Andronov-Hopf) Consider a weakly perturbed system near supercriticalAndronov-Hopf bifurcation (see Sect. 6.1.3)ż = (b + i)z + (−1 + di)z|z| 2 + ɛp(t) , z ∈ C .with b > 0. Let ε = b √ b/ɛ be small. Prove that the correspond<strong>in</strong>g phase modelis˙θ = 1 + d + ε Im {(1 + di)p(t)e −iθ } .When the forc<strong>in</strong>g p(t) is one-dimensional, i.e., p(t) = cq(t) with c ∈ C and scalarfunction q(t), the phase model has s<strong>in</strong>usoidal form˙θ = 1 + d + εs s<strong>in</strong>(θ − ψ)q(t) ,with the strength s and the phase shift ψ depend<strong>in</strong>g only on d and c.18. (Delayed coupl<strong>in</strong>g) Show that weakly coupled oscillatorsẋ i = f(x i ) + εn∑g ij (x i (t), x j (t − d ij ))with explicit axonal conduction delays d ij ≥ 0 have the phase modelj=1ϕ ′ i = ω i +n∑H ij (ϕ j − d ij − ϕ i )j≠iwhere ′ = d/dτ, τ = εt is the slow time, and H(χ) is def<strong>in</strong>ed by (10.16). Thus,explicit delays result <strong>in</strong> explicit phase shifts.19. Determ<strong>in</strong>e the existence and stability of synchronized states <strong>in</strong> the system˙ϕ 1 = ω 1 + c 1 s<strong>in</strong>(ϕ 2 − ϕ 1 )˙ϕ 2 = ω 2 + c 2 s<strong>in</strong>(ϕ 1 − ϕ 2 )as a function of the parameters ω = ω 2 − ω 1 and c = c 2 − c 1 .


510 Synchronization (see www.izhikevich.com)20. Consider the Kuramoto modelϕ i = ω +n∑c ij s<strong>in</strong>(ϕ j + ψ ij − ϕ i ) ,j=1where c ij and ψ ij are parameters. What can you say about its synchronizationproperties?21. Derive the self-consistency equation (10.22) for the Kuramoto model (10.20).22. Consider the phase deviation modelϕ ′ 1 = ω + c 1 H(ϕ 2 − ϕ 1 )ϕ ′ 2 = ω + c 2 H(ϕ 1 − ϕ 2 )with an even function H(−χ) = H(χ). Prove that the <strong>in</strong>-phase synchronizedstate, ϕ 1 = ϕ 2 , if it exists, cannot be exponentially stable. What can you sayabout the anti-phase state ϕ 1 = ϕ 2 + T/2?23. Prove that the <strong>in</strong>-phase synchronized state <strong>in</strong> a network of three or more pulsecoupledquadratic <strong>in</strong>tegrate-and-fire neurons is unstable.24. Prove (10.25).25. (Brown et al. 2004) Show that PRC for an oscillator near saddle homocl<strong>in</strong>ic orbitbifurcation scales as PRC (ϑ) ∼ e λ(T −ϑ) , where λ is the positive eigenvalue of thesaddle and T is the period of oscillation.26. Consider the quadratic <strong>in</strong>tegrate-and-fire neuron ẋ = ±1 + x 2 with the resett<strong>in</strong>g“ if x = +∞, then x ← x reset ”. Prove thatregime SNIC homocl<strong>in</strong>icmodel x ′ = +1 + x 2 x ′ = −1 + x 2 , (x reset > 1)x reset x x -1 1 x resetperiod T π/2 − atan x reset acoth x resetsolution x(t) − cot(t − T ) − coth(t − T )PRC Q(ϑ) s<strong>in</strong> 2 (ϑ − T ) s<strong>in</strong>h 2 (ϑ − T )15x reset =+1.1x reset =-1.100T00T


Solutions to Exercises, Chap. 10 511b 2a 2a 1b 1g(x, y)=0f(x, y)=0a 1a 1|g(a 1 )| |g(a 1 )|Figure 10.47: Left: Relaxation oscillator <strong>in</strong> the limit µ = 0 near the onset of oscillation.Middle and right: A magnification of a neighborhood of the jump po<strong>in</strong>t a 1 for variousg(a 1 ) and µ. Canard (French duck) solutions can appear when g(a 1 ) ≪ µ.where coth, acoth, and s<strong>in</strong>h are hyperbolic cotangent, hyperbolic <strong>in</strong>verse cotangentand hyperbolic s<strong>in</strong>e, respectively.27. [M.S.] Derive the PRC for an oscillator near saddle homocl<strong>in</strong>ic orbit bifurcationthat is valid dur<strong>in</strong>g the spike downstroke. Take advantage of the observation <strong>in</strong>Fig. 10.39 that the homocl<strong>in</strong>ic orbit consists of two qualitatively different parts.28. [M.S.] Derive PRC for a generic oscillator near fold limit cycle bifurcation.29. [M.S.] Simplify the connection function H for coupled relaxation oscillators(Izhikevich 2000) when the slow nullcl<strong>in</strong>e approaches the left knee, as <strong>in</strong> Fig. 10.47.Explore the range of parameters ε, µ, and |g(a 1 )| where the analysis is valid.30. [Ph.D.] Use ideas outl<strong>in</strong>ed <strong>in</strong> Sect. 10.4.5 to develop the theory of reduction ofweakly coupled bursters to phase models. Do not assume that burst<strong>in</strong>g trajectoryis periodic.Solutions to Chapter 101. In polar coord<strong>in</strong>ates, z = re iϑ , the system has the form˙ϑ = 1 , ṙ = r − r 3 .S<strong>in</strong>ce the phase of oscillation does not depend on the amplitude, the isochrons have the radialstructure depicted <strong>in</strong> Fig. 10.3.2. In polar coord<strong>in</strong>ates, the oscillator has the form˙ϑ = 1 + dr 2 , ṙ = r − r 3 .The second equation has an explicit solution r(t), such thatr(t) 2 =11 − (1 − 1/r(0) 2 )e −2t .


512 Solutions to Exercises, Chap. 10The phase difference between ϑ ˙ lc = 1 + d(1) 2 and ˙ϑ = 1 + dr(t) 2 grows as ˙χ = d(r(t) 2 − 1), andits asymptotic value isχ(∞) =∫ ∞Thus, on the χ-isochron, we have ϑ + d log r = χ.3. An example is the file isochrons.m0d(r(t) 2 − 1) = d log r(0) .function isochrons(F,phases,x0)% plot isochrons of a planar dynamical system x’=F(t,x)% at po<strong>in</strong>ts given by the vector ’phases’.% ’x0’ is a po<strong>in</strong>t on the limit cycle (2x1-vector)T= phases(end); % is the period of the cycletau = T/600; % time step of <strong>in</strong>tegrationm=200;% spatial gridk=5;% the number of skipped cycles[t,lc] = ode23s(F,0:tau:T,x0);dx=(max(lc)-m<strong>in</strong>(lc))’/m;center = (max(lc)+m<strong>in</strong>(lc))’/2;iso=[x0-m^0.5*dx, x0+m^0.5*dx];% forward <strong>in</strong>tegration% spatial resolution% center of the limit cycle% isochron’s <strong>in</strong>itial segmentfor t=0:-tau:-(k+1)*T% backward <strong>in</strong>tegrationfor i=1:size(iso,2)iso(:,i)=iso(:,i)-tau*feval(F,t,iso(:,i)); % move one stepend;i=1;while i1.5*m*dx) % check boundariesiso = [iso(:,1:i-1), iso(:,i+1:end)]; % removeelsei=i+1;end;end;i=1;while i 2% add a po<strong>in</strong>t <strong>in</strong> the middleiso = [iso(:,1:i), (iso(:,i)+iso(:,i+1))/2 ,iso(:,i+1:end)];end;if d < 0.5% remove the po<strong>in</strong>tiso = [iso(:,1:i), iso(:,i+2:end)];elsei=i+1;end;end;if (mod(-t,T)


Solutions to Exercises, Chap. 10 513hold off;The call of the function is isochrons(’F’,0:0.1:2*pi,[1;0]); withfunction dx = F(t,x);z=x(1)+1i*x(2);dz=(1+1i)*z-z*z*conj(z);dx=[real(dz); imag(dz)];4. (Hoppensteadt and Keener 1982) From calculus, B · C = |B| |C| cos(ψ). S<strong>in</strong>ce |B| = 1 andC = (A + cos ϑ, s<strong>in</strong> ϑ), see Fig. 10.46, we have B · C = A cos ϑ + cos 2 ϑ + s<strong>in</strong> 2 ϑ. Hence,cos ψ = (1+A cos ϑ)/ √ 1 + 2A cos ϑ + A 2 . When ϑ is <strong>in</strong> the upper (lower) half-plane, the phaseis delayed (advanced).5. An example is the file prc.mfunction PRC=prc(F,phases,x0,A)% plot phase-resett<strong>in</strong>g curve (PRC) of system x’=F(t,x) + A delta(t)% at po<strong>in</strong>ts given by the vector ’phases’.% ’x0’ is a po<strong>in</strong>t on the limit cycle with zero phase% A is the strength of stimulation (row-vector)% use peaks of spikes to f<strong>in</strong>d the phase differencesT= phases(end); % is the period of the cycletau = T/6000; % time step of <strong>in</strong>tegrationk=3;% the number of cycles needed to determ<strong>in</strong>e the new phasePRC=[];[tc,lc] = ode23s(F,0:tau:k*T,x0); % f<strong>in</strong>d limit cyclepeak=1+f<strong>in</strong>d(lc(2:end-1,1)>lc(1:end-2,1)&lc(2:end-1,1)>=lc(3:end,1));peak0 = tc(peak(end)); % the last peak is used for referencefor i=1:length(phases)[m,j]=m<strong>in</strong>(abs(phases(i)-tc));[t,x] = ode23s(F,phases(i):tau:k*T,lc(j,:)+A); % stimulatepeaks=1+f<strong>in</strong>d(x(2:end-1,1)>x(1:end-2,1)&x(2:end-1,1)>=x(3:end,1));PRC=[PRC, mod(T/2+peak0-t(peaks(end)),T)-T/2];subplot(2,1,2);drawnow;plot(phases(1:length(PRC)),PRC);xlabel(’phase of stimulation’);ylabel(’<strong>in</strong>duced phase difference’);subplot(2,1,1);plot(tc,lc(:,1),’r’,t,x(:,1),t(peaks(end)),x(peaks(end),1),’ro’);xlabel(’time’);ylabel(’membrane potential’);end;An example of a call of the function is PRC=prc(’F’,0:0.1:2*pi,[-1 0],[0.1 0]); withfunction dx = F(t,x);z=x(1)+1i*x(2);dz=(1+1i)*z-z*z*conj(z);dx=[real(dz); imag(dz)];6.grad ϑ(x) ==( ϑ(x + h1 ) − ϑ(x), . . . , ϑ(x + h )m) − ϑ(x)h 1h m( PRC1 (ϑ(x), h 1 ), . . . , PRC )m(ϑ(x), h m )h 1 h m


514 Solutions to Exercises, Chap. 10(<strong>in</strong> the limit h → 0)(Z1 (ϑ(x))h 1=, . . . , Z )m(ϑ(x))h mh 1h m= (Z 1 (ϑ(x)), . . . , Z m (ϑ(x))) = Z(ϑ(x)) .7. (Brown et al. 2004, Appendix A) Let x be a po<strong>in</strong>t on the limit cycle and z be an arbitrarynearby po<strong>in</strong>t. Let x(t) and z(t) be the trajectories start<strong>in</strong>g from the two po<strong>in</strong>ts, and y(t) =z(t) − x(t) be the difference. All equations below are valid up to O(y 2 ). The phase shift∆ϑ = ϑ(z(t)) − ϑ(x(t)) = grad ϑ(x(t)) · y(t) does not depend on time. Differentiat<strong>in</strong>g withrespect to time and tak<strong>in</strong>g grad ϑ(x(t)) = Z(ϑ(t)) (see previous exercise), results <strong>in</strong>0 = (d/dt) (Z(ϑ(t)) · y(t)) = Z ′ (ϑ(t)) · y(t) + Z(ϑ(t)) · Df(x(t))y(t)= Z ′ (ϑ(t)) · y(t) + Df(x(t)) ⊤ Z(ϑ(t)) · y(t)()= Z ′ (ϑ(t)) + Df(x(t)) ⊤ Z(ϑ(t)) · y(t) .S<strong>in</strong>ce y is arbitrary, Z satisfies Z ′ (ϑ) + Df(x(ϑ)) ⊤ Z(ϑ) = 0, i.e., the adjo<strong>in</strong>t equation (10.10).The normalization follows from (10.7).8. The solution to ˙v = b − v with v(0) = 0 is v(t) = b(1 − e −t ) with the period T = ln(b/(b − 1))determ<strong>in</strong>ed from the threshold cross<strong>in</strong>g v(T ) = 1. From v = b(1−e −ϑ ) we f<strong>in</strong>d ϑ = ln(b/(b−v)),hencePRC (ϑ) = ϑ new − ϑ = m<strong>in</strong> {ln(b/(b exp(−ϑ) − A), T } − ϑ .9. The system ˙v = 1 + v 2 with v(0) = −∞ has the solution (check by differentiat<strong>in</strong>g) v(t) =tan(t − π/2) with the period T = π. S<strong>in</strong>ce t = π/2 + atan v, we f<strong>in</strong>dPTC (ϑ) = π/2 + atan [A + tan(ϑ − π/2)]andPRC (ϑ) = PTC (ϑ) − ϑ = atan [A + tan(ϑ − π/2)] − (ϑ − π/2) .10. The system ˙v = b + v 2 with b > 0 and the <strong>in</strong>itial condition v(0) = v reset has the solution (checkby differentiat<strong>in</strong>g)v(t) = √ b tan( √ b(t + t 0 ))wheret 0 =1 √batan v reset√b.Equivalently,t = 1 √batan v √b− t 0 .From the condition v = 1 (peak of the spike), we f<strong>in</strong>dT = √ 1 atan √ 1 − t 0 = 1 (√ atan √ 1 − atan v )reset√ ,b b b b bHencePRC (ϑ) = ϑ new − ϑ = m<strong>in</strong> { 1 √batan [ A √b+ tan( √ b(ϑ + t 0 ))] − t 0 , T } − ϑ .11. Let ϑ denotes the phase of oscillator 1. Let χ n denote the phase of oscillator 2 just beforeoscillator 1 fires a spike, i.e., when ϑ = 0. This spike resets χ n to PTC 2 (χ n ). Oscillator 2 firesa spike when ϑ = T 2 −PTC 2 (χ n ), and it resets ϑ to PTC 1 (T 2 −PTC 2 (χ n )). F<strong>in</strong>ally, oscillator1 fires its spike when oscillator 2 has the phase χ n+1 = T 1 −PTC 1 (T 2 −PTC 2 (χ n )).


Solutions to Exercises, Chap. 10 51512. [MATLAB] An example is the file adjo<strong>in</strong>t.mfunction Q=adjo<strong>in</strong>t(F,t,x0)% f<strong>in</strong>ds solution to the Malk<strong>in</strong>’s adjo<strong>in</strong>t equation Q’ = -DF^t Q% at time-po<strong>in</strong>ts t with t(end) be<strong>in</strong>g the period% ’x0’ is a po<strong>in</strong>t on the limit cycle with zero phasetran=3; % the number of skipped cyclesdx = 0.000001; dy = 0.000001; % for evaluation of JacobianQ(1,:)=feval(F,0,x0)’;[t,x] = ode23s(F,t,x0);% <strong>in</strong>itial po<strong>in</strong>t;% f<strong>in</strong>d limit cyclefor k=1:tranQ(length(t),:)=Q(1,:); % <strong>in</strong>itial po<strong>in</strong>t;for i=length(t):-1:2% backward <strong>in</strong>tegrationL = [(feval(F,t(i),x(i,:)+[dx 0])-feval(F,t(i),x(i,:)))/dx,...(feval(F,t(i),x(i,:)+[0 dy])-feval(F,t(i),x(i,:)))/dy];Q(i-1,:) = Q(i,:) + (t(i)-t(i-1))*(Q(i,:)*L);end;end;Q = Q/(Q(1,:)*feval(F,0,x0)); % normalizationAn example of a call of the function is Q=adjo<strong>in</strong>t(’F’,0:0.01:2*pi,[1 0]); withfunction dx = F(t,x);z=x(1)+1i*x(2);dz=(1+1i)*z-z*z*conj(z);dx=[real(dz); imag(dz)];13. [MATLAB] We assume that PRC (ϑ) is given by its truncated Fourier series with unknownFourier coefficients. Then, we f<strong>in</strong>d the coefficients that m<strong>in</strong>imize the difference between predictedand actual <strong>in</strong>terspike <strong>in</strong>tervals. MATLAB file f<strong>in</strong>dprc.m takes the row vector of spikemoments, not count<strong>in</strong>g the spike at time zero, and the <strong>in</strong>put function p(t), determ<strong>in</strong>es thesampl<strong>in</strong>g frequency, the averaged period of oscillation and then calls file prcerror.m to f<strong>in</strong>dPRC.function PRC=f<strong>in</strong>dprc(sp,pp)global spikes p tau n% f<strong>in</strong>ds PRC of an oscillator theta’= 1 + PRC(theta)pp(t)% us<strong>in</strong>g the row-vector of spikes ’sp’ (when theta(t)=0)spikes = [0 sp];p=pp;tau = spikes(end)/length(p) % time step (sampl<strong>in</strong>g period)n=8; % The number of Fourier terms approximat<strong>in</strong>g PRCcoeff=zeros(1,2*n+1); % <strong>in</strong>itial approximationcoeff(2*n+2) = spikes(end)/length(spikes); % <strong>in</strong>itial periodcoeff=fm<strong>in</strong>search(’prcerror’,coeff);a = coeff(1:n) % Fourier coefficients for s<strong>in</strong>b = coeff(n+1:2*n) % Fourier coefficients for cosb0= coeff(2*n+1) % dc termT = coeff(2*n+2) % period of oscillationPRC=b0+sum((ones(floor(T/tau),1)*a).*s<strong>in</strong>((tau:tau:T)’*(1:n)*2*pi/T),2)+...sum((ones(floor(T/tau),1)*b).*cos((tau:tau:T)’*(1:n)*2*pi/T),2);


516 Solutions to Exercises, Chap. 10The follow<strong>in</strong>g program must be <strong>in</strong> the file prcerror.m.function err=prcerror(coeff)global spikes p tau na = coeff(1:n); % Fourier coefficients for s<strong>in</strong>b = coeff(n+1:2*n); % Fourier coefficients for cosb0= coeff(2*n+1); % dc termT = coeff(2*n+2); % period of oscillationerr=0;i=1;clf;for s=2:length(spikes)theta=0;while i*tau


Solutions to Exercises, Chap. 10 51716. W<strong>in</strong>free approach: Us<strong>in</strong>g results of Ex. 4,at A = 0.Z(ϑ) =∂∂A acos 1 + A cos ϑ√ = − s<strong>in</strong> ϑ1 + 2A cos ϑ + A2Kuramoto approach: S<strong>in</strong>ce grad ϑ(x) is orthogonal to the contour l<strong>in</strong>e of the function ϑ(x) atx, i.e., the isochron of x, and the results of Ex. 1 that isochrons are radial, we get grad (ϑ) =(− s<strong>in</strong> ϑ, cos ϑ) us<strong>in</strong>g purely geometrical considerations. S<strong>in</strong>ce p(t) is real, we need to keep onlythe first component.Malk<strong>in</strong> approach: Let us work <strong>in</strong> the complex doma<strong>in</strong>. On the circle z(t) = e it we get Df = i.S<strong>in</strong>ce Df ⊤ is equivalent to complex-conjugation <strong>in</strong> the complex doma<strong>in</strong>, we get ˙Q = i · Q,which has the solution Q(t) = Ce it . The free constant C = i is found from the normalizationcondition Q(0) ∗ i = 1, where ∗ means complex-conjugate.Alternatively, on the circle z(t) = e it , we have f(z(t)) = f(e it ) = ie it . From the normalizationcondition Q(t) ∗ f(z(t)) = 1 we f<strong>in</strong>d Q(t) = ie it = − s<strong>in</strong> ϑ + i cos ϑ.17. Rescal<strong>in</strong>g the state variable z = √ bu and the time, τ = ɛt, we obta<strong>in</strong> the reduced systemu ′ = (1 + i)u + (−1 + di)u|u| 2 + εp(t) .We can apply the theory only when ε is small. That is, the theory is guaranteed to work <strong>in</strong> avery weak limit ɛ ≪ b √ b ≪ 1. As it is often the case, numerical simulations suggest that thetheory works well outside the guaranteed <strong>in</strong>terval. Substitut<strong>in</strong>g u = re iθ <strong>in</strong>to this equation,r ′ e iθ + re iθ iθ ′ = (1 + i)re iθ + (−1 + di)r 3 e iθ + εp(t) ,divid<strong>in</strong>g by e iθ and separat<strong>in</strong>g real and imag<strong>in</strong>ary terms, we represent the oscillator <strong>in</strong> polarcoord<strong>in</strong>atesr ′ = r − r 3 + ε Re p(t)e −iθθ ′ = 1 + dr 2 + ε Im r −1 p(t)e −iθ .When ε = 0, this system has a limit cycle attractor r(t) = 1 and θ(t) = (1 + d)t, provided thatd ≠ −1. On the attractor, the solution to Malk<strong>in</strong>’s adjo<strong>in</strong>t equation (10.10),(Q ′ −2 0= −2d 0) ⊤ (0Q with Q(t) ·1 + d)= 1 ,is Q(t) = (d, 1)/(1+d). Indeed, the normalization condition results <strong>in</strong> Q 2 (t) = 1/(1+d). Hence,unique periodic solution of the first equation, Q ′ 1 = 2Q 1 − 2d/(1 + d), is Q 1 (t) = d/(1 + d). Onecan also use Kuramoto’s approach and the results of Ex. 2. The correspond<strong>in</strong>g phase model,ϑ ′ = 1 + ε{d Re p(t)e −i(1+d)ϑ + Im p(t)e −i(1+d)ϑ }/(d + 1) ,can be simplified via θ = (1 + d)ϑ (notice the font difference) to get the result.18. (Delayed coupl<strong>in</strong>g) Let ϑ(t) = t + ϕ(τ), where τ = εt is the slow time. S<strong>in</strong>ce ϑ(t − d) =t − d + ϕ(τ − εd) = t − d + ϕ(τ) + O(ε), we have x j (ϑ i (t − d ij )) = x j (t − d ij + ϕ(τ)) so thatwe can proceed as <strong>in</strong> Sect. 10.2.5 except that there is an extra term, −d ij , <strong>in</strong> (10.16). See alsoIzhikevich (1998).


518 Solutions to Exercises, Chap. 1019. Let χ = ϕ 2 − ϕ 1 ; then we have˙χ = ω − c s<strong>in</strong> χ .If |ω/c| ≤ 1, then there are two synchronized states, χ = arcs<strong>in</strong> (ω/c) and χ = π − arcs<strong>in</strong> (ω/c),one stable and the other unstable.20. From the theorem by Hoppensteadt and Izhikevich (1997) presented <strong>in</strong> Sect. 10.3.3 it followsthat Kuramoto’s model is a gradient system when c ij = c ji and ψ ij = −ψ ji . From Ermentrout’stheorem presented <strong>in</strong> the same section, it follows that the synchronized state ϕ i = ϕ j is stableif, e.g., all ψ ij = 0 and c ij > 0.21. S<strong>in</strong>ce the probability density function g(ω) is symmetric, the averaged frequency deviation ofthe network is zero, and rotat<strong>in</strong>g the coord<strong>in</strong>ate system, we can make the cluster phase ψ = 0.The network is split <strong>in</strong>to two populations: One oscillat<strong>in</strong>f with the cluster (|ω| < Kr), therebyform<strong>in</strong>g the cluster, and one drift<strong>in</strong>g <strong>in</strong> and out of the cluster. The latter does not contributeto the Kuramoto synchronization <strong>in</strong>dex, because contributions from different oscillators canceleach other on average. In the limit n → ∞, the sum (10.21) becomes the <strong>in</strong>tegral∫∫r = e iϕ(ω) g(ω)d ω ≈ e iϕ(ω) g(ω)d ω .|ω| 0 .(x)25. (Brown et al. 2004) The solution of x ′ = λx with x(0) = x 0 is x(t) = x 0 e λt . The period T =log(∆/x 0 )/λ is found from the condition x(T ) = ∆. Hence, Q(ϑ) = 1/(λx(ϑ)) = 1/(λx 0 e λϑ ) =e λ(T −ϑ) /(∆λ).26. Let us first consider the SNIC case ẋ = 1 + x 2 . The solution star<strong>in</strong>g with x(0) = x reset hasthe form (check by differentiat<strong>in</strong>g) x(t) = tan(t + t 0 ), where t 0 = atan x reset . The periodshould be found from the condition tan(T + t 0 ) = +∞, and it is T = π/2 − t 0 . Hence, x(t) =tan(t+π/2−T ) = − cot(t−T ). Now, Q(ϑ) = 1/(1+x(ϑ) 2 ) = 1/(1+cot 2 (ϑ−T )) = s<strong>in</strong> 2 (ϑ−T ).The homocl<strong>in</strong>ic case ẋ = −1 + x 2 is quite similar. The solution star<strong>in</strong>g with x(0) = x reset > 1has the form (check by differentiat<strong>in</strong>g) x(t) = − coth(t + t 0 ), where t 0 = acoth(−x reset ) = −acoth(x reset ). The period is found from the condition − coth(T +t 0 ) = +∞ result<strong>in</strong>g <strong>in</strong> T = −t 0 .Hence, x(t) = − coth(t − T ). F<strong>in</strong>ally, Q(ϑ) = 1/(−1 + x(ϑ) 2 ) = 1/(1 + coth 2 (ϑ − T )) =s<strong>in</strong>h 2 (ϑ − T ).

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!