P. C. Bressloff et al- Oscillatory waves in inhomogeneous neural media

  • Upload
    neerfam

  • View
    213

  • Download
    0

Embed Size (px)

Citation preview

  • 8/3/2019 P. C. Bressloff et al- Oscillatory waves in inhomogeneous neural media

    1/4

    Oscillatory waves in inhomogeneous neural media

    P. C. Bressloff1, S. E. Folias1, A. Prat2 and Y-X Li21Department of Mathematics, University of Utah, Salt Lake City, UT 84112

    2Departments of Mathematics and Zoology, University of British Columbia, Vancouver, BC, Canada V6T 1Z2(Dated: September 3, 2003)

    In this Letter we show that an inhomogeneous input can induce wave propagation failure in anexcitatory neural network due to the pinning of a stationary front or pulse solution. A subsequent

    reduction in the strength of the input can lead to a Hopf instability of the stationary solutionresulting in breather-like oscillatory waves.

    PACS numbers: 87.*, 05.45.-a, 05.45.Xt

    A number of theoretical studies have established theoccurrence of traveling fronts [1, 2] and traveling pulses[35] in one-dimensional excitatory neural networks mod-eled in terms of evolution equations of the form

    u(x, t)

    t+ u(x, t) =

    w(x|x)f(u(x, t))dx

    v(x, t) + I(x)

    1

    v(x, t)

    t+ v(x, t) = u(x, t) (1)

    where u(x, t) is a neural field that represents the localactivity of a population of excitatory neurons at positionx R, I(x) is an external input, is a synaptic timeconstant (assuming firstorder or exponential synapses),f(u) denotes an output firing rate function and w(x|x)is the strength of connections from neurons at x to neu-rons at x. The neural field v(x, t) represents some formof negative feedback recovery mechanism such as spikefrequency adaptation or synaptic depression, with , determining the relative strength and rate of feedback.

    (One can also incorporate higherorder synaptic and den-dritic processes by replacing u/t + u with a moregeneral linear differential operator Lu). It has been es-tablished [5] that there is a direct link between the abovemodel and experimental studies of wave propagation incortical slices where synaptic inhibition is pharmacologi-cally blocked [68]. Since there is strong vertical couplingbetween cortical layers, it is possible to treat a thin corti-cal slice as an effective onedimensional medium. Analy-sis of the model provides valuable information regardinghow the speed of a traveling wave, which is relativelystraightforward to measure experimentally, depends onvarious features of the underlying cortical circuitry.

    One of the basic assumptions in the analysis of travel-ing wave solutions of equation (1) is that the system isspatially homogeneous, that is, the external input I(x) isindependent of x and the synaptic weights depend onlyon the distance between pre-synaptic and post-synapticcells, w(x|x) = w(xx) with w a monotonically decreas-ing function of cortical separation. It can then be estab-lished that waves are in the form of traveling fronts in theabsence of any feedback, whereas traveling pulses tend tooccur when there is significant feedback [5]. However, the

    cortex is more realistically modeled as an inhomogeneousmedium. For example, inhomogeneities in the synapticweight distribution w are likely to arise due to the patchynature of long-range horizontal connections in superficiallayers of cortex [9]. Another important source of inho-mogeneity arises from external inputs induced by sensorystimuli, which may be modeled in terms of a nonuniforminput I(x). In this Letter we show that for appropriatechoices of input inhomogeneity, wave propagation failurecan occur due to the pinning of a stationary front or pulsesolution. More significantly, we find that these stationarysolutions can undergo a Hopf instability at a critical in-put amplitude, below which an oscillatory back-and-forthpattern of wave propagation or breather is observed.Our analysis predicts that the Hopf frequency dependson the relative strength and rate of feedback, but is in-dependent of the details of the weight distribution. Wealso show numerically how a secondary instability leadsto the generation of traveling waves. Analogous breather-like solutions have been found in inhomogeneous reaction

    diffusion systems [10, 11] and in numerical simulations ofa realistic model of fertilization calcium waves [12].First, let us consider traveling front solutions of equa-

    tion (1) in the case of zero input I(x) = 0 and ho-mogeneous weights w(x|x) = w(x x). For mathe-matical convenience, we take w(x) = (2d)1e|x|/d with

    w(y)dy = 1. The time and length scales are fixed bysetting = d = 1; typical values for these parameters are = 10msec and d = 1mm. As a further simplification,let f(u) = (u ) where is the Heaviside step func-tion and is a threshold. We then seek a traveling frontsolution of the form u(x, t) = U(), = x ct, c > 0,such that U(0) = , U() < for > 0 and U() >

    for < 0. The center of the wave is arbitrary due to thetranslation symmetry of the homogeneous system. Elim-inating the variable V() = v(x ct) by differentiatingequation (1) twice with respect to , leads to the secondorder differential equation

    c2U() + c[1 + ]U() [1 + ]U()

    = cw() W() (2)

    where denotes differentiation with respect to andW() =

    w(y)dy. The boundary conditions are

  • 8/3/2019 P. C. Bressloff et al- Oscillatory waves in inhomogeneous neural media

    2/4

    2

    U(0) = and U() = U. Here U are the homo-geneous fixed point solutions U = 1/1 + , U+ = 0. Itfollows that a necessary condition for the existence of afront solution is < U. The speed of a traveling frontsolution (if it exists) can then be obtained by solving theboundary value problem in the domains 0 and 0and matching the solutions at = 0. This leads to thebifurcation scenarios shown in figure 1. Such bifurca-

    tions also occur when the Heaviside output function isreplaced by a smooth sigmoid function, which can thenbe analyzed using perturbation methods, and for moregeneral monotonically decreasing weight distributions w[13]. Note that the bifurcation of the stationary frontshown in figure 1(a) is analogous to the front bifurcationstudied in reactiondiffusion equations, also known as thenonequilibrium Ising-Bloch (NIB) transition [10, 1416].Front bifurcations are of general interest, since they formorganizing centers for a variety of nontrivial dynamics in-cluding the formation of breathers in the presence of weakinput inhomogeneities (see below).

    0 0.5 1 1.5-1

    -0.5

    0

    0.5

    1

    0 0.5 1 1.5

    -2

    -1

    0

    1

    speedc

    = 1.0

    = 1.5

    (a) (b)

    speedc

    FIG. 1: Plot of wavefront speed c as a function of for fixed and = 0.25. Stable (unstable) branches are shown assolid (dashed) curves. (a) If 2(1 + ) = 1 then there existsa stationary front for all ; at a critical value of the sta-tionary front loses stability and bifurcates into a left and a

    right moving wave (b) If 2(1 + ) > 1 then there is a singleleft-moving wave for all and a pair of right-moving wavesthat annihilate in a saddle-node bifurcation. Left and rightmoving waves are reversed when 2(1 + ) < 1 (not shown).

    In the case of an inhomogeneous input, wave propaga-tion failure can occur due to the formation of a stable sta-tionary front solution. Stationary front solutions of equa-tion (1) for homogeneous weights and f(u) = (u )satisfy the equation

    (1 + )U(x) = x0

    w(x x)dx + I(x) (3)

    Suppose that I(x) is a monotonically decreasing functionof x. Since the system is no longer translation invariant,the position of the front is pinned to a particular locationx0 where U(x0) = . Monotonicity of I(x) ensures thatU(x) > for x < x0 and U(x) < for x > x0. Thecenter x0 satisfies (1 + ) = 1/2 + I(x0), which impliesthat in contrast to the homogeneous case, there exists astationary front over a range of threshold values (for fixed); changing the threshold simply shifts the position

    of the center x0. If the stationary front is stable then itwill prevent wave propagation. Stability is determinedby writing u(x, t) = U(x) + p(x, t) and v(x, t) = V(x) +q(x, t) and expanding equation (1) to first-order in (p, q):

    p(x, t)

    t= p(x, t) q(x, t)

    +

    w(x x)H(U(x))p(x, t)dx

    1

    q(x, t)

    t= q(x, t) + p(x, t) (4)

    The spectrum of the associated linear operator is foundby taking p(x, t) = etp(x) and q(x, t) = etq(x). Usingthe identity H(U(x)) = (xx0)/|U

    (x0)|, we obtainthe equation

    ( + 1)p(x) =w(x x0)

    |U(x0)|p(x0)

    p(x)

    + (5)

    Equation (5) has two classes of solution. The first con-sists of any function p(x) such that p(x0) = 0, for

    which the corresponding eigenvalues always have nega-tive real part. The second consists of solutions of theform p(x) = Aw(x x0), A = 0, for which the corre-sponding eigenvalues are

    =

    2 4(1 )(1 + )

    2(6)

    where

    = 1 + (1 + ), =1

    1 + 2D(7)

    with D = |I(x0)|. We have used the fact that I(x0) 0

    and w(0) = 1/2.Equation (6) implies that the stationary front (if it ex-ists) is locally stable provided that > 0 or, equivalently,the gradient of the inhomogeneous input at x0 satisfies

    D > Dc 1

    2

    1 + (8)

    Since D 0, it follows that the front is stable when > , that is, when the feedback is sufficiently weak orfast. On the other hand, if < then there is a Hopf bi-furcation at the critical gradient D = Dc. Consider as anexample the step inhomogeneity I(x) = (s/2) tanh(x),where s is the size of the step and determines its

    steepness. A stationary front will exist provided thats > s |1 2(1 + )|. The gradient D depends onx0, which is itself dependent on and . On eliminatingx0, we can write D = (s

    2 s2)/2s. Substituting intoequation (8) yields an expression for the critical value ofs that determines the Hopf bifurcation points:

    sc =1

    2

    1 + +

    1 +

    2+ 4s22

    . (9)

  • 8/3/2019 P. C. Bressloff et al- Oscillatory waves in inhomogeneous neural media

    3/4

    3

    0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 20

    0.2

    0.4

    0.60.8

    1

    1.2

    1.4

    1.6

    1.8

    2

    s

    = 0.5 = 1.0

    = 1.5

    FIG. 2: Stability phase diagram for a stationary front in thecase of a step input I(x) = s tanh(x)/2 where is thesteepness of the step and s its height. Hopf bifurcation lines(solid curves) in s parameter space are shown for variousvalues of . In each case the stationary front is stable above

    the line and unstable below it. The shaded area denotes theregion of parameter space where a stationary front solutiondoes not exist. The threshold = 0.25 and = 0.5.

    -20 200

    space x (in units of d)

    0

    100

    180

    time

    t(inunitsof)

    0

    0.8

    0.6

    0.4

    0.2

    activity u

    FIG. 3: Breather-like solution arising from a Hopf instabilityof a stationary front due to a slow reduction in the size s ofa step input inhomogeneity and exponential weights. Here = 0.5, = 0.5, = 1, = 0.25. The input amplitude s = 2at t = 0 and s = 0 at t = 180. The amplitude of the oscillationsteadily grows until it destabilizes at s 0.05, leading to the

    generation of a traveling front.

    The critical height sc is plotted as a function of for = 0.5 and various values of in Fig. 2. Note thatclose to the front bifurcation = a Hopf bifurcationoccurs in the presence of a weak inhomogeneity. Numer-ically one finds that reducing the input amplitude be-low the critical point induces a transition to a breather

    like oscillatory front solution, whose frequency of oscilla-tion is approximately equal to the critical Hopf frequencyH =

    ( ). This suggests that the bifurcation is

    supercritical. Note that the frequency of oscillations onlydepends on the size and rate of the negative feedback,but is independent of the details of the synaptic weightdistribution. As the input amplitude is further reduced,the breather itself becomes unstable and there is a sec-

    ondary bifurcation to a traveling front. This is illustratedin Fig. 3, which shows a space-time plot of the developingbreather as the input amplitude is slowly reduced.

    The above analysis can be extended to the case of sta-tionary pulse solutions in the presence of a unimodal in-put I(x) which, for concreteness, is taken to be a Gaus-

    sian of width centered at the origin I(x) = Iex2/22 .

    From symmetry arguments there exists a stationary pulsesolution U(x) of equation (1) centered at x = 0 withU(a/2) = and U() = 0:

    (1 + )U(x) = a/2

    a/2

    w(x x)dx + I(x) (10)

    The threshold and width a are related according to

    (1 + ) =

    I(a/2) +

    1 ea

    2

    G(a) (11)

    Plotting the function G(a) for a range of input ampli-tudes I, it can be shown that for (1 + ) < 0.5 thereexists a single pulse solution over the finite range of in-puts 0 I (1 + ), and no pulse solutions when

    I > (1 + ). On the other hand, when (1 + ) > 0.5there exist two solution branches as illustrated in Fig. 4,

    one corresponding to a narrow pulse and the other to abroad pulse. These two branches coalesce at the criticalpoint I = ISN where G(a) = (1 + ) and G(a) = 0.

    1 1.50

    1

    2

    3

    amplitude I

    widtha

    SN

    amplitude I

    HB

    (a)

    >

    1 1.5

    (b)

    <

    FIG. 4: Plot of pulse width a as a function of input amplitudeI obtained by numerically solving equation (11) for = 1, = 0.5 and = 1. The lower branch is unstable whereasthe upper branch is stable for large pulse width. (a) If > then the upper branch undergoes a saddle-node bifurcationat I = ISN. (b) If < then the upper branch undergoes aHopf bifurcation at I = IHB > ISN.

  • 8/3/2019 P. C. Bressloff et al- Oscillatory waves in inhomogeneous neural media

    4/4

    4

    Carrying out a linear stability analysis along similarlines to the case of a front leads to the following stabilityresults [13]: (i) The single pulse solution for (1+) < 0.5is unstable. (ii) The lower branch of solutions (narrowpulse) for (1 + ) > 0.5 is always unstable, whereas theupper branch (broad pulse) is stable for sufficiently largepulse width a. (iii) If > then the upper branch isstable for all I > ISN and undergoes a saddle node bi-

    furcation at I = ISN. (iv) If < then there exists acritical input amplitude IHB with IHB > ISN such thatthe upper branch is stable for I > IHB and undergoesa Hopf bifurcation at I = IHB . Numerically we findthat the Hopf instability of the upper branch induces abreather-like oscillatory pulse solution as illustrated inFig. 5. One finds that the associated Hopf frequency isagain given by H =

    ( ), which is independent

    of the pulsewidth a. For the parameter values used inFig. 5, we have 0.251 = 25Hz assuming that = 10msec. As the input amplitude I is slowly reducedbelow IHB , the oscillations steadily grow until a newinstability point is reached. Interestingly, the breatherpersists over a range of inputs beyond this secondary in-stability except that it now periodically emits pairs oftraveling pulses. Furthermore, in this parameter regimewe observe frequency-locking between the oscillations ofthe breather and the rate at which pairs of pulses areemitted from the breather. Note that although the ho-mogeneous network (I = 0) also supports the propaga-tion of traveling pulses, it does not support the existenceof a breather that can act as a source of these waves.

    -10 100

    space x (in units of d)

    0

    100

    200

    timet(inunitsof)

    0.2

    1.0

    0.8

    0.6

    0.4

    activity u

    0

    1.2

    FIG. 5: Breather-like solution arising from a Hopf instabilityof a stationary pulse due to a slow reduction in the amplitudeI of a Gaussian input and exponential weights. Here I = 5.5at t = 0 and I = 1.5 at t = 250. Other parameter valuesare = 0.03, = 2.5, = 0.3, = 1.0. The amplitude ofthe oscillation steadily grows until it undergoes a secondaryinstability at I 2, beyond which the breather persists andperiodically generates pairs of traveling pulses (only one ofwhich is shown). The breather itself disappears when I 1.

    Two major predictions of our analysis are (i) an inho-mogeneous input current can induce oscillatory behav-ior in the form of breathing fronts and pulses and (ii)the oscillation frequency is approximately independentof the details of the underlying synaptic weight distribu-tion, depending only on parameters that have a directbiological interpretation in terms of single cell recoverymechanisms. From an experimental perspective, our re-

    sults could be tested by introducing an inhomogeneouscurrent into a cortical slice and searching for these oscil-lations. One potential difficulty of such an experiment isthat persistent currents tend to burn out neurons. In thecase of traveling fronts, this might be avoided by oper-ating the system close to the front bifurcation of the ho-mogeneous network, see figure 1(a), such that only weakinhomogeneities would be needed to induce oscillations.An alternative approach might be to use some form ofpharmacological manipulation of NMDA receptors, forexample. Note that the usual method for inducing trav-eling waves in cortical slices (and in corresponding com-putational models) is to introduce short-lived current in-

    jections; once the wave is formed it propagates in a ho-mogeneous medium (neglecting the modulatory effects oflongrange horizontal connections [9]). In future work wewill generalize our results to the case of a smooth outputnonlinearity f and determine to what extent the oscil-lation frequency now depends on the form of the weightdistribution w. We will also consider extensions to targetwaves in twodimensional networks [13].

    [1] G. B. Ermentrout and J. B. Mcleod, Proc. Roy. Soc.

    Edin. A 123, 461 (1993).[2] M. A. P. Idiart and L. F. Abbott, Network 4, 285 (1993).[3] H. R. Wilson and J. D. Cowan, Kybernetik 13, 55 (1973).[4] S. Amari, Biol. Cybern. 27, 77 (1977).[5] D. Pinto and G. B. Ermentrout, SIAM J. Appl. Math

    62, 206 (2002).[6] R. D. Chervin, P. A. Pierce, and B. W. Connors, J. Neu-

    rophysiol. 60, 1695 (1988).[7] D. Golomb and Y. Amitai, J. Neurophysiol. 78, 1199

    (1997).[8] J.-Y. Wu, L. Guan, and Y. Tsau, J. Neurosci. 19, 5005

    (1999).[9] P. C. Bressloff, Physica D 155, 83 (2001).

    [10] M. Bode, Physica D 106, 270 (1997).

    [11] A. Prat and Y.-X. Li (2003), submitted for publication.[12] Y.-X. Li (2003), submitted for publication.[13] S. E. Folias and P. C. Bressloff (2003), unpublished.[14] J. Rinzel and D. Terman SIAM J. Appl. Math. 42, 1111

    (1982).[15] A. Hagberg and E. Meron Nonlinearity 7, 805 (1994).[16] A. Hagberg and E. Meron and I. Rubinstein and B. Zalt-

    man, Phys. Rev. Lett. 76, 427 (1996).