• Aucun résultat trouvé

Nonlinear analysis of periodic waves in a neural field model

N/A
N/A
Protected

Academic year: 2021

Partager "Nonlinear analysis of periodic waves in a neural field model"

Copied!
24
0
0

Texte intégral

(1)

HAL Id: hal-03019102

https://hal.archives-ouvertes.fr/hal-03019102

Submitted on 23 Nov 2020

HAL is a multi-disciplinary open access

archive for the deposit and dissemination of

sci-entific research documents, whether they are

pub-lished or not. The documents may come from

teaching and research institutions in France or

abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est

destinée au dépôt et à la diffusion de documents

scientifiques de niveau recherche, publiés ou non,

émanant des établissements d’enseignement et de

recherche français ou étrangers, des laboratoires

publics ou privés.

model

S Budzinskiy, A Beuter, V Volpert

To cite this version:

(2)

Nonlinear analysis of periodic waves in a neural field model

S. Budzinskiy,1, 2, 3 A. Beuter,4, 5,a)and V. Volpert6, 7, 3

1)Faculty of Computational Mathematics and Cybernetics, Lomonosov Moscow State University, Leninskie Gory 1,

119991 Moscow, Russia

2)Marchuk Institute of Numerical Mathematics RAS, Gubkina St. 8, 119333 Moscow,

Russia

3)Peoples’ Friendship University of Russia (RUDN University), Miklukho-Maklaya St. 6, 117198 Moscow,

Russia

4)Bordeaux INP, Bordeaux, France 5)CorStim SAS, Montpellier, France

6)Institut Camille Jordan, UMR 5208 CNRS, University Lyon 1, 69622 Villeurbanne, France 7)INRIA Team Dracula, INRIA Lyon La Doua, 69603 Villeurbanne, France

(Dated: 5 August 2020)

Various types of brain activity, including motor, visual, and language, are accompanied by the propagation of periodic waves of electric potential in the cortex, possibly providing the synchronization of the epicenters involved in these activities. One example is cortical electrical activity propagating during sleep and described as traveling waves1. These waves modulate cortical excitability as they progress. Clinically-related examples include cortical spreading depression in which a wave of depolarization propagates in migraine but also in stroke, hemorrhage or traumatic brain injury2. Here we consider the possible role of epicenters and explore a neural field model with two nonlinear integro-differential equations for the distributions of activating and inhibiting signals. It is studied with symmetric connectivity functions characterizing signal exchange between two populations of neurons, excitatory and inhibitory. Bifurcation analysis is used to investigate the emergence of periodic traveling waves and of standing oscillations from the stationary, spatially homogeneous solutions, and the stability of these solutions. Both types of solutions can be started by local oscillations indicating a possible role of epicenters in the initiation of wave propagation.

The concept of periodic traveling waves appeared in the early 1970’s and since that time investigators have actively explored the mechanisms, computational principles and functional role of these waveforms3,4. This paper focuses on how modeling information can be used to guide electri-cal stimulation in the targeted manipulation of brain net-works in human patients affected by neurological disor-ders. We explore a neural field model with two nonlinear integro-differential equations and perform a bifurcation analysis. Results suggest that it is possible to determine the direction of propagation and that epicenters may play a role in the initiation of wave propagating in the cortex. These results will guide the determination of the optimal localization to apply external brain stimulation (forcing term) to restore wave propagation in damaged cortical tis-sue.

I. INTRODUCTION

All types of brain activity are accompanied by formation of spatiotemporal patterns of electric potential in the cerebral cortex, in particular, wave fronts and periodic waves5–7. They provide subthreshold depolarization to individual neurons and increase their spiking probability. They possibly synchronize different parts of the brain and organize spatial phase distri-butions in a population of neurons8. Traveling waves can

a)Electronic mail: anne.beuter@wanadoo.fr

also be detected from the phase relations between oscillations recorded in different cortical regions9.

EEG recordings from the cortical surface in awake neuro-surgical patients performing a memory task can show focal re-gions or clusters of 2-15 Hz oscillations10. These oscillations

formed traveling waves that propagated at about 0.25-0.75 m/s and correlated with the subject’s performance. In other words, they guide the spatial propagation of neural activity demon-strating large scale spatially coordinated oscillations.

The relevance of traveling waves is actively explored in patients with stroke, subarachnoid hemorrhage or traumatic brain injury. The occurrence of spreading depression or de-polarization in these pathologies is a well documented phe-nomenon. Spreading depression or depolarization of neu-rons and glial cells is also reported in patients with migraine2. However, the relevance of traveling waves is also explored in healthy subjects. Indeed traveling waves have been observed in processes involving cognition (such as perception, learning, recognition, memory, language) (see for example Ref. 10 for working memory or Ref. 11 for language) but also in motor behaviors9,12or vision13.

For each human subject the examination of slow waves of cortical activity during sleep-awake states using simultaneous scalp EEG and intracranial recordings allowed the identifica-tion of the set of intracranial contacts involved in a larger per-centage of detected events14. These contacts are called «hubs»

because the wave had large probability of passing through the region close to the contact. Using probabilities they were able to reconstruct a preferential propagation network for each sub-ject. Slow waves have been reported to propagate across cor-tical areas at about 1 m/s with multiple propagation paths and several points of origin14. These waves appear to shape and

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

(3)

strengthen neuronal networks. There are numerous other evi-dences of wave propagation during brain activity11,15.

There are two main classes of cortical activity which are de-scribed by neural mass models and neural field models (see, e.g., Ref. 16). The former is based on a discrete represen-tation of brain networks with spatially homogeneous neural elements connected to each other. In neural field models, cerebral cortex is considered as a continuous medium with a nonlocal interaction due to neuron connectivity. Such mod-els were first introduced in Ref. 17 followed by many others for one or two neuron populations18–20, discrete or distributed time delay20,21, linear adaptation22, refractoriness23, and other questions22,24–27.

In this work we consider the model ∂ u ∂ t = P1,1∗ ψ1(u) − P1,2∗ ψ2(v) − σ u, ∂ v ∂ t = P2,1∗ ψ1(u) − P2,2∗ ψ2(v) − σ v, (1) where Pi, j(x) = ai, je−bi, j|x|, i, j = 1, 2,

ai, j, bi, j, and σ are positive constants, ψ1(u) and ψ2(v) are

bounded smooth growing (sigmoid, i.e. having an S-shaped curve) functions on the whole real axis, ∗ denotes convolution. The variables u and v characterize the activity of the neural population of excitatory and inhibitory neurons, respectively. System (1) represents a particular case without time delay of the model studied in Ref. 28. Considered on the whole real axis, under some natural conditions, it has a stationary, spa-tially homogeneous solution (u0, v0). This solution can be

stable or unstable depending on the values of parameters. In particular, it can lose its stability due to a simple real eigen-value or a pair of complex conjugate eigeneigen-values crossing the imaginary axis. Linear stability analysis of the stationary ho-mogeneous solution was carried out in Ref. 28. Here we are interested in nonlinear bifurcation analysis allowing the de-termination of stability of the bifurcating solutions. We will obtain the conditions providing the existence and stability of standing and periodic traveling waves.

Furthermore, we are interested in the initiation of the pe-riodic waves by space localized oscillations. From the bio-logical point of view, they correspond to the oscillations in the hubs, which process received information and send it to other hubs in the form of traveling waves. We show that wave initiation occurs depending on the properties of the localized oscillations.

II. SPECTRUM OF THE LINEARIZED EQUATIONS

Let (u0(σ ), v0(σ )) be a pair of constants satisfying

σ u0= 2 a1,1 b1,1 ψ1(u0) − 2 a1,2 b1,2 ψ2(v0), σ v0= 2 a2,1 b2,1 ψ1(u0) − 2 a2,2 b2,2 ψ2(v0).

Then (u0(σ ), v0(σ )) are solutions to (1). By shifting the

func-tions

u(x,t) = u0(σ ) + ˜u(x,t), v(x,t) = v0(σ ) + ˜v(x,t),

linearizing the equations (1), and dropping the tildes, we get ∂ u ∂ t = ψ 0 1(u0(σ ))P1,1∗ u − ψ20(v0(σ ))P1,2∗ v − σ u(x,t), ∂ v ∂ t = ψ 0 1(u0(σ ))P2,1∗ u − ψ20(v0(σ ))P2,2∗ v − σ v(x,t).

Looking for solutions in the form of

u(x,t) = exp(λ t)u(x), v(x,t) = exp(λ t)v(x) and after applying the Fourier transform, we obtain the char-acteristic equations ψ10(u0(σ )) ˆP1,1(ξ ) ˆu(ξ ) − ψ20(v0(σ )) ˆP1,2(ξ ) ˆv(ξ ) = (σ + λ ) ˆu(ξ ), ψ10(u0(σ )) ˆP2,1(ξ ) ˆu(ξ ) − ψ20(v0(σ )) ˆP2,2(ξ ) ˆv(ξ ) = (σ + λ ) ˆv(ξ ), where ˆ Pi, j(ξ ) = 2ai, jbi, j b2i, j+ ξ2. Let µ = µ(ξ ; σ ) = σ + λ (ξ ; σ ) ∈ C and ˆ Qi,1(ξ ; σ ) = ψ10(u0(σ )) ˆPi,1(ξ ), ˆ Qi,2(ξ ; σ ) = ψ20(v0(σ )) ˆPi,2(ξ ).

Then the characteristic polynomial becomes µ2− ˆQ1,1(ξ ) − ˆQ2,2(ξ )µ −

− ˆQ1,1(ξ ) ˆQ2,2(ξ ) − ˆQ1,2(ξ ) ˆQ2,1(ξ ) = 0

with two roots µ (ξ ) =1 2 ˆ Q1,1(ξ ) − ˆQ2,2(ξ ) ± ±1 2 q ˆ Q1,1(ξ ) + ˆQ2,2(ξ ) 2 − 4 ˆQ1,2(ξ ) ˆQ2,1(ξ ).

If the cross-connectivity functions P1,2(x) and P2,1(x) are

strong enough (their amplitudes a1,2 and a2,1 are large

enough), we can ensure that the expression under the square root is always negative

ˆ

Q1,1(ξ ) + ˆQ2,2(ξ ) 2

−4 ˆQ1,2(ξ ) ˆQ2,1(ξ ) < 0, −∞ < ξ < ∞,

which leads to complex values of µ(ξ ) such that Reµ(ξ ) =1 2 Qˆ1,1(ξ ) − ˆQ2,2(ξ ) , Imµ(ξ ) = ±1 2 q 4 ˆQ1,2(ξ ) ˆQ2,1(ξ ) − ˆQ1,1(ξ ) + ˆQ2,2(ξ ) 2 . Consider a function f (ξ ) : R → R that is related to the real part of the characteristic values µ(ξ ):

f(ξ ) = α1 β1+ ξ2

− α2 β2+ ξ2

, αi> 0, βi> 0.

We are interested in its extreme values and the corresponding frequency ξ where they are attained.

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

(4)

Lemma 1. If α1= α2or β1= β2then| f (ξ )| ≤ | f (0)| for all

ξ .

Proof. Let α1= α2= α. Then

f(ξ ) = α (β2− β1) (β1+ ξ2)(β2+ ξ2)

.

Unless β1= β2 and f (ξ ) ≡ 0, | f (ξ )| is monotonically

de-creasing on ξ > 0 and inde-creasing on ξ < 0. Now let β1= β2= β . Then

f(ξ ) =α1− α2 β + ξ2.

Unless α1= α2 and f (ξ ) ≡ 0, | f (ξ )| is monotonically

de-creasing on ξ > 0 and inde-creasing on ξ < 0. Let α16= α2and β16= β2; we introduce

β3=

α1β2− α2β1

α1− α2

and note that it can take any real value, not necessarily positive ones. Then the derivative of f (ξ ) is

f0(ξ ) = 2ξ (α1− α2)× ×(β1+ ξ 2)(β 2+ ξ2) − (β3+ ξ2)(β1+ β2+ 2ξ2) (β1+ ξ2)2(β2+ ξ2)2 . It vanishes at ξ = 0 and, possibly, at

ξ2= −β3±

p

(β3− β1)(β3− β2) ≡ η±

if the right hand side is real and positive. Lemma 2. Both η+and η−are real whenever

β3≤ min(β1, β2) or β3≥ max(β1, β2).

The value η−is negative when it is real, i.e. whenever

β3≤ min(β1, β2) or β3≥ max(β1, β2).

The value η+is nonpositive whenever

β1β2

β1+ β2

≤ β3≤ min(β1, β2) or β3≥ max(β1, β2).

Proof. The first statement is obvious. Let us assume that η+, η−∈ R. For the second statement,

1. if β3≥ 0 then η−< 0 if and only if

β3> −

p

(β3− β1)(β3− β2),

so 0 ≤ β3≤ min(β1, β2) or β3≥ max(β1, β2);

2. if β3< 0 then η−< 0 if and only if

β1β2− β3(β1+ β2) > 0,

so simply β3< 0.

For the third statement, η+≤ 0 if and only if

β1β2− β3(β1+ β2) ≤ 0

whence the result follows.

Lemma 3. Let α16= α2, β16= β2, and β3< β1β2/(β1+ β2).

The derivative f0(ξ ) vanishes at only 3 points:

ξ = 0 and ξ = ± q

−β3+

p

(β3− β1)(β3− β2).

Proof. Since α16= α2and β16= β2, f0(ξ ) turns to zero if and

only if ξ = 0 or ξ2= η±. Lemma 2 states that η−< 0 and

η+> 0 when β3< β1β2/(β1+ β2). So f0(ξ ) = 0 at ξ = 0 and ξ = ±√η+. Since f0(ξ ) = 2(α1− α2) β1β2− β3(β1+ β2) β122 ξ + O(ξ 2)

in the vicinity of the origin, the function f (ξ ) has global max-ima at

ξ2= −β3+

p

(β3− β1)(β3− β2)

when α1> α2, and global minima when α1< α2.

Let

α1= ψ10(u0(σ ))a1,1b1,1, β1= b21,1

and

α2= ψ20(v0(σ ))a2,2b2,2, β2= b22,2,

then Reµ(ξ ) = f (ξ ) and the largest real part supξReµ(ξ ) is attained at ±ξ∗(σ ): ξ∗(σ ) = r −β3(σ ) + q (β3(σ ) − b21,1)(β3(σ ) − b22,2), β3(σ ) = ψ10(u0(σ ))a1,1b1,1b22,2− ψ 0 2(v0(σ ))a2,2b2,2b21,1 ψ10(u0(σ ))a1,1b1,1− ψ20(v0(σ ))a2,2b2,2 .

Thus supξReλ (ξ ; σ ) = Reλ (ξ∗(σ )) = f (ξ∗(σ )) − σ .

Theorem 1. Let a12, a21 1 and

b22 b11 <ψ 0 1(u0(σ ))a11 ψ20(v0(σ ))a22 < b11 b22 3 .

Then the spectrum is complex Imλ (ξ ; σ ) 6= 0 and Reλ (ξ ; σ ) < Reλ (ξ∗(σ ); σ ) for all ξ 6= ±ξ∗(σ ).

Proof. Follows from Lemmas 1-3 once we rewrite the con-straint β3< β1β2/(β1+ β2).

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

(5)

III. HOPF BIFURCATION AND NORMAL FORM ANALYSIS

Fix a value σ∗> 0 and let σ ∈ R be its small perturbation.

With a slight abuse of notation we will write f (σ ) for func-tions of σ that we would previously call f (σ∗+ σ ). Assume

that

Reλ (ξ ; 0) < Reλ (ξ∗(0); 0), ξ 6= ξ∗(0), (H1)

and

λ (ξ∗(0); 0) = ±iω∗6= 0. (H2)

Condition (H1) follows from Theorem 1 while the Hopf bifur-cation condition (H2) can be satisfied with a careful choice of parameters. We also assume the transversality condition

 d dσReλ (ξ∗(0); σ )  σ =0 = = −1 +1 2ψ 00 1(u0(0))u00(0) ˆP11(ξ∗(0))− −1 2ψ 00 2(v0(0))v00(0) ˆP22(ξ∗(0)) 6= 0. (H3)

That is the pair of imaginary eigenvalues passes the imaginary axis with nonzero speed.

Let ξ∗= ξ∗(0) and l∗= 2π/ξ∗. Consider the Lebesgue

space L2(0, l∗) with standard inner product and norm:

h f , giL2(0,l)= Z l∗ 0 f gdx, k f kL2(0,l)= q h f , f iL2(0,l).

The complex exponentials en=√1leinξ∗xform an orthonormal

basis in L2(0, L

∗). Also, consider the Sobolev space of l∗

-periodic twice-differentiable functions H2per(0, l∗) with inner

product h f , giH2

per(0,l∗)=

n∈Z

(1 + n4ξ∗4)h f , eniL2(0,l)hen, giL2(0,l)

and Euclidean norm. We consider smooth functions to make the analysis extendable to diffusive models.

Let w=u v  ∈ X = H2 per(0, l∗) 2

with inner product

hw, ˜wiX= hw1, ˜w1iH2

per(0,l∗)+ hw2, ˜w2iH2 per(0,l∗)

where wjand ˜wjare the components of two-dimensional

vec-tor functions.

Let Mk(σ ) ∈ R2×2be diagonal matrices

Mk(σ ) = 1 k! " ψ1(k)(u0(σ )) 0 0 −ψ2(k)(v0(σ )) #

andPn∈ R2×2be Fourier filtering matrices

Pn= ˆ P11(nξ∗) ˆP12(nξ∗) ˆ P21(nξ∗) ˆP22(nξ∗)  .

Any function w ∈ X ⊂ (L2(0, l∗))2can be represented by its

Fourier series w=

n∈Z wnen, wn= hw 1, eniL2(0,l) hw2, eniL2(0,l)  ∈ C2. We can rewrite the initial equations as

wt= L(σ )w + F(w, σ ), w∈ X, where L(σ )w =

n∈Z (PnM1(σ ) − σ∗− σ )wnen and F(w, σ ) =

k≥2n∈Z

PnMk(σ )(w k)nen.

We use to denote the Hadamard (element-wise) product of vectors and write w 2= w w, w 3= w w w, etc., for Hadamard products of w with itself.

The characteristic equation at σ = 0 decouples into (PnM1(0) − σ∗− λ )wn= 0, n∈ Z,

and λ ∈ C is a characteristic value if there exist n ∈ Z and wn6= 0 ∈ C2that satisfy the equation above. Owing to (H2),

there is a unique pair of imaginary characteristic values ±iω∗

corresponding to n = ±1. They are degenerate because the connectivity functions are symmetric and the system as a whole is O(2)-equivariant. Let ζ16= 0 ∈ C2be an eigenvector:

(P1M1(0) − σ∗− iω∗)ζ1= 0,

ζ1=ψ20(v0(0)) ˆP12(ξ∗) ψ10(u0(0)) ˆP11(ξ∗) − σ∗− iω∗T.

Consider vector functions

q1= ζ1e1, q2= ζ1e−1, q3= ζ1e1, q4= ζ1e−1.

In pairs, they span the eigenspaces

N (L(0) − iω∗) = span{q1, q2} ⊂ X,

N (L(0) + iω∗) = span{q3, q4} ⊂ X;

together they span the center subspace Q = span{q1, q2, q3, q4} ⊂ X.

The basis {qj} is not orthogonal so we will use a biorthogonal

system to get a projection operator ontoQ. Let ζ1∗6= 0 ∈ C2

be an eigenvector of the adjoint matrix (P1M1(0) − σ∗− iω∗)Hζ1∗= 0

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

(6)

normalized so that hζ1, ζ1∗i = 1. Further, let q∗1= 1 1 + ξ4 ∗ ζ1∗e1, q∗2= 1 1 + ξ4 ∗ ζ1∗e−1, q∗3= 1 1 + ξ4 ∗ ζ1∗e1, q∗4= 1 1 + ξ4 ∗ ζ1∗e−1

be the basis vectors of the eigenspaces N (L(0)∗ + iω∗) = span{q∗1, q ∗ 2} ⊂ X, N (L(0)∗− iω ∗) = span{q∗3, q ∗ 4} ⊂ X.

It is easy to show that {q∗j} and {qj} are biorthogonal systems

hqi, q∗jiX=

(

1, i= j, 0, i6= j. We can thus define a projection operator

π : X →Q, πw =

4

j=1

hw, q∗jiXqj

and a complementary subspace ˜

Q = {w ∈ X : πw = 0}. Then X =Q ⊕ ˜Q.

For a given w ∈ X , let z ∈ C4(note that z

4= z1and z2= z3

as the functions are real-valued) be the vector of coefficients of πw with respect to the basis {qj}. Let γ be the velocity of

the critical characteristic value: γ =  d dσλ (ξ∗; σ )  σ =0 .

According to the center manifold theorem, the equations for z1and z3on the manifold have the following normal form29:

˙z1= (iω∗+ γσ )z1+ c1z1|z1|2+ c2z1|z3|2+ h.o.t.,

˙z3= (iω∗+ γσ )z3+ c1z3|z3|2+ c2z3|z1|2+ h.o.t.,

where h.o.t. =O(σ(|z1|3+ |z3|3) + σ2(|z1| + |z3|) + (|z1| +

|z2|)5). See Appendix A for the computation of the cubic

co-efficients c1and c2of the normal form.

The relations between c1and c2determine the stability of

periodic traveling wave and standing wave solutions29. To see

this, we make a polar change of coordinates, z1= ρ1eiθ1, z3= ρ3eiθ3,

and keep two equations out of four: ˙

ρ1= ρ1(σ Reγ + Rec1ρ12+ Rec2ρ32) + h.o.t., ˙

ρ3= ρ3(σ Reγ + Rec1ρ32+ Rec2ρ12) + h.o.t..

(2)

These are the amplitude equations as ρ1and ρ3represent the

amplitudes of counterpropagating periodic traveling waves. The system (2) can have up to four equilibria (see Table I),

TABLE I: Equilibria of the amplitude equations (2).

Type of solution ρ1 ρ3

Constant 0 0

Periodic traveling wave

q −σ Reγ

Rec1 0

Periodic traveling wave 0

q −σ Reγ Rec1 Standing wave q − σ Reγ Rec1+Rec2 q − σ Reγ Rec1+Rec2

and the stability of these equilibria defines the stability of pe-riodic traveling waves and standing waves.

Let σ Reγ > 0 then

Rec1< 0, Re(c1+ c2) < 0, Re(c1− c2) < 0

leads to asymptotically stable standing waves while Rec1< 0, Re(c1+ c2) < 0, Re(c1− c2) > 0

produces asymptotically stable periodic traveling waves.

IV. ONE CLASS OF PARAMETERS

The net result of Sections II-III are the explicit expressions of the normal form coefficients, which can be easily evaluated for a given set of parameters. However, these expressions are too cumbersome to aid in the search of suitable parameters. Theorem 1 is also of merely theoretical interest: the condition a1,2, a2,1 1 which ensures that the whole spectrum lies in

C \ R is too vague to be verifiable. In this section, we will present and semi-rigorously analyze a particular class of pa-rameters that can be studied in more details.

Let ψ1= ψ2= ψ be a sigmoid function. In our analysis,

we will use a number of its properties such as

ψ0(0) > 0, ψ00(0) = 0, ψ000(0) < 0, ψ0000(0) = 0. Let R > 0 be the ratio of parameters

a1,2 b1,2 =a2,1 b2,1 =a2,2 b2,2 = R

and suppose that the ratio a1,1/b1,1is perturbed slightly

a1,1

b1,1

= R + ε, ε ∈ R. Next, assume that the parameters bi, jare equal

b1,1= b1,2= b2,1= b, b> 0,

except for b2,2that is smaller:

b2,2= νb, 0 < ν < 1.

Lemma 4. For all small values of ε, system (1) has constant solutions u0= 2 ψ (0) σ∗  1 + 2Rψ 0 (0) σ∗  ε +O(ε2), v0= 4R ψ (0) σ∗ ψ0(0) σ∗ ε +O(ε2).

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

(7)

Proof. At ε = 0, we have

σ∗u0= σ∗v0= 2R(ψ(u0) − ψ(v0))

and u0= v0= 0. By the inverse function theorem there is a

smooth branch of solutions (u0(ε), v0(ε)) whose first

deriva-tives satisfy the equation " σ∗− 2Rψ0(0) 2Rψ0(0) −2Rψ0(0) σ∗+ 2Rψ0(0) # "du 0 dε(0) dv0 dε(0) # = " 2ψ(0) 0 # . Note that the Jacobian matrix is full-rank for all admissible values of R, σ∗, and ψ0(0), being a sum of a scalar matrix and

a rank-1 matrix. Its inverse can be computed, for example, with the Sherman–Morrison formula.

A. Spectrum

By the properties of sigmoid functions,

ψ0(u0) = ψ0(0) +O(ε2). ψ0(v0) = ψ0(0) +O(ε2). Then ˆ Q1,1(ξ ) = 2ψ0(0)(R + ε) b2 b2+ ξ2+O(ε 2), ˆ Q1,2(ξ ) = ˆQ2,1(ξ ) = 2ψ0(0)R b2 b2+ ξ2+O(ε 2), ˆ Q2,2(ξ ) = 2ψ0(0)R ν2b2 ν2b2+ ξ2+O(ε 2).

Recall that the characteristic values are µ (ξ ) =1 2 Qˆ1,1(ξ ) − ˆQ2,2(ξ ) ± ±1 2 q ˆ Q1,1(ξ ) + ˆQ2,2(ξ ) 2 − 4 ˆQ1,2(ξ ) ˆQ2,1(ξ ).

Consider the expression under the square root ˆ Q1,1(ξ ) + ˆQ2,2(ξ ) 2 − 4 ˆQ1,2(ξ ) ˆQ2,1(ξ ) = = 2ψ 0 (0)Rb2 (b2+ ξ2) 2" 1 + ν2 b 2+ ξ2 ν2b2+ ξ2 2 + + 2  1 + ν2 b 2+ ξ2 ν2b2+ ξ2  ε − 4R2  +O(ε2). It has the same sign as

 1 + ν2 b 2+ ξ2 ν2b2+ ξ2  −2 −ε R+O(ε 2)= = ε R− (1 − ν 2) ξ2 ν2b2+ ξ2+O(ε 2). Since ξ2

ν2b22 is monotonically increasing as a function of ξ2,

there is a value ξ02= ν 2b2 1 − ν2 ε R+O(ε 2),

if 0 < ε/R < 1 − ν2, such that the expression under the square root is positive for 0 ≤ ξ2< ξ02and negative for ξ2> ξ02. It follows that

µ (ξ ) ∈ (

R, 0 ≤ ξ2< ξ02,

C \ R, ξ2> ξ02.

If ε < 0, the whole spectrum lies in C \ R and we can apply Theorem 1 to show that

Reµ(ξ ) < Reµ(ξ∗), ξ 6= ±ξ∗

when ε +O(ε2) > R(ν2− 1).

Suppose ε ≥ 0. Note that even though the spectrum is not completely complex, we can still apply Theorem 1. Indeed, the frequency ξ∗equals

ξ∗2= νb2  1 − 1 + ν 2(1 − ν) ε R  +O(ε2)

and clearly ξ∗2> ξ02. Then Reµ(ξ ) < Reµ(ξ∗) for all ξ2> ξ02,

ξ 6= ±ξ∗, provided that

ν2< 1 +ε R+O(ε

2) < 2 − ν2< 1

ν2.

Under the same conditions, ˆQ1,1(ξ ) − ˆQ2,2(ξ ) is growing on

(0, ξ0) and

ˆ

Q1,1(ξ ) − ˆQ2,2(ξ ) ≤ 4ψ0(0)ε +O(ε2), 0 ≤ ξ < ξ0.

Using monotonicity properties we can also derive that ˆ Q1,1(ξ ) + ˆQ2,2(ξ ) 2 − 4 ˆQ1,2(ξ ) ˆQ2,1(ξ ) < < Qˆ1,1(0) + ˆQ2,2(0) 2 − 4 ˆQ1,2(ξ0) ˆQ2,1(ξ0) = = (4ψ0(0)R)21 + ν 2 1 − ν2 ε R+O(ε 2)

on (0, ξ0). We thus get an upper bound

µ (ξ ) < 2ψ0(0)R s 1 + ν2 1 − ν2 ε R+O(ε), 0 ≤ ξ < ξ0, which shows that µ(ξ ) = Reµ(ξ ) < Reµ(ξ∗) for ξ2< ξ02. As

a result,

Reµ(ξ ) < Reµ(ξ∗), ξ 6= ±ξ∗.

B. Hopf bifurcation

For the Hopf bifurcation to occur, we must have µ (ξ∗) = σ∗+ iω∗, ω∗6= 0.

Straightforward calculations show that Reµ(ξ∗) = ψ0(0)R 1 + ν  1 − ν +ε R  +O(ε2), Imµ(ξ∗) = ψ0(0)R 1 + ν p (1 − ν)(3 + ν) −ε R  +O(ε2),

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

(8)

whence we get the bifurcation condition σ∗= ψ0(0)R 1 + ν  1 − ν +ε R  +O(ε2).

In Figure 1, we show Hopf bifurcation curves in the (R, ν) parameter space for different values of σ∗/ψ0(0).

0 2 4 6 8 10 12 14 R 0.0 0.2 0.4 0.6 0.8 1.0 */ 0(0) = 0.5 */ 0(0) = 1 */ 0(0) = 2 */ 0(0) = 5

FIG. 1: Curves in the (R, ν) parameter space where Hopf bifurcation occurs for different values of σ∗/ψ0(0) and ε = 0.

C. Normal form

Let σ be a small perturbation of the critical parameter value σ∗. Then the real part of the velocity of the critical

character-istic value is negative Reγ =  d dσReλ (ξ∗(0); σ )  σ =0 = −1 +O(ε2) < 0

and the condition σ Reγ > 0 transforms into σ < 0.

Consider matrices Mk(0). It follows from our derivations that M1(0) = " ψ0(0) +O(ε2) 0 0 −ψ0(0) +O(ε2) # , M2(0) = 1 2 " ψ000(0)u0+O(ε2) 0 0 −ψ000(0)v0+O(ε2) # , M3(0) = 1 6 " ψ000(0) +O(ε2) 0 0 −ψ000(0) +O(ε2) # .

In particular, kM2(0)k =O(ε). Note also that

P1= " 2R 1+ν 1 + 2−ν 2−2ν ε R  2R 1+ν 1 + ν 2−2ν ε R  2R 1+ν 1 + ν 2−2ν ε R  2Rν 1+ν 1 + 1 2−2ν ε R  # +O(ε2).

Then the eigenvector ζ1equals

ζ1= ψ0(0)R 1 + ν " 2 + ν 1−ν ε R 1 + ν − ip(1 − ν)(3 + ν) + ( 1 1−ν− i) ε R # +O(ε2).

For simplicity, we will only write the O(1) terms in what follows. In that case, we have

ζ1∗= 1 + ν ψ0(0)R   1 4+ i 4 1+ν √ (1−ν)(3+ν) −i 2 1 √ (1−ν)(3+ν)  +O(ε).

As kζ1k =O(1) and kM2(0)k =O(ε), the vectors hj, j =

1, . . . , 5 from Appendix A are all small, khjk =O(ε). Based

on the property of the Hadamard product, kx yk ≤ kxkkyk, we can conclude that

kM2(0)(ζ1 h1)k =O(ε2), kM2(0)(ζ1 h2)k =O(ε2),

kM2(0)(ζ1 h3)k =O(ε2), kM2(0)(ζ1 h4)k =O(ε2),

kM2(0)(ζ1 h5)k =O(ε2).

As a consequence, the normal form coefficients take a much simpler form: c1= 1 2l∗ D M3(0)(ζ1 ζ1 ζ1),P1Tζ ∗ 1 E +O(ε2), c2= 1 l∗ D M3(0)(ζ1 ζ1 ζ1),P1Tζ1∗ E +O(ε2).

Recall that for periodic traveling waves to be asymptoti-cally stable, three conditions need to be met:

Rec1< 0, Re(c1+ c2) < 0, Re(c1− c2) > 0.

In our case we have

Re(c1+ c2) = 3Rec1+O(ε2),

Re(c1− c2) = −Rec1+O(ε2).

This means that the stability of periodic traveling waves de-pends on a single condition

Rec1< 0.

Direct calculations give

c1= 1 3l∗  ψ0(0)R 1 + ν 3 ψ000(0) ψ0(0)  1 − ν + ip(1 − ν)(3 + ν)+O(ε). Therefore, since ν < 1, ψ0(0) > 0, and ψ000(0) < 0, the real part is negative, Rec1= 1 3l∗  ψ0(0)R 1 + ν 3 ψ000(0) ψ0(0)(1 − ν) +O(ε) < 0, and periodic traveling waves are asymptotically stable for all parameters of this class as long as |ε| is not too large.

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

(9)

TABLE II: Example of parameters.

a11 a12 a21 a22 b11 b12 b21 b22 σ∗

3.05 3.00 3.00 0.30 1.00 1.00 1.00 0.10 1.00

V. NUMERICAL EXAMPLES

For the numerical examples, we choose ψ1= ψ2= ψ with

ψ (u) = 2

πarctan(0.6782u) + 1.

Consider the parameters from Table II. The constant solution is u0= 0.404, v0= 0.287. Upon linearization about (u0, v0)

we can find that the critical spatial frequency is ξ∗= 0.318

and that λ (ξ∗) = ±1.86i (see Figure 2). The computation of

1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 Re ( ) 2.0 1.5 1.0 0.5 0.0 0.5 1.0 1.5 2.0 Im ( ) (± *)

FIG. 2: Distribution of characteristic values λ (ξ ) on the complex plain for the parameters from Table II. the normal form gives

c1≈ −0.0182 − 0.0386i, c2≈ −0.0365 − 0.0773i,

which means that the periodic traveling waves are asymptoti-cally stable.

For numerical verification, we ran numerical simulations on the interval (0, 2π/ξ∗) with periodic boundary conditions. We

computed the convolutions using the fast Fourier transform and the convolution theorem.

In Figure 3, we present numerical simulations for box-like initial states:

u0(x) = 2χ{x < π/ξ∗} − 1, v0(x) = 2χ{x < π/ξ∗} − 1.

Here, χ{Ω} is the indicator function of the set Ω such that χ {Ω}(x) = 1 if x ∈ Ω and χ {Ω}(x) = 0 if x 6∈ Ω. In the for-mulas above, both u0and v0are equal to 1 on (0, π/ξ∗) and

to -1 on (π/ξ∗, 2π/ξ∗). As the simulations show, the solution

retains the symmetry of the initial state and becomes a stand-ing wave. However, if we shrink or enlarge the positive half to (0, π/ξ∗± ε) for one of the components (let it be v0), the

solution turns into a periodic traveling wave since the initial state breaks the reflection symmetry of the system.

It is also important to understand how the system evolves from zero initial state in the presence of a forcing term. Namely, we are interested in time-oscillating point sources like

f(x,t) = δ (x − x0) sin(t),

where δ (x) is the δ -function. So we want to solve the follow-ing initial value problem:

∂ u ∂ t = P1,1∗ ψ1(u) − P1,2∗ ψ2(v) − σ u + f1(x,t), ∂ v ∂ t = P2,1∗ ψ1(u) − P2,2∗ ψ2(v) − σ v + f2(x,t), u(x, 0) = 0, v(x, 0) = 0. In Figure 4, we use fj(x,t) = 0.1δ (x − xj) sin(ωt), j= 1, 2,

as the forcing terms. If x1= x2, that is the point sources

are placed at the exactly same positions, the zero initial state evolves into a standing wave. Once again, this is because the reflection symmetry is not broken. Meanwhile, should the point sources be placed at different (possibly very close) posi-tions, x16= x2, the solution is a periodic traveling wave, which

propagates in the direction determined by the relative position of the two sources.

In Figure 5, we explore the case where the forcing terms have different oscillation frequencies:

f1(x,t) = 0.1δ (x − π/2ξ∗) sin(At),

f2(x,t) = 0.1δ (x) sin(t).

We observed that for a fixed distance between the forcing terms, the propagation direction of the traveling waves de-pends on the ratio of the forcing term frequencies A > 0. Namely, there is a threshold value A∗such that waves

propa-gate in different directions for A < A∗and A > A∗. Moreover,

as A → A∗, the solution spends more time in the form of a

standing wave, which suggests that at A = A∗the solution is a

standing wave.

VI. DISCUSSION

With the development of modern technologies it has been possible to demonstrate the existence of periodic traveling waves in cortex during many human activities7. These propa-gating waves help explain brain dynamics observed at the pop-ulation level30. For example, Huang et al.31,32studied brain wave dynamics for years using voltage-sensitive dye imaging and demonstrated the presence of spiral dynamics in the intact mammalian brain during pharmacologically induced oscilla-tions as well as during sleep-like states. These authors showed that traveling waves are emergent properties which can influ-ence not only brain oscillation frequencies and spatial coher-ence but they can also modulate their amplitude. As a result

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

(10)

0 5 10 15 x U 0 20 40 60 80 100 120 140 time 0 5 10 15 x V 1.0 0.5 0.0 0.5 1.0 1.5 (a) v0(x) = 2χ{x < π/ξ∗} − 1 0 5 10 15 x U 0 20 40 60 80 100 120 140 time 0 5 10 15 x V 1.0 0.5 0.0 0.5 1.0 1.5 (b) v0(x) = 2χ{x < π/ξ∗− ε} − 1 0 5 10 15 x U 0 20 40 60 80 100 120 140 time 0 5 10 15 x V 1.0 0.5 0.0 0.5 1.0 1.5 (c) v0(x) = 2χ{x < π/ξ∗+ ε} − 1

FIG. 3: Numerical simulation with the initial state u0(x) = 2χ{x < π/ξ∗} − 1 and σ = 0.95. Level lines of the solution u(x,t)

(upper row) and v(x,t) (lower row) are shown for three different initial states v0(x). Standing waves are observed for the

symmetric initial state (a), periodic traveling waves for the asymmetric initial states (b,c).

0 5 10 15 x U 0 20 40 60 80 100 120 140 time 0 5 10 15 x V 1.0 0.5 0.0 0.5 1.0 1.5 (a) x1= 0, x2= 0 0 5 10 15 x U 0 20 40 60 80 100 120 140 time 0 5 10 15 x V 1.0 0.5 0.0 0.5 1.0 1.5 (b) x1= ε, x2= 0 0 5 10 15 x U 0 20 40 60 80 100 120 140 time 0 5 10 15 x V 1.0 0.5 0.0 0.5 1.0 1.5 (c) x1= 0, x2= ε

FIG. 4: Numerical simulation with the forcing terms fj(x,t) = 0.1δ (x − xj) sin(ωt), j = 1, 2, and σ = 0.92. Level lines of the

solution u(x,t) (upper row) and v(x,t) (lower row) are shown for three different forcing terms. Standing waves are observed for the symmetric forcing term (a), periodic traveling waves for the asymmetric forcing terms (b,c).

these waves «may contribute to both normal cortical process-ing and to pathological patterns of activity such as those found in epilepsy» (Ref. 30, P. 1591). Interestingly Stead et al.33

observed «microseizures» (spiral waves) on isolated research microelectrodes in both epileptic patients (in regions prone to generate epileptic seizures) but also in control subjects. They conclude that only «the density of these phenomena distin-guishes normal from epileptic brain» (Ref. 33, P. 2789).

Although the functional role of traveling waves is not com-pletely elucidated yet, new techniques based on targeted neu-romodulation of traveling waves using electric fields are al-ready being explored as potential therapeutic strategies. For example, Whalen et al.2 were able to control the modula-tion, suppression and prevention or to increase propagation velocity of spreading depression with DC electrical fields in brain slices thus demonstrating «the potential feasibility of electrical control and prevention of spreading depression» (Ref. 2, P. 8769). Using coupled oscillators Ermentrout and Kleinfeld34 also studied electrical waves in the cortex and

speculated as to their computational role.

Neural field models provide an appropriate tool to study these waves but the results differ for different models. A sin-gle equation with symmetric connectivity functions for acti-vating and inhibitory signals does not have solutions in the

form of periodic traveling waves. In the case of asymmetric functions, such solutions exist, they are stable, and solutions with different frequencies (speeds) are observed for the same values of parameters35. The last property is important for the understanding of the existence of the waves with different fre-quencies during brain activity.

A single equation with symmetric connectivity functions can also manifest periodic waves in the presence of time delay36. The periodic waves bifurcate from the stationary,

spatially homogeneous solution being unstable, and they come stable if time delay exceeds some critical value. As be-fore, the coexistence of waves with different frequencies is ob-served. Another important characterization of periodic waves related to the choice of initial conditions should also be taken into account. We will return to this question below.

The models with two different neuron populations, excita-tory and inhibiexcita-tory, can also describe the propagation of pe-riodic waves17,18,21,22,28. In this work we considered a par-ticular case of the model Ref. 28 without time delay and with symmetric connectivity functions. Nonlinear bifurcation anal-ysis shows two types of solutions: standing oscillating solu-tions with the behavior qualitatively similar to the product of two periodic functions, cos(ωt) cos(kx), and periodic travel-ing waves similar to cos(ωt + kx). Their stability depends on

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

(11)

0 5 10 15 x U 0 50 100 150 200 250 300 time 0 5 10 15 x V 1.0 0.5 0.0 0.5 1.0 (a) A = 1 0 5 10 15 x U 0 50 100 150 200 250 300 time 0 5 10 15 x V 1.0 0.5 0.0 0.5 1.0 1.5 (b) A = 1.973469 0 5 10 15 x U 0 50 100 150 200 250 300 time 0 5 10 15 x V 1.0 0.5 0.0 0.5 1.0 1.5 (c) A = 1.973470

FIG. 5: Numerical simulation with the forcing terms f1(x,t) = 0.1δ (x − π/2ξ∗) sin(At), f2(x,t) = 0.1δ (x) sin(t), and σ = 0.92.

Level lines of the solution u(x,t) (upper row) and v(x,t) (lower row) are shown for three different forcing terms. The traveling waves move to the right for A below the threshold value (a,b) and to the left for A above the threshold value (c).

the values of parameters, and they are mutually exclusive.

From the biological point of view we are interested not only in the existence and stability of periodic traveling waves but also in their initiation, that is, under what conditions (initial, boundary, etc), the solution will approach the wave. This question can be potentially important to determine the role of epicenters (hubs) in the functioning of brain connectome. The epicenters receive the information from other epicenters or from the external body organs, they process this information and send it further downstream. Assuming that the synchro-nization of epicenters is effectuated by cortical waves, we can suppose that the waves are initiated locally by the activated epicenter. In the case of a single equation, such initiation appears to be possible for asymmetric connectivity function but not for symmetric connectivity functions and time delay36. The latter requires the initial conditions close to the wave on the whole space interval. The model considered in this work admits wave initiation by local forcing terms (additional terms in the equations). If the forcing term is symmetric, then stand-ing oscillatstand-ing solutions are observed. In the case of asymmet-ric forcing term, under the parameters for which the periodic wave is stable, solution converges to the wave.

Let us note that the direction of propagation is one of the important parameters characterizing cortical waves. We show that this direction is determined by the forcing term. In the 2D setting we expect that we can obtain any direction be-tween 0 and 2π. Hence, the epicenters can exchange directed signals12. Wave initiation by local forcing terms is observed in the case of brain stimulation with electrodes37. External brain stimulation can be used in order to restore wave propagation in a partially damaged tissue35,36. Recent perspectives propose to control brain traveling waves in human subjects by stim-ulating preidentified nodes using electrical stimulation. One of these initiatives propose to modulate the activity of cortical epicenters of the language network to treat chronic conditions such as post stroke aphasia38,39.

ACKNOWLEDGMENTS

Funding was provided by "RUDN University Program 5-100” (SB, implementation of the model; VV, designed the model and supervised analyses and interpretations), by CorStim (AB, designed and conceptualized the research), by the Moscow Center for Fundamental and Applied Mathemat-ics (SB, calculations; agreement with the Ministry of Educa-tion and Science of the Russian FederaEduca-tion No. 075-15-2019-1624). The manuscript was written and revised with input from all three authors.

DATA AVAILABILITY

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

Appendix A: Computation of the normal form coefficients

To compute the cubic coefficients c1∈ C and c2∈ C, we

introduce a function ˜

F(w, σ ) = F(w, σ ) + L(σ )w − L(0)w − σ L0(0)w

and use the formulas given in Ref. 29. The computation in-volves second and third-order Fréchet derivatives of ˜F(w, σ ) ( ˜Fww(0, 0) and ˜Fwww(0, 0)) at w = 0 and σ = 0: ˜ F2(w1, w2) =

n∈Z PnM2(0)(w1 w2)nen, ˜ F3(w1, w2, w3) =

n∈Z PnM3(0)(w1 w2 w3)nen.

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

(12)

It remains to evaluate the following expressions: g1= 1 2(2iω∗− L(0)) −1F˜ 2(q1, q1), g2= −L(0)−1F˜2(q1, q1), g3= (2iω∗− L(0))−1F˜2(q1, q3), g4= −L(0)−1F˜2(q1, q3), g5= −L(0)−1F˜2(q3, q3).

We treat them one by one:

1. ˜ F2(q1, q1) = 1 √ l∗ P2M2(0)(ζ1 ζ1)e2, g1= h1e2, h1∈ C2, (2iω∗−P2M1(0))h1= 1 2√l∗P2 M2(0)(ζ1 ζ1); 2. ˜ F2(q1, q1) = 1 √ l∗ P0M2(0)(ζ1 ζ1)e0, g2= h2e0, h2∈ C2, (P0M1(0))h2= − 1 √ l∗P0 M2(0)(ζ1 ζ1); 3. ˜ F2(q1, q3) = 1 √ l∗P2 M2(0)(ζ1 ζ1)e2, g3= h3e2, h3∈ C2, (2iω∗−P2M1(0))h3= 1 √ l∗P2 M2(0)(ζ1 ζ1); 4. ˜ F2(q1, q3) = 1 √ l∗ P0M2(0)(ζ1 ζ1)e0, g4= h4e0, h4∈ C2, (P0M1(0))h4= − 1 √ l∗P0 M2(0)(ζ1 ζ1); 5. ˜ F2(q3, q3) = 1 √ l∗P0 M2(0)(ζ1 ζ1)e0, g5= h5e0, h5∈ C2, (P0M1(0))h5= − 1 √ l∗P0 M2(0)(ζ1 ζ1).

We also need two terms with third-order derivatives: ˜ F3(q1, q1, q1) = 1 l∗P1 M3(0)(ζ1 ζ1 ζ1)e1, ˜ F3(q1, q3, q3) = 1 l∗P1 M3(0)(ζ1 ζ1 ζ1)e1.

The normal form coefficients are then given by the follow-ing formulas29: c1= 1 2h ˜F3(q1, q1, q1) + 2 ˜F2(q1, g1) + 2 ˜F2(q1, g2), q ∗ 1iX, c2= h ˜F3(q1, q3, q3) + ˜F2(q3, g3) + ˜F2(q3, g4) + ˜F2(q1, g5), q∗1iX. It follows that c1=  1 2l∗P1 M3(0)(ζ1 ζ1 ζ1) + 1 √ l∗P1 M2(0)(ζ1 h1) + 1 √ l∗P1 M2(0)(ζ1 h2), ζ1∗  =√1 l∗  1 2√l∗ M3(0)(ζ1 ζ1 ζ1) + M2(0)(ζ1 h1) + M2(0)(ζ1 h2),P1Tζ ∗ 1  , c2=  1 l∗P1 M3(0)(ζ1 ζ1 ζ1) + 1 √ l∗ P1M2(0)(ζ1 h3) + 1 √ l∗ P1M2(0)(ζ1 h4) + 1 √ l∗ P1M2(0)(ζ1 h5), ζ1∗  =√1 l∗  1 √ l∗ M3(0)(ζ1 ζ1 ζ1) + M2(0)(ζ1 h3) + M2(0)(ζ1 h4) + M2(0)(ζ1 h5),P1Tζ ∗ 1  .

1M. Massimini, R. Huber, F. Ferrarelli, S. Hill, and G. Tononi, “The Sleep Slow Oscillation as a Traveling Wave,” Journal of Neuroscience 24, 6862– 6870 (2004).

2A. J. Whalen, Y. Xiao, H. Kadji, M. A. Dahlem, B. J. Gluckman, and S. J. Schiff, “Control of Spreading Depression with Electrical Fields,” Scientific Reports 8, 1–9 (2018).

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

(13)

3N. Kopell and L. N. Howard, “Plane Wave Solutions to Reaction-Diffusion Equations,” Studies in Applied Mathematics 52, 291–328 (1973). 4N. Kopell and L. N. Howard, “Horizontal Bands in the Belousov Reaction,”

Science 180, 1171–1173 (1973).

5A. Bahramisharif, M. A. J. van Gerven, E. J. Aarnoutse, M. R. Mercier, T. H. Schwartz, J. J. Foxe, N. F. Ramsey, and O. Jensen, “Propagating Neocortical Gamma Bursts Are Coordinated by Traveling Alpha Waves,” Journal of Neuroscience 33, 18849–18854 (2013).

6M. Halgren, I. Ulbert, H. Bastuji, D. Fabó, L. Er˝oss, M. Rey, O. Devinsky, W. K. Doyle, R. Mak-McCully, E. Halgren, L. Wittner, P. Chauvel, G. Heit, E. Eskandar, A. Mandell, and S. S. Cash, “The generation and propagation of the human alpha rhythm,” Proceedings of the National Academy of Sci-ences 116, 23772–23782 (2019).

7D. M. Alexander, A. R. Nikolaev, P. Jurica, M. Zvyagintsev, K. Mathiak, and C. van Leeuwen, “Global Neuromagnetic Cortical Fields Have Non-Zero Velocity,” PLoS ONE 11 (2016), 10.1371/journal.pone.0148413. 8J.-Y. Wu, X. Huang, and C. Zhang, “Propagating Waves of Activity in the

Neocortex: What They Are, What They Do,” The Neuroscientist 14, 487– 502 (2008).

9L. Muller, F. Chavane, J. Reynolds, and T. J. Sejnowski, “Cortical trav-elling waves: Mechanisms and computational principles,” Nature Reviews Neuroscience 19, 255–268 (2018).

10H. Zhang, A. J. Watrous, A. Patel, and J. Jacobs, “Theta and Alpha Oscil-lations Are Traveling Waves in the Human Neocortex,” Neuron 98, 1269– 1281.e4 (2018).

11J. Rapela, “Traveling waves appear and disappear in unison with produced speech,” arXiv:1806.09559 [q-bio] (2018), arXiv:1806.09559 [q-bio]. 12K. Takahashi, M. Saleh, R. D. Penn, and N. Hatsopoulos,

“Propagat-ing Waves in Human Motor Cortex,” Frontiers in Human Neuroscience 5 (2011), 10.3389/fnhum.2011.00040.

13D. Lozano-Soldevilla and R. VanRullen, “The Hidden Spatial Dimension of Alpha: 10-Hz Perceptual Echoes Propagate as Periodic Traveling Waves in the Human Brain,” Cell Reports 26, 374–380.e4 (2019).

14V. Botella-Soler, M. Valderrama, B. Crépon, V. Navarro, and M. Le Van Quyen, “Large-Scale Cortical Dynamics of Sleep Slow Waves,” PLoS ONE 7 (2012), 10.1371/journal.pone.0030757.

15J. Weaver, “How brain waves help us make sense of speech,” PLoS Biology 11 (2013).

16D. Pinotsis, P. Robinson, P. beim Graben, and K. Friston, “Neural masses and fields: Modeling the dynamics of brain activity,” Frontiers in Compu-tational Neuroscience 8 (2014), 10.3389/fncom.2014.00149.

17H. R. Wilson and J. D. Cowan, “A mathematical theory of the functional dynamics of cortical and thalamic nervous tissue,” Kybernetik 13, 55–80 (1973).

18S. Atasoy, I. Donnelly, and J. Pearson, “Human brain networks function in connectome-specific harmonic waves,” Nature Communications 7, 1–10 (2016).

19S. Heitmann, P. Gong, and M. Breakspear, “A computational role for bista-bility and traveling waves in motor cortex,” Frontiers in Computational Neuroscience 6 (2012), 10.3389/fncom.2012.00067.

20N. A. Venkov, S. Coombes, and P. C. Matthews, “Dynamic instabilities in scalar neural field equations with space-dependent delays,” Physica D: Nonlinear Phenomena 232, 1–15 (2007).

21F. M. Atay and A. Hutt, “Neural Fields with Distributed Transmission Speeds and Long-Range Feedback Delays,” SIAM Journal on Applied Dy-namical Systems 5, 670–698 (2006).

22G. B. Ermentrout, S. E. Folias, and Z. P. Kilpatrick, “Spatiotemporal Pat-tern Formation in Neural Fields with Linear Adaptation,” in Neural Fields: Theory and Applications, edited by S. Coombes, P. beim Graben, R. Pot-thast, and J. Wright (Springer, Berlin, Heidelberg, 2014) pp. 119–151. 23H. G. E. Meijer and S. Coombes, “Travelling waves in a neural field

model with refractoriness,” Journal of Mathematical Biology 68, 1249– 1268 (2014).

24G. B. Ermentrout and J. B. McLeod, “Existence and uniqueness of trav-elling waves for a neural network,” Proceedings of the Royal Society of Edinburgh Section A: Mathematics 123, 461–478 (1993/ed).

25Z. Chen, B. Ermentrout, and X.-J. Wang, “Wave Propagation Mediated by GABAB Synapse and Rebound Excitation in an Inhibitory Network: A Reduced Model Approach,” Journal of Computational Neuroscience 5, 53– 69 (1998).

26D. J. Pinto and G. B. Ermentrout, “Spatially Structured Activity in Synapti-cally Coupled Neuronal Networks: I. Traveling Fronts and Pulses,” SIAM Journal on Applied Mathematics 62, 206–225 (2001).

27D. J. Pinto and G. B. Ermentrout, “Spatially Structured Activity in Synap-tically Coupled Neuronal Networks: II. Lateral Inhibition and Standing Pulses,” SIAM Journal on Applied Mathematics 62, 226–243 (2001). 28J. Senk, K. Korvasová, J. Schuecker, E. Hagen, T. Tetzlaff, M. Diesmann,

and M. Helias, “Conditions for wave trains in spiking neural networks,” Physical Review Research 2, 023174 (2020).

29S. A. van Gils and J. Mallet-Paret, “Hopf bifurcation and symmetry: Trav-elling and standing waves on the circle,” Proceedings of the Royal Society of Edinburgh: Section A Mathematics 104, 279–307 (1986).

30A. Keane and P. Gong, “Propagating Waves Can Explain Irregular Neural Dynamics,” Journal of Neuroscience 35, 1591–1605 (2015).

31X. Huang, W. C. Troy, Q. Yang, H. Ma, C. R. Laing, S. J. Schiff, and J.-Y. Wu, “Spiral Waves in Disinhibited Mammalian Neocortex,” Journal of Neuroscience 24, 9897–9902 (2004).

32X. Huang, W. Xu, J. Liang, K. Takagaki, X. Gao, and J.-y. Wu, “Spiral Wave Dynamics in Neocortex,” Neuron 68, 978–990 (2010).

33M. Stead, M. Bower, B. H. Brinkmann, K. Lee, W. R. Marsh, F. B. Meyer, B. Litt, J. Van Gompel, and G. A. Worrell, “Microseizures and the spatiotemporal scales of human partial epilepsy,” Brain 133, 2789–2797 (2010).

34G. B. Ermentrout and D. Kleinfeld, “Traveling Electrical Waves in Cortex: Insights from Phase Dynamics and Speculation on a Computational Role,” Neuron 29, 33–44 (2001).

35N. Bessonov, A. Beuter, S. Trofimchuk, and V. Volpert, “Estimate of the travelling wave speed for an integro-differential equation,” Applied Mathe-matics Letters 88, 103–110 (2019).

36N. Bessonov, A. Beuter, S. Trofimchuk, and V. Volpert, “Cortical waves and post-stroke brain stimulation,” Mathematical Methods in the Applied Sciences 42, 3912–3928 (2019).

37A. Opitz, A. Falchier, C.-G. Yan, E. M. Yeagle, G. S. Linn, P. Megevand, A. Thielscher, R. D. A, M. P. Milham, A. D. Mehta, and C. E. Schroeder, “Spatiotemporal structure of intracranial electric fields induced by transcra-nial electric stimulation in humans and nonhuman primates,” Scientific Re-ports 6, 1–11 (2016).

38A. Beuter, A. Balossier, S. Trofimchuk, and V. Volpert, “Modeling of post-stroke stimulation of cortical tissue,” Mathematical Biosciences 305, 146– 159 (2018).

39A. Beuter, A. Balossier, F. Vassal, S. Hemm, and V. Volpert, “Cortical stimulation in aphasia following ischemic stroke: Toward model-guided electrical neuromodulation,” Biological Cybernetics 114, 5–21 (2020).

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

(14)

0

2

4

6

8

10

12

14

R

0.0

0.2

0.4

0.6

0.8

1.0

*

/

0

(0) = 0.5

*

/

0

(0) = 1

*

/

0

(0) = 2

*

/

0

(0) = 5

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

(15)

1.4

1.2

1.0

0.8

0.6

0.4

0.2

0.0

Re ( )

2.0

1.5

1.0

0.5

0.0

0.5

1.0

1.5

2.0

Im

(

)

*

)

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

(16)

0

5

10

15

x

U

0

20

40

60

80

100

120

140

time

0

5

10

15

x

V

1.0

0.5

0.0

0.5

1.0

1.5

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

(17)

0

5

10

15

x

U

0

20

40

60

80

100

120

140

time

0

5

10

15

x

V

1.0

0.5

0.0

0.5

1.0

1.5

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

(18)

0

5

10

15

x

U

0

20

40

60

80

100

120

140

time

0

5

10

15

x

V

1.0

0.5

0.0

0.5

1.0

1.5

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

(19)

0

5

10

15

x

U

0

20

40

60

80

100

120

140

time

0

5

10

15

x

V

1.0

0.5

0.0

0.5

1.0

1.5

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

(20)

0

5

10

15

x

U

0

20

40

60

80

100

120

140

time

0

5

10

15

x

V

1.0

0.5

0.0

0.5

1.0

1.5

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

(21)

0

5

10

15

x

U

0

20

40

60

80

100

120

140

time

0

5

10

15

x

V

1.0

0.5

0.0

0.5

1.0

1.5

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

(22)

0

5

10

15

x

U

0

50

100

150

200

250

300

time

0

5

10

15

x

V

1.0

0.5

0.0

0.5

1.0

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

(23)

0

5

10

15

x

U

0

50

100

150

200

250

300

time

0

5

10

15

x

V

1.0

0.5

0.0

0.5

1.0

1.5

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

(24)

0

5

10

15

x

U

0

50

100

150

200

250

300

time

0

5

10

15

x

V

1.0

0.5

0.0

0.5

1.0

1.5

This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.

Références

Documents relatifs

Nonlinear Schrödinger equations, standing waves, orbital stability, instability, blow-up, variational methods..

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

In this paper, we have studied the existence of traveling wave solutions and spreading properties for single-layer delayed neural field equations when the kinetic dynamics are

The synchronous recycling has been shown to be effective with multipass delay lines as well as with Fabry-Perot cavities, the resonant amplitudes of output modulated

On one hand, we put the one- dimensional theory, or in other words the stability theory for longitudinal perturbations, on a par with the one available for systems of Korteweg

Gérard, Lenzmann, Pocovnicu and Raphaël [17] deduced the invertibility of the linearized operator for the half-wave equation around the profiles Q β when β is close enough to 1,

As for the nonlinear stability, the only result we are aware of is due to Angulo [1], who proved very recently that the family of dnoidal waves of the focusing NLS equation is

In contrast to the KP-I equation, now we show that the spectrum of A a (ℓ) is purely imaginary when ℓ and a sufficiently small, i.e., the small periodic waves of the KP-II equation