Results

-axis were initiated in a two-dimensional medium of $$6.25\,\text {cm} \times 6.25\,\text {cm}$$ size, using the smooth Fenton-Karma model with parameter set FK2 (see Table A.2, p. 184), isotropic diffusion of $$D=1\,{\text {cm}^2/\text {s}}$$ and periodic boundary conditions. Due to the periodic boundary conditions, the plane wave does not die out at the medium boundary but continuously reenters the domain from the opposite side. For the parameters used here, a stable periodic activation pattern develops, such that $$\lambda _i\le 0$$ is expected for all $$i$$, meaning that no perturbation should grow with time. In Fig. 3.1, a snapshot of the system is plotted along with the perturbations belonging to the leading Lyapunov exponents. As indicated in the figure, there is one neutrally stable direction with $$\lambda _1=0$$ and all other $$\lambda _i<0$$, $$i\ge 2$$. This is entirely consistent with the Lyapunov spectrum of a periodic orbit in an autonomous low-dimensional system, where the neutrally stable direction corresponds to a shift along the trajectory, indicating time invariance of the dynamical system. However, in this case, the Lyapunov vectors have a spatial structure: They are concentrated at the wave front and back and show a characteristic vertical oscillation in space. The first vector (second column) indeed corresponds to a shift along the trajectory, accelerating the plane wave towards a future state. A cut along the $$y$$-axis of the first 15 Lyapunov vectors, shown in Fig. 3.2, indicates that the decay rate (cf. figure legend) of the perturbations increases with the spatial frequency. The harmonic spatial shape of the perturbations suggests to attribute their existence to the diffusive spatial coupling of the excitable medium. The emergence of the observed spectrum will be explained in the following.


A328193_1_En_3_Fig1_HTML.gif


Fig. 3.1
Example perturbations of a plane wave. The plane wave is traveling from left to right in a square medium with periodic boundary conditions. In the leftmost column, a snapshot of the medium state at an arbitrary point in time is depicted, the five columns on the right show snapshots of five associated Lyapunov vectors (at the same time), ordered by growth rate. For the first Lyapunov vector $$\lambda _1\approx 0$$, since the plane wave is a periodic orbit in an autonomous system. The higher-index perturbations have negative growth rates, indicated below each column (see Fig. 3.2 for details). The red line in the upper left plot indicates the domain of a one-dimensional cable which, if considered separately, would exhibit identical dynamics. Simulations done in a $$6.25\,\text {cm}\times 6.25\,\text {cm}$$ domain with the smooth Fenton-Karma model using parameter set FK2 (see Table A.2, p. 184), isotropic diffusion of $$D=1\,{\text {cm}^2/\text {s}}$$ and periodic boundary conditions. The space step is $$h=521\,{\upmu \text {m}}$$, the time step $$\Delta t = 12.5\,{\upmu \text {s}}$$. Lyapunov vectors were allowed to align for $$2{,}800\,\text {ms}$$ before calculating the spectrum according to Sect. 2.​2.​6 for another $$2{,}800\,\text {ms}$$ (roughly 13 periods)


A328193_1_En_3_Fig2_HTML.gif


Fig. 3.2
Wave front modulation for a plane wave. The plot shows a cut along the wave front (parallel to the $$y$$-axis) through the $$\delta u$$-component of the first 15 Lyapunov vectors. The position $$x_0$$ of the cut is indicated by arrows in the first row of Fig. 3.1. Model and numerical parameters are identical to Fig. 3.1


Indeed, a plane wave in the two-dimensional system can be viewed as the natural extension of a one-dimensional scenario: Because of the absence of spatial variations in the $$y$$-direction, the two-dimensional version of Eq. (2.​72), which describes the actual dynamics, reduces to a one-dimensional problem:


$$\begin{aligned} \frac{\partial u}{\partial t}= D \Bigg (\frac{\partial ^{2} u}{\partial x^{2}} + \underbrace{\frac{\partial ^{2} u}{\partial y^{2}}}_{=0}\Bigg ) + f(u, \mathbf {h}) = D \frac{\partial ^{2} u}{\partial x^{2}} + f(u, \mathbf {h}) \end{aligned}$$

(3.1)
This is the equation for a one-dimensional cable (indicated by a red line in Fig. 3.1). The Lyapunov spectrum of $$N$$ uncoupled copies of such a cable would consist of $$N$$ copies of the spectrum of a single cable. However, the diffusion operator $$\frac{\partial ^{2} }{\partial y^{2}}$$ which couples the cables in $$y$$-direction stabilizes the trajectory where $$u(x,y,t)$$ is identical for all $$y$$ and thus shifts some Lyapunov exponents in the negative direction.

Following the above picture, the structure of the Lyapunov spectrum can be explained in the context of transverse Lyapunov exponents for a coupled system of identical oscillators, where one oscillator corresponds to the aforementioned periodic one-dimensional cable. Its $$n=3N_x$$ degrees of freedom are the values of the three dynamic variables in every of the $$N_x$$ computational points. $$N$$ of these oscillators (where $$N=N_y$$ is the number of computational points in $$y$$-direction in the two-dimensional simulation) exhibiting the periodic dynamics of a traveling wave are diffusively coupled in $$y$$-direction. Assume for a moment that all degrees of freedom of neighboring oscillators are coupled in the same way via diffusion, corresponding to the following system of equations for the Fenton-Karma model [(replacing Eq. (2.​40)]:


$$\begin{aligned} \frac{\partial u}{\partial t}&=\nabla \cdot \underline{ D}\nabla u-J_\text {fi}(u,v)-J_\text {so}(u)-J_\text {si}(u,w)\end{aligned}$$

(3.2a)



$$\begin{aligned} \frac{\partial v}{\partial t}&=H_1(u,v)+\nabla \cdot \underline{ D}\nabla v\end{aligned}$$

(3.2b)



$$\begin{aligned} \frac{\partial w}{\partial t}&=H_2(u,w)+\nabla \cdot \underline{ D}\nabla w \end{aligned}$$

(3.2c)


A328193_1_En_3_Fig3_HTML.gif


Fig. 3.3
Lyapunov spectrum for a traveling wave in a cable. The wave is traveling from left to right in a cable modeled by the modified system 3.2a with periodic boundary conditions. The first column displays the spatial distribution of the model variables $$u$$, $$v$$ and $$w$$. In the second column, the first four Lyapunov vectors are plotted. Corresponding Lyapunov exponents are indicated in the legend. Numerical parameters are identical to Fig. 3.1 except for a vanishing spatial extent of the domain in $$y$$-direction

$$H_1(u,v)$$ and $$H_2(u,v)$$ denote the original right hand side for $$v$$ and $$w$$ of Eq. (2.​40). A snapshot of the dynamics in one cable together with the leading Lyapunov spectrum is shown in Fig. 3.3. With this modified system of equations, the transverse Lyapunov exponents of the numerically discretized system viewed as a collection of coupled cables can be calculated analytically, following Heagy et al. [6]: They derived an expression for a system of coupled oscillators with an additive coupling term for the $$j$$th oscillator of the form $$c[\mathbf {z}_{j-1}-2\mathbf {z}_j+\mathbf {z}_{j+1}]$$, where $$\mathbf {z}_j$$ is the state vector of the $$j$$th oscillator with its $$n$$ degrees of freedom. This corresponds exactly to one possible discretization of a second derivative such as the $$y$$-component of the Laplacian. Equations (2.​86) and (2.​87) state that the coupling strength in this context is $$c=D\frac{1}{h^2}=D\frac{N_y^2}{L_y^2}$$. The result of Heagy et al. now is that the transverse Lyapunov exponents can be calculated as follows:


$$\begin{aligned} \lambda _{i}^{k} = \lambda _i^0-4c\sin ^2\left( \frac{\pi k}{N_y}\right) , \end{aligned}$$

(3.3)
$$1\le k\le N_y-1$$ is the index denoting the spatial mode, and $$\lambda _i^0$$ with $$1\le i\le n$$ are the Lyapunov exponents for a single cable without coupling, which were illustrated in Fig. 3.3, such that $$\lambda _i^0 = \lambda ^\text {1d}_i$$. In the continuum limit $$N_y\rightarrow \infty $$, the origin of this shift becomes immediately clear, because for $$k\ll N_y$$


$$\begin{aligned} \lambda _i^k \approx \lambda _i^0 -4c\left( \frac{\pi k}{N_y}\right) ^2 = \lambda _i^0 - D\left( \frac{2\pi k}{L_y}\right) ^2. \end{aligned}$$

(3.4)
The last term can be recognized as the eigenvalues of the diffusion operator which is responsible for the coupling and whose eigenspaces are two-dimensional:


$$\begin{aligned} D\frac{\partial ^{2} }{\partial y^{2}} \left( A\sin \left( \gamma _ky\right) + B\cos \left( \gamma _ky\right) \right) = -D\gamma _k^2 \left( A\sin \left( \gamma _ky\right) + B\cos \left( \gamma _ky\right) \right) , \end{aligned}$$

(3.5)
where $$\gamma _k = \frac{2\pi k}{L_y}$$. The fact that the shift of the transverse Lyapunov exponents is given by the eigenvalues of the $$y$$-component of the Laplace operator is, of course, not a coincidence: Let $$\delta \mathbf {U}^\text {1d}(x)=[\delta u^\text {1d}(x), \delta v^\text {1d}(x), \delta w^\text {1d}(x)]$$ be a Lyapunov vector of the one-dimensional cable associated with a Lyapunov exponent of $$\lambda ^\text {1d}$$ in the co-moving frame of reference (hence the time independence). Let $$\mathbf {L}^\text {1d}$$ denote the linearized dynamics of the cable in the co-moving frame (i.e. the linearization around the fixed point) and $$\mathbf {L} = \mathbf {L}^\text {1d} + \frac{\partial ^{2} }{\partial y^{2}}$$ that of the two-dimensional system with complete coupling as in Eq. (3.2a).1 Then, it is known that $$\delta \mathbf {U}^\text {1d}$$ is an eigenvector of $$\mathbf {L}^\text {1d}$$ with eigenvalue $$\lambda ^\text {1d}$$. It is seen immediately that with an eigenfunction $$f_k(y)$$ of the Laplace operator with eigenvalue $$\gamma _k$$, the definition


$$\begin{aligned} \delta \mathbf {U}(x,y) = \delta \mathbf {U}^\text {1d}(x)f_k(y) \end{aligned}$$

(3.6)
yields an eigenvector $$\delta \mathbf {U}(x,y)$$ of $$\mathbf {L}$$ with eigenvalue $$\lambda ^\text {1d} -D\gamma _k^2$$. Thus, each choice of $$k$$ yields a different copy of the spectrum of the one dimensional system in the two-dimensional medium, shifted by $$-D\gamma _k^2$$, which is exactly the result of Eq. (3.4).

Since $$\lambda _1^\text {1d}=0$$ for a single cable, a number of Lyapunov exponents equal to the second term of Eq. (3.3) are expected to appear in between $$\lambda _1^\text {1d}$$ and $$\lambda _2^\text {1d}$$ (if there is a large enough gap between them). To verify this conjecture, simulations of the full two-dimensional system with the modified Eq. (3.2a) were carried out. As shown in Table 3.1, the Lyapunov spectrum of the two-dimensional system corresponds exactly (within numerical errors) to that expected from Eq. (3.3) for a system of coupled one-dimensional cables: “Copies” of both $$\lambda ^\text {1d}_1$$ and $$\lambda ^\text {1d}_2$$ can be found in the spectrum with their values shifted by the second term of Eq. (3.3).


Table 3.1
Lyapunov spectrum of a plane wave for complete coupling
































































































































































$$\lambda _i^\text {1d}=\lambda _i^0$$

$$\lambda ^\text {2d}=\lambda _i^k$$

$$(\lambda _i^k-\lambda _i^0)/(-4c\sin ^2(\frac{\pi k}{N}))$$

i

k

$$(0.2\pm 4.8)\times 10^{-4}$$

$$(-0.1\pm 4.7)\times 10^{-4}$$
 
1

0
 
$$(-9.5\pm 4.8)\times 10^{-4}$$

$$1.0\pm 0.9$$

1

1
 
$$(-9.8\pm 4.8)\times 10^{-4}$$

$$1.0\pm 0.9$$

1

1
 
$$(-4.0\pm 0.5)\times 10^{-3}$$

$$1.0\pm 0.2$$

1

2
 
$$(-4.0\pm 0.5)\times 10^{-3}$$

$$1.0\pm 0.2$$

1

2
 
$$(-9.0\pm 0.5)\times 10^{-3}$$

$$1.0\pm 0.1$$

1

3
 
$$(-9.0\pm 0.5)\times 10^{-3}$$

$$1.0\pm 0.1$$

1

3
 
$$(-1.60\pm 0.05)\times 10^{-2}$$

$$0.99\pm 0.06$$

1

4
 
$$(-1.60\pm 0.05)\times 10^{-2}$$

$$0.99\pm 0.06$$

1

4
 
$$(-2.49\pm 0.05)\times 10^{-2}$$

$$0.99\pm 0.04$$

1

5
 
$$(-2.49\pm 0.05)\times 10^{-2}$$

$$0.99\pm 0.04$$

1

5
 
$$(-3.58\pm 0.05)\times 10^{-2}$$

$$0.99\pm 0.03$$

1

6
 
$$(-3.58\pm 0.05)\times 10^{-2}$$

$$0.99\pm 0.03$$

1

6
 
$$(-4.86\pm 0.05)\times 10^{-2}$$

$$0.99\pm 0.02$$

1

7
 
$$(-4.86\pm 0.05)\times 10^{-2}$$

$$0.99\pm 0.02$$

1

7
 
$$(-6.33\pm 0.05)\times 10^{-2}$$

$$0.99\pm 0.01$$

1

8
 
$$(-6.33\pm 0.05)\times 10^{-2}$$

$$0.99\pm 0.01$$

1

8
 
$$(-7.98\pm 0.05)\times 10^{-2}$$

$$0.99\pm 0.01$$

1

9
 
$$(-7.98\pm 0.05)\times 10^{-2}$$

$$0.99\pm 0.01$$

1

9

$$(-8.53\pm 0.02)\times 10^{-2}$$

$$(-8.53\pm 0.01)\times 10^{-2}$$
 
2

0
 
$$(-8.63\pm 0.04)\times 10^{-2}$$

$$1.0\pm 0.6$$

2

1
 
$$(-8.63\pm 0.04)\times 10^{-2}$$

$$1.0\pm 0.5$$

2

1

$$(-8.6\pm 0.2)\times 10^{-2}$$

$$(-8.6\pm 0.2)\times 10^{-2}$$
 
3

0

$$(-8.6\pm 0.2)\times 10^{-2}$$

$$(-8.7\pm 0.1)\times 10^{-2}$$
 
4

0


Lyapunov exponents are listed in descending order for plane waves using parameter set FK2 and diffusion in all model variables according to Eq. (3.2a, 3.2b, 3.2c). The first column contains the Lyapunov exponents for a one-dimensional cable (cf. Fig. 3.3), the second those for the full two-dimensional system. The third column indicates the ratio of the observed shift of the Lyapunov exponents to that one expected from Eq. (3.3). The fourth and fifth column denote the values of $$i$$ and $$k$$ used for Eq. (3.3)

In Ref. [6], Heagy et al. note that the case of non-complete diffusive coupling is not analytically tractable. This corresponds to the case of the original system (2.​40), since only the $$u$$ variables of neighboring cables are coupled. However, even in this situation, the Lyapunov spectra of the plane wave and of the traveling wave in a cable are still closely connected by the spectrum of the Laplace operator: The third column in Table 3.2 indicates that all additional Lyapunov exponents between $$\lambda _1^\text {1d}$$ and $$\lambda _2^\text {1d}$$ differ from those expected from Eq. (3.3) only by an approximately constant factor. Also, the two-fold degeneracy of the Lyapunov exponents is preserved. Indeed, even the spatial structure of Lyapunov vectors depicted in Figs. 3.1 and 3.2 resembles that expected from Eq. (3.6), namely different oscillatory modulations of the first Lyapunov vector. The similarity of the two scenarios (complete coupling and $$u$$-coupling) indicates the dominant role of the activation variable $$u$$ for the formation of wave patterns in excitable media. After all, activation (or, in the Fenton-Karma model: action potentials) can only be initiated from the resting state by a perturbation to $$u$$, but not via any other variable.


Table 3.2
Lyapunov spectrum of a plane wave for single-variable coupling
































































































































































$$\lambda _i^\text {1d}=\lambda _i^0$$

$$\lambda ^\text {2d}=\lambda _i^k$$

$$(\lambda _i^k-\lambda _i^0)/(-4c\sin ^2(\frac{\pi k}{N}))$$

i

k

$$(-0.8\pm 2.6)\times 10^{-4}$$

$$(-0.8\pm 2.6)\times 10^{-4}$$
 
1

0
 
$$(-1.5\pm 0.3)\times 10^{-3}$$

$$1.4\pm 0.5$$

1

1
 
$$(-1.5\pm 0.3)\times 10^{-3}$$

$$1.4\pm 0.5$$

1

1
 
$$(-5.9\pm 0.3)\times 10^{-3}$$

$$1.4\pm 0.1$$

1

2
 
$$(-5.9\pm 0.3)\times 10^{-3}$$

$$1.4\pm 0.1$$

1

2
 
$$(-1.32\pm 0.03)\times 10^{-2}$$

$$1.45\pm 0.06$$

1

3
 
$$(-1.32\pm 0.03)\times 10^{-2}$$

$$1.45\pm 0.06$$

1

3
 
$$(-2.35\pm 0.03)\times 10^{-2}$$

$$1.46\pm 0.03$$

1

4
 
$$(-2.35\pm 0.03)\times 10^{-2}$$

$$1.46\pm 0.03$$

1

4
 
$$(-3.70\pm 0.03)\times 10^{-2}$$

$$1.47\pm 0.02$$

1

5
 
$$(-3.70\pm 0.03)\times 10^{-2}$$

$$1.47\pm 0.02$$

1

5
 
$$(-5.37\pm 0.03)\times 10^{-2}$$

$$1.49\pm 0.01$$

1

6
 
$$(-5.37\pm 0.03)\times 10^{-2}$$

$$1.49\pm 0.01$$

1

6
 
$$(-7.40\pm 0.03)\times 10^{-2}$$

$$1.51\pm 0.01$$

1

7
 
$$(-7.40\pm 0.03)\times 10^{-2}$$

$$1.51\pm 0.01$$

1

7

$$(-8.6\pm 0.5)\times 10^{-2}$$

$$(-8.6\pm 0.6)\times 10^{-2}$$
 
2

0

$$(-8.6\pm 0.5)\times 10^{-2}$$

$$(-8.6\pm 0.6)\times 10^{-2}$$
 
3

0

$$(-8.6\pm 0.5)\times 10^{-2}$$

$$(-8.6\pm 0.6)\times 10^{-2}$$
 
4

0

$$(-8.6\pm 0.5)\times 10^{-2}$$

$$(-8.6\pm 0.6)\times 10^{-2}$$
 
5

0

$$(-8.6\pm 0.5)\times 10^{-2}$$

$$(-8.6\pm 0.5)\times 10^{-2}$$
 
6

0

$$(-8.6\pm 0.6)\times 10^{-2}$$

$$(-8.6\pm 0.6)\times 10^{-2}$$
 
7

0

$$(-8.6\pm 0.5)\times 10^{-2}$$

$$(-8.6\pm 0.5)\times 10^{-2}$$
 
8

0

$$(-8.6\pm 0.5)\times 10^{-2}$$

$$(-8.6\pm 0.5)\times 10^{-2}$$
 
9

0

$$(-8.6\pm 0.5)\times 10^{-2}$$

$$(-8.6\pm 0.5)\times 10^{-2}$$
 
10

0


Lyapunov exponents are listed in descending order for plane waves in the original Fenton-Karma model of Eq, (2.​40) using parameter set FK2. The first column contains the Lyapunov exponents for a one-dimensional cable (cf. Fig. 3.3), the second those for the full two-dimensional system. The third column indicates the ratio of the observed shift of the Lyapunov exponents to that one expected from Eq. (3.3). The fourth and firth column denote the values of $$i$$ and $$k$$ used for Eq. (3.3)



3.1.2 Rigidly Rotating Spiral Waves


Another type of attractor that can be found in various examples of excitable media (with no-flux boundary conditions) is that of a single spiral wave. In its simplest form of a rigidly rotating wave, the trajectory of the system corresponds to a periodic orbit. Therefore, at a first glance and from the dynamical systems point of view, the attractor seems to be similar to the one considered in Sect. 3.1.1. However, it is known that spiral waves are characterized by phase singularities (see Sect. 2.​1.​7), a topological property which represents a localized wave source that does not require specific anatomical features, as opposed to the plane wave that requires the periodic medium considered in Sect. 3.1.1. This qualitative difference between the two attractors becomes manifest in the spectrum of Lyapunov exponents by the appearance of two additional Lyapunov exponents that are equal to zero: In Fig. 3.4, the subspace belonging to $$\lambda _1=\lambda _2=\lambda _3=0$$ for a spiral wave in the Barkley model is illustrated with three meaningful directions in tangent space. Due to the degeneracy, Lyapunov stability analysis yields three arbitrary Lyapunov vectors in this subspace. In Fig. 3.4, three orthonormalized linear combinations of these arbitrary Lyapunov vectors are depicted, indicating that the two additional zero exponents are a result of the translational symmetry in the system: If the boundary is far away from the rotational center of the spiral—i.e. the two are electrotonically decoupled—a shift of the wave in either of the two available directions will not decay but represents a neutrally stable perturbation that is picked up by the linear stability analysis. The first Lyapunov vector plotted in Fig. 3.4 (second column) corresponds to a perturbation along the trajectory of the system, i.e. parallel to $$\frac{\partial (u,v)}{\partial t}$$. This is the neutrally stable direction in tangent space expected for a periodic orbit in every autonomous dynamical system, in this case accelerating the dynamics towards a future state. Consequently, the other two linearly-independent directions indeed correspond to a shift of the spiral wave in two different directions. For example, the perturbation in the third column of Fig. 3.4 accelerates wave parts moving (roughly) to upper left of the medium and decelerates those parts moving to the lower right, effectively shifting the whole spiral towards the upper left.

A328193_1_En_3_Fig4_HTML.gif


Fig. 3.4
Lyapunov vectors of a rigidly rotating spiral wave. Depicted are the medium state (first column) and three Lyapunov vectors (second to fourth column) spanning the degenerated subspace for $$\lambda _1=\lambda _2=\lambda _3=0$$ (values obtained in the simulation indicated below each column). Linear combinations were chosen such that the second column represents a shift along the trajectory of the system. The third and fourth column correspond to translational Lyapunov vectors that are perpendicular both to each other and to the second column. Simulations done using parameter set B2, see p. 184. The numerical space step is $$h=1/6$$, the domain size $$20\times 20$$ and the time step $$\Delta t=6.25\times 10^{-4}$$. Lyapunov vectors were allowed to align for a time span of $$35$$ time units, before the Lyapunov spectrum was calculated for $$350$$ time units (roughly 100 spiral periods) according to Sect. 2.​2.​6

The periodic orbit of a steadily rotating spiral therefore seems to be embedded in an invariant manifold of total dimension three, where the two directions transverse to the flow represent the independence of the dynamics from position of the spiral tip. Although the finding that spiral waves are subject to a Euclidean symmetry (ignoring boundary effects) seems trivial, note that Lyapunov spectra are computed using a spatially and temporally discretized system. In contrast, the translational, Euclidean invariance of the dynamics is a continuous symmetry. A priori, it is at least questionable whether this continuous symmetry is also present in the discretized system. The analysis shown here confirms that this is indeed the case.

A328193_1_En_3_Fig5_HTML.gif


Fig. 3.5
Numerical artifacts in the Lyapunov spectrum (Barkley model). Plot of the upper Lyapunov spectrum of a stable spiral wave against the numerical resolution in the Barkley model. As expected, there is a limiting spectrum for large $$N/L$$ (good numerical resolution). Note the initially stable translational perturbations ($$\lambda _2$$ and $$\lambda _3$$) which become marginally stable as the resolution increases. Thus, at sufficiently high resolutions, the translational invariance of the PDE is preserved in the discretized system, whereas pinning of the spiral to the numerical grid is observed at low resolutions. Simulations done using parameter set B2, see p. 184. The numerical space step is indicated in the plot, the domain size is $$20 \times 20$$, the time step $$\Delta t=6.25\times 10^{-4}$$. Lyapunov vectors were allowed to align for $$35$$ time units, before the Lyapunov spectrum was calculated for $$350$$ time units (roughly 100 spiral periods) according to Sect. 2.​2.​6


A328193_1_En_3_Fig6_HTML.gif


Fig. 3.6
Numerical artifacts in the Lyapunov spectrum (Fenton-Karma model). Dependence of the upper Lyapunov spectrum of a stable spiral wave on numerical resolution in the smooth Fenton-Karma model. In contrast to the Barkley model (Fig. 3.5), the lower part of the spectrum ($$\lambda _i$$, $$i>3$$” src=”/wp-content/uploads/2016/11/A328193_1_En_3_Chapter_IEq208.gif”></SPAN>) shows increased stability for low resolutions. Note the different y-axis scalings in the <SPAN class=EmphasisTypeItalic>lower</SPAN> and <SPAN class=EmphasisTypeItalic>upper</SPAN> half of the plot. Simulations done using parameter set FK3 (see p. 184) with isotropic diffusion of <SPAN id=IEq209 class=InlineEquation><IMG alt=. The numerical space step is indicated in the plot, the domain size is $$3.125\,\text {cm}\times 3.125\,\text {cm}$$, the time step $$\Delta t=0.0125\,\text {ms}$$. Lyapunov vectors were allowed to align for $$2{,}800\,\text {ms}$$, before the Lyapunov spectrum was calculated for $$2{,}800\,\text {ms}$$ (roughly 22 spiral periods) according to Sect. 2.​2.​6

The correctness of the numerical approximation in this particular aspect (with respect to the original PDE) depends on the chosen numerical resolution, as shown in Fig. 3.5. For high resolutions, the upper Lyapunov spectrum becomes independent of the numerical resolution and the structure of the neutrally stable subspace is preserved. For low resolutions, the Euclidean symmetry is broken and the neutrally stable subspace becomes one-dimensional: certain positions of the spiral wave are favored and stabilized by the numerical grid, hence yielding negative Lyapunov exponents for the translational perturbations. In other words, the spiral wave tends to pin to the numerical lattice. This effect has already been noticed by comparing spiral tip trajectories for different numerical resolutions [7]. For stable spiral waves without a visible core (i.e. with a stationary tip), however, this strategy is inappropriate, since a retrospective analysis cannot determine whether or not the tip could potentially have been located at a slightly different position. The plot in Fig. 3.5 also indicates that a low numerical resolution does not have identical effects on all directions in tangent space: in contrast to $$\lambda _2$$ and $$\lambda _3$$, higher-index Lyapunov exponents increase towards lower resolutions, meaning that the corresponding perturbations become less stable. This effect is, however, not universal: Fig. 3.6 indicates that for specific parameters in the Fenton-Karma model, a low numerical resolution has a stabilizing effect on $$\lambda _i$$, $$i>3$$” src=”/wp-content/uploads/2016/11/A328193_1_En_3_Chapter_IEq217.gif”></SPAN> as well. Other authors have observed that a coarse numerical resolution can both induce [<CITE><A href=7, 8] and prevent [9, 10] spiral wave break-up, which would be associated with one or more Lyapunov exponents being elevated above or reduced below zero, respectively. In particular for investigating control strategies for spiral wave activity in excitable media, Lyapunov stability analysis therefore presents a valuable tool to directly assess the accuracy of the numerical approximation. In this way, the faithful reproduction of stability properties in the numerical realization of the PDE can be ensured without relying on the traditional method of checking the sensitivity of indirect indicators to changes in resolution, such as spiral tip trajectories or periods, conduction velocity, action potential duration and the rate of change of the membrane potential during the upstroke phase [11]. The big advantage of the analysis lies in its implicit check for possible perturbations, whereas all other indicators mentioned above only operate on the attractor itself.


3.1.3 Multiple Spiral Waves


In Sect. 3.1.2, it was shown that Lyapunov stability analysis can detect symmetries in an activation pattern. At a second glance, this is quite surprising for the translational symmetries, since the domain under consideration (a bounded region with no-flux boundary conditions) does not have this symmetry. Consequently, it can be concluded that the part of the activation pattern responsible for the observed zero exponents (and possibly also for other exponents in the upper Lyapunov spectrum) is actually much smaller than the domain size and thus protected from the influence of the boundary.2 This does not contradict the extension of the corresponding Lyapunov vectors throughout the whole domain (see Fig. 3.4), as translations and rotations—the latter being equal to a perturbation along the trajectory—nevertheless act globally, as long as the spiral wave can reach the whole domain.

A328193_1_En_3_Fig7_HTML.gif


Fig. 3.7
Lyapunov exponents of multiple stable spirals. The number of spirals can be detected from the Lyapunov spectrum: a Exemplary snapshots of the activator variable $$u$$ of the Barkley model from simulations with one to twelve spirals exhibiting periodic dynamics. Black color indicates activation. Positions of phase singularities are marked by red dots. b Number of zero Lyapunov exponents for the above simulations. Every spiral contributes three such exponents, indicating the waves’ mutual independence in terms of angle and position (all symmetries are preserved individually). c Lyapunov spectrum of 50 leading exponents plotted against the number of spirals. Dashed lines indicate change of Lyapunov exponents for each additional spiral, Lyapunov exponents are colored randomly by their index. Numbers above gray bands indicate, how many Lyapunov exponents have a value within the corresponding interval of growth rates. Counts that are affected by the limited number of Lyapunov exponents (50) have been replaced by “x”. Simulation done using parameters B2, see p. 184. The numerical space step is $$h=1/5$$, the domain size $$30\times 30$$ and the time step $$\Delta t=0.002$$. Lyapunov vectors were allowed to align for a time span of $$50$$ time units, before the Lyapunov spectrum was calculated for another $$350$$ time units (roughly 100 spiral periods) according to Sect. 2.​2.​6

One way to assess the extension of the essential part of the spiral in terms of symmetry and stability would be to investigate the dependence of the spectrum on system size. Here, a different approach is chosen: different numbers of spiral waves are initiated in a medium of constant size, using the Barkley model with parameter set B2 (see Table A.1). Figure 3.7a shows example snapshots of the $$u$$-variable of the Barkley model for one to twelve spiral waves. The images correspond the state of the medium after an initial transient phase of spiral movement and possible disappearance of phase singularities from the medium. From the plot in Fig. 3.7b, it can be concluded that the specific signature of three zero Lyapunov exponents for a spiral wave must be carried by a relatively small region around the phase singularity. For one to twelve spiral waves, the number of zero Lyapunov exponents corresponds exactly to three times the number of spirals in the domain. This means that all the symmetries are preserved for each spiral individually. More specifically, this rules out the possibility that any of the spiral waves are phase-synchronized or that the position of one phase singularity is constrained by interaction with the surrounding ones (in which case some zero Lyapunov exponents should become negative). As indicated in Fig. 3.7c, this additive behavior of the Lyapunov spectrum does in fact extend to the next (more negative) two pairs of Lyapunov exponents. For the number of spirals and parameters considered here, the entire leading Lyapunov spectrum for $$n$$ spirals is therefore identical to that one expected for $$n$$ copies of a system with one spiral wave. This is remarkable, since different $$n$$ actually represent coexisting attractors in the same system. On the one hand, the behavior of the Lyapunov spectrum reveals that these attractors have a very specific structure. On the other hand, it confirms the intuitive interpretation of spiral waves in excitable media as objects.

A328193_1_En_3_Fig8_HTML.gif


Fig. 3.8
Single-spiral perturbations in a system of multiple spirals. Multiple spiral waves on a periodic orbit as independent entities. Along with the state of a medium containing three periodically rotating waves (first column), four specific directions in the degenerate subspace of neutral perturbations are shown. $$\delta (u,v)_\text {traj}$$ is the perturbation along the trajectory of the complete system. $$\delta (u,v)_{{\text {spiral}}{\{1,2,3\}}}$$ are perturbations advancing each spiral individually in time, thereby separating the areas controlled by different spiral waves. Numerical and model parameters are identical to those used in Fig. 3.7

To emphasize that the directions in the more and more degenerated subspaces in tangent space can indeed be assigned to individual spiral waves, Fig. 3.8 shows four specific Lyapunov vectors from the subspace with growth rate zero in a three-spiral simulation. The first one corresponds to the perturbation along the trajectory, while the other three indicate that the time- or phase-shifting symmetry is indeed preserved for each spiral wave individually, as if they were (almost) independent systems. As an interesting by-product, these specific vectors yield a decomposition of the medium into domains controlled by the individual spirals. At least for this periodic attractor, Lyapunov stability analysis therefore provides a way to objectively define “domains of influence” of individual spirals in contrast to visual inspection of the dynamics [12].

It should be noted that the complete preservation of symmetries does not imply that phase singularities in the medium do not interact: as mentioned above, the simulations were allowed to converge to a state with stationary phase singularities exhibiting periodic dynamics. In this transient time, spiral waves can have interacted with each other or pushed each other out of the medium. All the above analysis means is that the phase singularities have settled down outside their mutual interaction ranges, probably because for this specific model and parameters B1, the spiral waves are repelling each other. In contrast, the formation of pairs or clusters of spiral waves that settle down in a minimum of an attracting potential could have been detected, provided the force survives the linearization of the original equations.


3.1.4 Transition to Meandering


One classical use case for Lyapunov exponents is bifurcation analysis: Bifurcations are qualitative changes of the attractor structure along a certain path in parameter space. Locally in phase space, certain directions change their stability properties from stable to unstable or vice versa. Therefore, exactly at the bifurcation, there are typically one or more additional neutrally stable directions that can be detected via Lyapunov stability analysis (and yield additional zero Lyapunov exponents). A typical and well-understood bifurcation in excitable media is the transition from a rigidly rotating spiral to a meandering spiral, which has been shown to occur via a Hopf bifurcation [1315] introducing a second frequency into the system. D. Barkley obtained this important finding by carrying out a linear stability analysis on the fixed point representing a rigidly rotating spiral in the co-rotating frame of constant angular velocity. In this frame of reference, two eigenvalues of the Jacobian (at the fixed point) cross the imaginary axis at the critical bifurcation parameter value.

A328193_1_En_3_Fig9_HTML.gif


Fig. 3.9
Bifurcation detection. Transition from a rigidly rotating spiral ($$a \gtrsim 0.760$$) to a meandering spiral ($$a \lesssim 0.760$$) for parameter set B3, see p. 184. Spiral tip paths detected by the method outlined in Sect. 2.​2.​5 with $$(u^*,v^*)=(0.5,0.4)$$ are shown above the plot centered at the corresponding values of $$a$$. The bifurcation point can be clearly detected from the spectrum of Lyapunov exponents. For meandering, the degenerate subspace of growth rate zero becomes four- instead of three-dimensional, as expected from the Hopf nature of the bifurcation. Inset plot enlarged around the bifurcation point. The numerical space step is $$h=1/8$$, the domain size $$18.75\times 18.75$$ and the time step $$\Delta t=0.002$$. Lyapunov vectors were allowed to align for a time span of $$80$$ time units, before the Lyapunov spectrum was calculated for $$700$$ time units (roughly 200 spiral periods independent of $$a$$) according to Sect. 2.​2.​6


A328193_1_En_3_Fig10_HTML.gif


Fig. 3.10
Meandering modes for a rigidly rotating spiral. Bifurcating subspace in tangent space for parameter set B3 with $$a=0.766$$, i.e. before the meandering bifurcation (c.f. Fig. 3.9). Shown is the medium state (leftmost column) and are two arbitrary perpendicular vectors in the degenerated subspace belonging to the fourth and fifth Lyapunov exponent (values indicated below each column)

Figure 3.9 shows the Lyapunov spectrum for a single spiral wave in the Barkley model using parameter set B3. The transition to meandering occurs from right to left for a parameter value of $$a\approx 0.760$$ and is identical to the one considered in a publication by Barkley [14]. Towards the bifurcation point (from the right), two Lyapunov exponents approach zero, corresponding to the two-dimensional eigenspace for the fixed-point problem considered by Barkley. At the bifurcation, the rigid rotation of the spiral becomes unstable via the above-mentioned Hopf bifurcation. Since Lyapunov spectra only characterize the attractor the system is currently exploring, the two Lyapunov exponents do not become positive (as the real parts of the eigenvalues for the fixed-point analysis do). Instead, after the bifurcation (towards the left), there are now four zero Lyapunov exponents: two for the periodic or quasi-periodic dynamics on a torus and two—as before—as a result of the Euclidean symmetry (see Sects. 3.1.2 and 3.1.3). The Lyapunov vectors belonging to the two exponents approaching zero towards the critical parameter value correspond to the bifurcating directions in tangent space. Thus, the two vectors shown in Fig. 3.10 span the two-dimensional bifurcating subspace. As the Lyapunov vectors depicted in Fig. 3.4, which represent the Euclidian symmetry, the Lyapunov vectors for the meandering modes are concentrated at the wave front and back. However, while the symmetry modes have an angular periodicity of the wave front modulation of approximately $$2\pi $$ (because of its action as a spatial shift), the meandering modes exhibit a higher spatial frequency. Note that this is similar to the increasing spatial frequency for higher-index modes for plane waves as depicted in Fig. 3.1.

Of course, Lyapunov stability analysis cannot replace linear stability analysis of the corresponding fixed point problem: the most important disadvantage is that the imaginary parts of the eigenvalues cannot be reconstructed from the Lyapunov exponents. Furthermore, as mentioned above, Lyapunov analysis is always restricted to the attractor the system is currently evolving on, while a linear stability analysis can determine the stability properties of arbitrary points in phase space, no matter whether they are stable or not.3 The latter might, however, also be considered an advantage, since the fixed points (or other manifolds) do not have to be determined separately from the stability analysis. As an example, the angular wave speed of the co-rotating frame along with the fields of $$u$$, $$v$$ and $$w$$ of the Fenton-Karma model are obtained using a Newton-Raphson method in Ref. [16], and the dominant eigenvalues and eigenmodes at the fixed point have to be determined by further sophisticated methods. In contrast, Lyapunov exponents and vectors can be determined by standard procedures without any further algorithmic effort. Note also that the ability to detect and characterize this particular meandering bifurcation via linear stability analysis is based on the geometrically simple trajectory from which it originates, which makes it easily possible to transform a periodic orbit into a fixed point. While this can in the general case always be done by defining appropriate Poincaré sections, switching from the continuous to a time-discrete system might hide important details. In contrast, Lyapunov analysis could also be applied for meandering transitions where the (non-meandering) spiral trajectory is not just a circular motion.


3.1.5 Circular Heterogeneities


So far, the stability analysis has been applied to homogeneous excitable media, only. However, an important problem in the context of the control and development of undesired activation patterns in excitable media is how heterogeneity influences their stability (see introduction, Sect. 1.​6). One more specific question one might ask is: If a spiral wave is prone to some sort of instability, does heterogeneity stabilize or destabilize it?

The scenario in which this question will be answered uses a specific parameter range in the Barkley model, in which $$a$$ and $$\epsilon $$ are fixed, whereas the excitation threshold $$b$$ is varied. Figure 3.11 shows the Lyapunov spectra for a spiral wave in a homogeneous medium with the different values of $$b$$. For the whole range, the spiral wave rotates steadily, indicated by three Lyapunov exponents which are equal to zero (similar to the spiral in Sect. 3.1.4 before the meandering bifurcation, see Fig. 3.9). The two Lyapunov exponents associated with meandering come very close to zero towards $$b\approx 0.13$$, but instead of bifurcating, they stabilize again for further increasing $$b$$. Finally, for very high $$b$$, the medium loses its ability to form spiral waves and so the Lyapunov spectrum becomes all negative, indicating a stable fixed point which corresponds to the stable, homogeneous resting state of the medium.

A328193_1_En_3_Fig11_HTML.gif


Fig. 3.11
Lyapunov spectrum for different homogeneous excitation thresholds. Parameter set B4 (see Table A.1, p. 184) is used with the indicated value of $$b$$ throughout the medium. The range of the excitation threshold $$b$$ covers all values used in the heterogeneous scenario (cf. Fig. 3.12). The numerical space step is $$h=1/3$$, the domain size $$40\times 40$$ and the time step $$\Delta t=0.0025$$. Lyapunov vectors were allowed to align for a time span of $$35$$ time units, before the Lyapunov spectrum was calculated for $$150$$ time units according to Sect. 2.​2.​6

The heterogeneity that will be considered is a circular region of radius $$R$$ close to the spiral tip, in which the excitation threshold is increased from $$b=0.08$$ (which is used in the rest of the medium) to $$b_\mathrm{het}$$. It is well established that spiral waves are usually drawn to regions of longer spiral wave periods or lower excitability [1719], although exceptions are known [20]. For the parameters used here, the expected behavior is indeed observed and the spiral wave is attracted by the heterogeneity if initiated close to it. A snapshot of the long time behavior for an exemplary simulation is shown in Fig. 3.12a. From Sect. 3.1.2 it is known that usually, two Lyapunov exponents which are equal to zero indicate the translational symmetry of the activation pattern. Consequently, for a heterogeneity which pins the spiral wave to a specific position as in Fig. 3.12a, these two exponents should become negative, similar to the pinning effect of the numerical grid observed in Fig. 3.5.

However, a parameter scan for $$b_\mathrm{{het}}$$ and $$R$$ reveals that the behavior of the system is much richer than expected. Figure 3.12c shows the Lyapunov spectrum of the dynamics for a small heterogeneity of $$R=1.2$$ in model length units, plotted against the parameter $$b$$ within the heterogeneity. $$b_\mathrm{{het}}=0.08$$ corresponds to a homogeneous medium. As expected, the two Lyapunov exponents $$\lambda _2$$ and $$\lambda _3$$ indicating spatial symmetry become negative with increasing excitation threshold $$b$$. The heterogeneity thus seems to create one specific favorable location for the spiral and the “pinning force” to this location increases with the excitation threshold $$b_\mathrm{{het}}$$ (indicated by the more and more stable symmetry modes).4 One can already see in the same plot that the destabilization of the meandering modes which was visible for homogeneously increased $$b$$ (see Fig. 3.11) is not fully suppressed: $$\lambda _4$$ and $$\lambda _5$$, which were identified as the meandering modes in Sect. 3.1.2, are slightly destabilized, although not enough to lead to a bifurcation and less than in the homogeneous case (see Fig. 3.11). However, if the heterogeneity is larger, as is the case for Fig. 3.12d ($$R=2$$), exactly this happens. Not only is the rate with which the symmetry modes become stable much faster, but at the same time, the meandering modes are destabilized much stronger, leading to a bifurcation close to $$b_\mathrm{{het}}=0.115$$ which results in one additional zero Lyapunov exponent above $$b_\mathrm{{het}}=0.115$$ (similar to meandering in a homogeneous medium). Note that, for this to happen, the Lyapunov exponents which correspond to the symmetry and meandering modes have to reverse their order (see value of $$b$$ marked by “” in Fig. 3.12d). This will be discussed in detail at the end of this section. Visual inspection of the spiral tip trajectories confirms that up to $$b_\mathrm{{het}}=0.115$$, the spiral wave is indeed held in a fixed position by the heterogeneity, as seen in Fig. 3.12a, whereas above the bifurcation value, the tip starts to meander. This effect is seen more easily for an even larger heterogeneity ($$R=5$$), with a Lyapunov spectrum as plotted in Fig. 3.12e. The corresponding spiral tip trajectory for the meandering region with two zero Lyapunov exponents is shown in Fig. 3.12b.

A328193_1_En_3_Fig12_HTML.gif


Fig. 3.12
Symmetry break by heterogeneity. a, b Example snapshots of the Barkley model variable $$u$$ of spirals interacting with a circular heterogeneity of increased excitation threshold $$b$$ (region indicated by yellow color). Parameter set B4 (see Table A.1, p. 184) is used with $$b=0.08$$ throughout the medium, except for the circular region, whose parameter $$b_\mathrm{{het}}$$ is indicated by “” and “$$\times $$” in (d) and (e), respectively. The spiral tip path detected using the method outlined in Sect. 2.​2.​5 with $$(u^*,v^*)=(0.48,0.38)$$ is plotted as a red line. ce Leading Lyapunov spectra for three different radii of the heterogeneity ($$R=1.2, 2, 5$$, respectively) plotted against the value of the model parameter $$b=b_\mathrm{{het}}$$ within the circular region. $$b_\mathrm{{het}}=0.08$$ corresponds to a homogeneous medium. The numerical parameters are identical to those noted in the caption of Fig. 3.11

For small $$R$$, the heterogeneity therefore acts as a pinning center, despite the fact that a homogeneous medium of the same increased excitation threshold is close to a meandering instability (cf. Fig. 3.11). For larger $$R$$, however, this potential instability does not only reappear but is even enhanced, such that an actual bifurcation can take place which is not observed in the homogeneous case. Another interesting phenomenon for $$R=5$$ is that the symmetry modes are not stabilized for moderate values of $$b_\mathrm{{het}}$$, indicating that within the heterogeneity, the spiral can still evolve freely if the heterogeneity is not too strong. Also, note that the Lyapunov spectra for moderate $$b_\mathrm{{het}}$$ (below the bifurcation value) look very similar to the homogeneous case in Fig. 3.11. The value of $$b_\mathrm{{het}}$$ at which the spiral wave feels that the modification of $$b$$ is not global and deviates from the homogeneous behavior therefore increases with increasing $$R$$. This is consistent with the fact that the case of $$R\rightarrow \infty $$ is identical to the homogeneous scenario of Fig. 3.11.

For a stable spiral wave in a homogeneous medium it was shown in Sect. 3.1.2 that the leading Lyapunov spectrum consists of three neutrally stable modes which represent a perturbation along the trajectory of the system and spatial shifts in the two available directions due to Euclidian symmetry. The fourth and fifth Lyapunov exponent were associated with the modes which were found to be responsible for the meandering bifurcation considered in Sect. 3.1.4. However, as seen in this section, the relative order of the symmetry and meandering modes in terms of growth rate is not universal, but can be reversed by interaction of the spiral wave with a heterogeneity. In Fig. 3.12d, it is seen that this reversal can take place when both modes are actually stable, which makes it impossible to differentiate them by observing the dynamics of the spiral wave. By following the change of Lyapunov exponents from the homogeneous case ($$b_\mathrm{{het}}=0.08$$) to the value of $$b$$ indicated by “”, it was implicitly assumed that the trend in growth rate for both modes continues and that indeed the Lyapunov exponents of the meandering modes approach a value of zero (later confirmed when the bifurcation takes place at about $$b_\mathrm{{het}}=0.115$$). However, since the ordering of Lyapunov exponents from largest to smallest alone defines their indices, one would not be able to follow this (heuristic) reasoning if only the spectrum at $$b_\mathrm{{het}}=0.1$$ (“”) was known. Fortunately, this is possible making use of one quantity disregarded so far this section: the Lyapunov vectors. Figure 3.13 shows the Lyapunov vectors for the five leading exponents in the system with $$R=2$$ and $$b_\mathrm{{het}}=0.1$$. The first one corresponds to the perturbation along the trajectory, whereas the spatial symmetry modes can be found as the fourth and fifth Lyapunov vector. $$\delta (u,v)_4$$ represents a shift of the wave pattern towards the lower right, whereas $$\delta (u,v)_5$$ shifts the spiral wave towards the lower left. In contrast, the second and third Lyapunov vector exhibit a higher-frequency modulation of the wave front characteristic for the meandering modes (see Sect. 3.1.4). In this way, the reversal of stability order necessary for the pinned meandering bifurcation observed in this section can be objectively confirmed.

A328193_1_En_3_Fig13_HTML.gif


Fig. 3.13
Lyapunov vectors of a spiral wave interacting with a circular heterogeneity. Shown are the Lyapunov vectors corresponding to the first five Lyapunov exponents for the parameters indicated by “” in Fig. 3.12. The region of increased excitation threshold $$b$$ is marked by yellow color. The medium state is depicted in the leftmost column. The five columns each show the $$u$$ and $$v$$ components of the Lyapunov vectors. Values of Lyapunov exponents are printed below each column


3.1.6 Random Heterogeneities


One of the advantages of Lyapunov stability analysis compared to the linear stability analysis of fixed points is that it can be applied to arbitrary trajectories. Therefore, the heterogeneity in the system does not have to be “simple” for the method to work. Thinking about the complex system of the heart and its inherent heterogeneity, not only abrupt changes of tissue properties in localized regions are conceivable, as considered in Sect. 3.1.5. Rather, real-world (and in particular biological) systems are usually subject to natural variability which leads to a random distribution of parameters. Following this line of thoughts, the excitation threshold of the medium is now modified randomly. To compare the results with those in Sect. 3.1.5, random variations of $$b$$ from the base value of $$b_0=0.08$$ towards higher excitation thresholds are imposed, where a well-defined correlation length serves as a similar measure as the radius of the heterogeneity in Sect. 3.1.5. To construct such a randomly varying $$b(x,y)$$, so-called Gaussian random fields are utilized. These are fully defined by their auto-correlation function, which makes it particularly easy to obtain different realizations. If $$\sigma $$ is the intended correlation length, let


$$\begin{aligned} R(x,y)=\exp \left( \frac{x^2+y^2}{2\sigma ^2}\right) \end{aligned}$$

(3.7)
be the imposed (spatial) auto-correlation function. Via the Wiener-Khinchin theorem, the Fourier transform of $$R$$ is the power spectrum of the underlying field $$\beta _\text {full}(x,y)$$, so


$$\begin{aligned} \left| \mathcal F_2[\beta _\text {full}]\right| = \sqrt{\mathcal F_2[R]}, \end{aligned}$$

(3.8)
where $$\mathcal F_2$$ denotes the two-dimensional Fourier transform in space. As the Fourier transform of the field $$\beta _\text {full}$$ is only determined up to its amplitude by this relation, different realizations of the field can be obtained by choosing random phases $$\theta (k_x,k_y)$$ for each mode (numerically: for each computational point) and transforming the spectrum back to real space:


$$\begin{aligned} \beta _\text {full}=\mathrm {Re}\left( \mathcal F_2^{-1}\left[ \theta \cdot \sqrt{\mathcal F_2[R]}\right] \right) , \end{aligned}$$

(3.9)
where $$\mathrm {Re}$$ is the real part. The standard deviation of the obtained realization $$\beta _\text {full}(x,y)$$ across the whole domain is then normalized to one and negative regions are set to zero to give the final modulation $$\beta (x,y)$$:


$$\begin{aligned} \beta (x,y)=\frac{1}{\text {std}(\beta _\text {full})}{\left\{ \begin{array}{ll} 0 &{}\quad \beta _\text {full}(x,y)<0\\ \beta _\text {full}(x,y) &{}\quad \beta _\text {full}(x,y)>0 \end{array}\right. }\qquad \forall x,y \end{aligned}$$” src=”/wp-content/uploads/2016/11/A328193_1_En_3_Chapter_Equ12.gif”></DIV></DIV><br />
<DIV class=EquationNumber>(3.10)</DIV></DIV>The complete process of construction is illustrated in Fig. <SPAN class=InternalRef><A href=3.14. The field $$b(x,y)$$ for the simulation is then defined as


$$\begin{aligned} b(x,y)=b_0+(b_\text {target}-b_0)\beta (x,y) \qquad \forall x,y. \end{aligned}$$

(3.11)


A328193_1_En_3_Fig14_HTML.gif


Fig. 3.14
Construction of random excitation threshold. a Imposed two-dimensional gaussian auto-correlation function $$R(x,y)$$ with a defined correlation length $$\sigma $$ [see Eq. (3.7)]. b Amplitude of the spectrum of the underlying random field, via Wiener-Khinchin theorem from (a) as in Eq. (3.8). c Random phase $$\theta (x,y)$$ determining the realization of the field. d Realization of the field according to Eq. (3.9). e Normalized and capped version $$\beta (x,y)$$ of the field in (d) used to modify the excitation threshold. f Same as (e), but for a different realization, i.e. different phases in (c)

The number $$b_\text {target}$$ is by construction the value of $$b$$ which is one standard deviation of $$\beta _\text {full}$$ above the base value $$b_0$$ of the excitation threshold. It serves here as a measure for the amplitude of the heterogeneity and has a meaning similar to $$b_\mathrm{{het}}$$ in Sect. 3.1.5. Simulations for a range of correlation lengths $$\sigma $$ and target excitabilities $$b_\text {target}$$ were carried out. Because of the stochasticity of the system, for every fixed $$\sigma $$ and $$b_\text {target}$$ Lyapunov exponents were calculated using 40 different realizations of $$b(x,y)$$. The ensemble averages for three selected values of $$\sigma $$ along with two exemplary snapshots are shown in Fig. 3.15. For rather small-scale heterogeneities ($$\sigma =0.71$$), Fig. 3.15c shows that little more happens for increasing $$b_\text {target}$$ than the gradual stabilization of the Euclidian symmetry modes. This is in a way expected, because the spiral wave interacts with the first heterogeneity it encounters and pins to it. The gradual effect is a result of an increasing average amplitude of the heterogeneity with increasing $$b_\text {target}$$. The meandering modes are only slightly destabilized and Lyapunov exponents with even higher indices, $$\lambda _i$$, $$i>5$$” src=”/wp-content/uploads/2016/11/A328193_1_En_3_Chapter_IEq342.gif”></SPAN> (shown in gray), only increase for the extreme values of <SPAN id=IEq343 class=InlineEquation><IMG alt=. Note also that the standard deviation of the ensemble Lyapunov exponents increases towards higher $$b_\text {target}$$, because there is a larger variability in the amplitude of heterogeneities the spiral can encounter.

A328193_1_En_3_Fig15_HTML.gif


Fig. 3.15
Symmetry break by random heterogeneity. a, b Example snapshots of the Barkley model variable $$u$$ of spirals interacting with the random heterogeneity of increased excitation threshold $$b$$ (deviation $$b(x,y)-b_0$$ from the base value indicated by yellow color). Base parameters correspond to parameter set B4 (see Table A.1, p. 184) with $$b_0=0.08$$. The value of $$b_\text {target}$$ for the example snapshots is indicated by “” and “$$\times $$” in (c) and (e), respectively. The spiral tip path detected using the method outlined in Sect. 2.​2.​5 with $$(u^*,v^*)=(0.48,0.38)$$ is plotted as a red line. ce Leading Lyapunov spectra for three different correlation lengths $$\sigma $$ plotted against the strength $$b_\text {target}$$ of the heterogeneity. $$b_\text {target}=0.08$$ corresponds to a homogeneous medium. The plotted Lyapunov exponents are ensemble averages, error bars indicate standard deviation. The numerical space step is $$h=1/3$$, the domain size $$40\times 40$$ and the time step $$\Delta t=0.0025$$. Lyapunov vectors were allowed to align for a time span of $$60$$ time units, before the Lyapunov spectrum was calculated for $$525$$ time units according to Sect. 2.​2.​6

For heterogeneities on larger spatial scales ($$\sigma =1.56$$), Fig. 3.15d shows a similar behavior. However, the symmetry modes are stabilized more rapidly and the meandering modes are considerably more destabilized compared to the first case. Thus, although there is, on average, no bifurcation, some of the behavior for a discrete heterogeneity (cf. Fig. 3.12) is maintained. A prominent difference is seen in the higher-index Lyapunov exponents, which clearly increase for higher $$b_\text {target}$$, closing the otherwise large gap below the meandering modes. This is probably due to the heterogeneities with which the spiral tip is not directly interacting and which are distributed across the rest of the medium—a feature that does not exist for the discrete heterogeneities of Sect. 3.1.5.

Figure 3.15e indicates that all these features are preserved for even larger heterogeneities ($$\sigma =2.97$$), except for one: for small $$b_\text {target}$$, the pinning force (as measured by the stabilization of the symmetry modes) stays relatively weak, before intensifying from $$b_\text {target}\approx 0.11$$ upwards. This is again a somewhat masked version of the effect seen for large heterogeneities in Fig. 3.12e: There, no symmetry break was observed up to a certain threshold value of the heterogeneity amplitude. Hence, even for the random case, the larger the typical size of the heterogeneities, the stronger this heterogeneity needs to be to cause a considerable pinning effect, although the meandering modes are still destabilized immediately. The latter could make the spiral wave prone to other destabilizing factors, like a global parameter shift of the medium or random stimuli.

It should be emphasized that all the Lyapunov spectra considered in this section were ensemble averages, although it was confirmed that no meandering is observed for any single realization of heterogeneity. Nevertheless, it would be interesting to find out, whether the reversal of the growth rates for symmetry and meandering modes as observed in Sect. 3.1.5 can similarly occur in this generalized scenario. Unfortunately, the Lyapunov vectors which turned out to be very useful for circular heterogeneities are less indicative in this case. The first five Lyapunov vectors for the system whose parameters are marked by “$$\times $$” in Fig. 3.15 are shown in Fig. 3.16. While the first Lyapunov vector can clearly be identified as a perturbation along the trajectory, the next four Lyapunov vectors appear to be more or less random modulations of the wave front with no discernible regularity. This is likely to be due to the irregular propagation of the spiral wave, which causes fluctuating growth rates of which the Lyapunov exponents (which nevertheless have an exact two-fold degeneracy) are only averages. Thus, the Lyapunov vectors are also time-dependent beyond a simple rotation.

A328193_1_En_3_Fig16_HTML.gif


Fig. 3.16
Lyapunov vectors of a spiral wave interacting with random heterogeneity. Shown are the Lyapunov vectors corresponding to the first five Lyapunov exponents for the parameters indicated by “$$\times $$” in Fig. 3.15. The deviation $$b(x,y)-b_0$$ from the base value is indicated by yellow color. The medium state is depicted in the leftmost column. The five columns each show the $$u$$ and $$v$$ components of the Lyapunov vectors. Values of Lyapunov exponents are printed below each column

In summary, the above analysis shows that the interpretation of the Lyapunov spectrum for this specific case does not change dramatically when making the transition more realistic and asymmetric forms of heterogeneity. Considering the global nature of this heterogeneity (as opposed to the localized disk in Sect. 3.1.5), this implies that the modes with a Lyapunov exponent close to zero (for the homogeneous case) mainly respond to the interactions located at the spiral center. This is in good agreement with the localization of response functions in a region close to the spiral tip [21, 22]. Only for large and strong heterogeneities, the outer regions of the spiral wave become important, leading to a destabilization of the lower Lyapunov spectrum. This is intuitive to some extent, because these heterogeneities might eventually cause breakup in the waves sweeping over them. The fact that no meandering was observed for random heterogeneities suggests that the sharp drop of the excitation threshold for the single, isolated heterogeneity considered in Sect. 3.1.5 is a necessary prerequisite for this effect.


3.1.7 Heterogeneities in Spatio-Temporal Chaos


In Sects. 3.1.5 and 3.1.6, the interaction of a single spiral wave with heterogeneities of the medium parameters was examined. As outlined in the introduction (see Sect. 1.​4), spiral waves underlie tachycardias in the heart and their stability properties are of major interest because they may degenerate into a life-threatening state of spatio-temporal chaos. It is reasonable to assume that the chances of terminating such turbulent activity with gentle defibrillation methods depend on the complexity (or chaoticity) of the dynamics and that this complexity depends on the properties of the tissue and the degree of heterogeneity in the system. However, different consequences of heterogeneity are conceivable: On the one hand, it could increase complexity, because it adds diversity to the substrate. On the other hand, it could decrease complexity due to the additional spatial structure imposed. The third and probably most plausible possibility is that, assuming the whole range of heterogeneous parameters gives rise to spatio-temporal chaos, the chaotic dynamics leads to an average complexity in between those observed for homogeneous systems within the same range of parameters. Therefore, the subject of this last topic of Sect. 3.1 is the question, how spatial heterogeneities can alter the dynamics during spatio-temporal chaos beyond what can be inferred from individual homogeneous systems with different parameters.

For chaotic attractors, Lyapunov stability analysis provides a means to quantitatively assess complexity, namely the Lyapunov dimension $$D_\text {L}$$, which was introduced by J. A. Kaplan and J. L. Yorke [23]. It was conjectured by Kaplan and Yorke that for typical (non-pathological) systems, $$D_\text {L}$$ is equal to the Rényi dimension of order 1 [24], which is also called the information dimension and is a fractal dimension for the attractor in phase space. A direct calculation of the information dimension is almost impossible for very high-dimensional attractors, as it relies on the estimation of the asymptotic probability distribution of trajectories in phase space and the time until densities sufficient for this estimation are reached increases with increasing dimensionality. $$D_\text {L}$$, however, is a function of the spectrum of Lyapunov exponents and can thus easily be calculated, if the Lyapunov exponents are known. It is defined as


$$\begin{aligned} D_\text {L} = k+\frac{1}{|\lambda _{k+1}|}\underbrace{\sum _{i=1}^k\lambda _i}_*, \end{aligned}$$

(3.12)
where $$k$$ is the largest index such that the sum (*) is still non-negative. Consequently, for non-chaotic attractors with $$\lambda _i\le 0$$ for all $$i$$, the Lyapunov dimension $$D_L$$ is equal to the number of Lyapunov exponents which are equal to zero. Heuristically, in the general case, $$D_\text {L}$$ is the number of degrees of freedom necessary to describe the dynamics, such that the resulting set in phase space is, on average, neither expanding nor contracting. For this reason, it can be used to estimate an effective number of degrees of freedom for high- or infinite-dimensional systems. In fact, as noted in the introduction in Sect. 1.​7, in one of the two studies which have made use of Lyapunov exponents in excitable media so far, M. C. Strain and H. S. Greenside did exactly that [25]: In the Bär-Eiswrith model during spatio-temporal chaos, $$D_\text {L}$$ was found to be an extensive quantity and the average number of degrees of freedom per phase singularity was found to be a function of the model parameter $$\epsilon $$ in the chaotic regime. The following results build on this knowledge by considering heterogeneities in the same model parameter.

Simulations are carried out in a medium of size $$40\times 40$$ with periodic boundary conditions, using the Bär-Eiswirth model with parameters $$a=0.84$$, $$b=0.07$$ and $$\epsilon $$ varying between $$0.08$$ and $$0.12$$. For the whole range of $$\epsilon $$, spiral waves are known to break up leading to a state of spatio-temporal chaos [26]. In all cases, the activity is initiated by creating two spiral waves which subsequently break up as illustrated in Fig. 3.17a. Note that the initial condition is slightly asymmetric to prevent mirror symmetry which would persist even after breakup. The simulation is allowed to equilibrate for 1,500 model time units (roughly 300 dominant periods) before Lyapunov exponents and phase singularity statistics are obtained for another 1,500 time units. Phase singularities are detected every $$1$$ time units using the method introduced in Sect. 2.​1.​7 with $$(u^*,v^*)=(0.65, 0.45)$$. In contrast to Ref. [25], this fixed definition of phase is used independent of the value of $$\epsilon $$, to be able to obtain comparable results later in the heterogeneous case. The Lyapunov dimension $$D_\text {L}$$ is calculated for the obtained Lyapunov spectra according to Eq. (3.12). In order to assess the accuracy of the dimension estimation, the calculated Lyapunov spectrum is shifted up (down) by the error estimate introduced in Sect. 2.​2.​6 and the Lyapunov dimension is calculated for the shifted spectra to yield an upper (lower) bound for $$D_\text {L}$$.

A328193_1_En_3_Fig17_HTML.gif


Fig. 3.17
Spatio-temporal chaos in the Bär-Eiswrith model. a Snapshots of the model variables $$u$$ and $$v$$ for the beginning of an exemplary simulation showing the initiation protocol for spatio-temporal chaos (time progresses from left to right). Two spiral waves quickly break up and a turbulent state is formed. Red dots correspond to phase singularities detected by the method outlined in Sect. 2.​2.​5 with $$(u^*,v^*)=(0.65,0.45)$$. The value of $$\epsilon $$ is marked by “$$\times $$” in (b) and (c). b Mean number of phase singularities in simulations exhibiting spatio-temporal chaos, initiated as in (a), using parameter set BE1 (see Table A.1, p. 184) with different values of $$\epsilon $$. The number of phase singularities is calculated every 1 model time units from $$t=1{,}500$$ to $$t=3{,}000$$ to allow for equilibration. Error bars indicate standard deviation. c Kaplan-Yorke dimension $$D_\text {L}$$ for the same parameters as in (b). Error bars indicate the possible range of $$D_\text {L}$$ for Lyapunov spectra shifted by the error as calculated according to Sect. 2.​2.​6. The numerical space step is $$h=1/3$$, the domain size $$40\times 40$$ and the time step $$\Delta t=0.002$$. Lyapunov vectors were allowed to align for a time span of $$200$$ time units, before the Lyapunov spectrum was calculated for 1,500 time units (roughly 300 dominant periods) according to Sect. 2.​2.​6

Figure 3.17b shows that, in the parameter range considered here, neither the number nor the number fluctuation of phase singularities depends strongly on the parameter $$\epsilon $$.5 Roughly 10 phase singularities are present on average in the simulation domain. In contrast, as indicated by Fig. 3.17c, the Lyapunov dimension $$D_\text {L}$$ increases steadily with increasing $$\epsilon $$. This is plausible, as the transition to spiral breakup occurs just below $$\epsilon =0.07$$ [26] and is consistent with the results reported in Ref. [25]. Beyond the critical value, further increase of $$\epsilon $$ thus increases the chaoticity of the system.

What happens now if different values of $$\epsilon $$ from this range are combined in a single medium? Similarly to the results presented in Sect. 3.1.5, circular heterogneities are considered. Since multiple spiral waves are present in the system, not only one but several circular heterogeneities, in which the value of $$\epsilon $$ is modified, are inserted into the system. For simplicity, here, a regular grid of $$4\,\times \, 4$$ heterogeneities is chosen, such that the total number of 16 heterogeneities is on the same order as the average number of phase singularities as calculated in Fig. 3.17b. Simulations were carried out for three different levels $$\epsilon _\mathrm{{het}}$$ within the circular regions, whose radius radii $$R$$ was also varied. In the rest of the medium, $$\epsilon $$ was kept at its base value of $$0.08$$. Phase singularity statistics and the Lyapunov dimension $$D_\text {L}$$ are calculated as above. Snapshots of an examplary simulation for heterogeneities with $$\epsilon _\mathrm{{het}}=0.12$$ and $$R=3$$ and the time evolution of the number of phase singularities can be found in Fig. 3.18.

A328193_1_En_3_Fig18_HTML.gif


Fig. 3.18
Spatio-temporal chaos in a heterogeneous system. a Example snapshots of the Bär-Eiswirth model variable $$u$$ of spatio-temporal chaos interacting with circular heterogeneities, within which $$\epsilon $$ is increased to $$\epsilon _\mathrm{{het}}$$ (indicated by yellow color). Red dots correspond to phase singularities detected by the method outlined in Sect. 2.​2.​5 with $$(u^*,v^*)=(0.65,0.45)$$. The radius $$R=3$$ of the heterogeneities and $$\epsilon _\mathrm{{het}}=0.12$$ are indicated by “$$+$$” in Fig. 3.19. Parameter set BE1 (see Table A.1, p. 184) is used; outside the yellow regions $$\epsilon =0.08$$. b Number of phase singularities for the entire simulation time. Statistics (shown in Fig. 3.19) are calculated from $$t=1{,}500$$ onwards. The “$$+$$” at $$t=0$$ indicates that the data corresponds to the equally marked parameters in Fig. 3.19

Only gold members can continue reading. Log In or Register to continue

Stay updated, free articles. Join our Telegram channel

Nov 21, 2016 | Posted by in CARDIOLOGY | Comments Off on Results

Full access? Get Clinical Tree

Get Clinical Tree app for offline access