Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Chaotic dynamics creates and destroys branched flow

Alexandre Wagemakers Nonlinear Dynamics, Chaos and Complex Systems Group, Departamento de Física, Universidad Rey Juan Carlos, Tulipán s/n, 28933 Móstoles, Madrid, Spain alexandre.wagemakers@urjc.es    Aleksi Hartikainen Computational Physics Laboratory, Tampere University, Tampere 33720, Finland aleksi.hartikainen@tuni.fi    Alvar Daza Nonlinear Dynamics, Chaos and Complex Systems Group, Departamento de Física, Universidad Rey Juan Carlos, Tulipán s/n, 28933 Móstoles, Madrid, Spain Department of Physics, Harvard University, Cambridge, Massachusetts 02138, USA alvar.daza@urjc.es    Esa Räsänen Computational Physics Laboratory, Tampere University, Tampere 33720, Finland Department of Physics, Harvard University, Cambridge, Massachusetts 02138, USA esa.rasanen@tuni.fi    Miguel A.F. Sanjuán Nonlinear Dynamics, Chaos and Complex Systems Group, Departamento de Física, Universidad Rey Juan Carlos, Tulipán s/n, 28933 Móstoles, Madrid, Spain miguel.sanjuan@urjc.es
(June 14, 2024)
Abstract

The phenomenon of branched flow, visualized as a chaotic arborescent pattern of propagating particles, waves, or rays, has been identified in disparate physical systems ranging from electrons to tsunamis, with periodic systems only recently being added to this list. Here, we explore the laws governing the evolution of the branches in periodic potentials. On one hand, we observe that branch formation follows a similar pattern in all non-integrable potentials, no matter whether the potentials are periodic or completely irregular. Chaotic dynamics ultimately drives the birth of the branches. On the other hand, our results reveal that for periodic potentials the decay of the branches exhibits new characteristics due to the presence of infinitely stable branches known as superwires. Again, the interplay between branched flow and superwires is deeply connected to Hamiltonian chaos. In this work, we explore the relationships between the laws of branched flow and the structures of phase space, providing extensive numerical and theoretical arguments to support our findings.

preprint: APS/123-QED

I Introduction

Electrons, lasers, tsunamis, and ants have at least one thing in common: they all display branched flow [1, 2, 3, 4, 5]. Whenever a wave propagates through a weakly refracting medium, flow is expected to accumulate along certain directions, forming structures called branches. Among the different examples, tsunamis serve as the most dramatic illustration of the powerful implications of branched flow. Branches can carry an unusual amount of energy towards unpredictable points in space.

Despite being a ubiquitous phenomenon with remarkable effects, basic knowledge of the mechanisms behind branched flow is still lacking. It has been well established that branched flow occurs in the semiclassical limit [6]. This means that the wavelength should be small compared to the characteristic length of the refracting medium. Deflections must be small too, so that branched flow is mostly made of forward scattering. However, the precise relations between the flow patterns and the shape of the potentials producing them still remains unclear. Indeed, almost all the literature on branched flow focuses solely on the so-called random potentials. It is worth mentioning that the term random potential might be confusing, since branched flow is a completely deterministic process. However, given its wide use in the existing literature, here we refer to random potentials when talking about smooth potentials lacking any spatial order. In all the examples cited so far, waves evolve in these smooth landscapes with no perceptible spatial pattern.

Spatial disorder of the potential is not a requirement for the dynamics of branched flow. In fact, periodic potentials have recently been identified to produce branched flow, opening up a whole new realm of possibilities [7]. From superlattices such as twisted bilayer graphene [8] to photonic crystals [9], branched flow could be observed if the right conditions are met. Furthermore, in periodic potentials, some branches remain indefinitely stable, forming so-called superwires. Fig. 1 shows branched flow (in red) and superwires (in blue) as the result of a plane wave evolving in a periodic potential. In the superwires, it is the dynamics that prevent the spreading of the flow, unlike being energetically trapped as in a waveguide. In fact, superwires are related to stable dynamics, while branched flow is related to chaotic dynamics, as can be observed in the phase space portrait of Fig. 1. Potential applications of superwires include superconductivity and beam focusing.

Early works investigated the laws of branched flow only in random potentials [10, 11]. Therefore, it is natural to ask how these laws translate in the case of a periodic potential. Our goal in this work is to understand how branches emerge and disappear in periodic potentials. A simple visual inspection might indicate that branched flow looks similar at both random and periodic potentials, but the presence of superwires implies important differences in the dynamics that require detailed investigation. In this paper, we discuss how prior knowledge on branched flow needs to be revised for periodic media.

Refer to caption
Figure 1: Branched flow in the Fermi periodic potential of Eq. (12). In the left panel, red tones indicate branched flow related to chaotic trajectories (positive maximum Lyapunov exponent), whereas blue tones are used for superwires (zero maximum Lyapunov exponent). The right panel depicts the Poincaré section along the direction transverse to the potential y~~𝑦\tilde{y}over~ start_ARG italic_y end_ARG using the same color coding. Superwires are caused by KAM stable islands (in blue) while branched flow is related to the chaotic sea (in red).

The manuscript is organized as follows. First, the dynamics of branched flow and the numerical techniques used to characterize it are described in Sec. II. Our main findings are included in Sec. III, where we study the laws of the birth and death of the branches in periodic potentials. Finally, we discuss the implications and perspectives of our work in Sec. IV.

II Methods

Here, we detail the numerical and theoretical tools employed along the manuscript.

II.1 Semiclassical dynamics of branched flow

There are two conditions that make branched flow a semiclassical phenomenon [6]. First, the wavelength must be smaller than the characteristic length of the potential. Second, the potential height must be sufficiently small, allowing the wave to be gently deflected without backscattering:

λ𝜆\displaystyle\lambdaitalic_λ lpotmuch-less-thanabsentsubscript𝑙𝑝𝑜𝑡\displaystyle\ll l_{pot}≪ italic_l start_POSTSUBSCRIPT italic_p italic_o italic_t end_POSTSUBSCRIPT (1)
Vmaxsubscript𝑉𝑚𝑎𝑥\displaystyle V_{max}italic_V start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT Ewave.much-less-thanabsentsubscript𝐸𝑤𝑎𝑣𝑒\displaystyle\ll E_{wave}.≪ italic_E start_POSTSUBSCRIPT italic_w italic_a italic_v italic_e end_POSTSUBSCRIPT . (2)

In this regime, wave propagation can be replaced by ray-tracing of the appropriate manifolds. For example, a plane wave can be described as a set of parallel rays with fixed energy, or a point source can be modeled using rays coming out from a single point in space in every possible direction.

The small deflection condition still enables a further simplification. It is possible to neglect longitudinal dynamics and concentrate only in the transverse direction. In this work, we focus on two-dimensional branched flow. Therefore, by ignoring the longitudinal direction, we are reducing the problem to just one dimension. This is the quasi-2D approximation [10], where the motion associated to the transverse (y𝑦yitalic_y) and longitudinal (x𝑥xitalic_x) directions is described by

py˙=Vy˙subscript𝑝𝑦𝑉𝑦\displaystyle\dot{p_{y}}=-\frac{\partial V}{\partial{y}}over˙ start_ARG italic_p start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT end_ARG = - divide start_ARG ∂ italic_V end_ARG start_ARG ∂ italic_y end_ARG (3)
y˙=py˙𝑦subscript𝑝𝑦\displaystyle\dot{y}=p_{y}over˙ start_ARG italic_y end_ARG = italic_p start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT (4)
px˙=0˙subscript𝑝𝑥0\displaystyle\dot{p_{x}}=0over˙ start_ARG italic_p start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_ARG = 0 (5)
x˙=px=k.˙𝑥subscript𝑝𝑥𝑘\displaystyle\dot{x}=p_{x}=k.over˙ start_ARG italic_x end_ARG = italic_p start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT = italic_k . (6)

Here k𝑘kitalic_k is a constant that can be set equal to one without loss of generality and V(x,y)𝑉𝑥𝑦V(x,y)italic_V ( italic_x , italic_y ) represents the two-dimensional potential. By construction, there is no backscattering in the quasi-2D approximation.

Given the trivial dynamics along the longitudinal axis, we can reduce the model even further and get the a discrete map for the transverse dynamics

py(n+1)subscript𝑝𝑦𝑛1\displaystyle p_{y}(n+1)italic_p start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ( italic_n + 1 ) =py(n)(Vy)(x,y)=(xn,yn)absentsubscript𝑝𝑦𝑛subscript𝑉𝑦𝑥𝑦subscript𝑥𝑛subscript𝑦𝑛\displaystyle=p_{y}(n)-\left(\frac{\partial V}{\partial y}\right)_{(x,y)=(x_{n% },y_{n})}= italic_p start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ( italic_n ) - ( divide start_ARG ∂ italic_V end_ARG start_ARG ∂ italic_y end_ARG ) start_POSTSUBSCRIPT ( italic_x , italic_y ) = ( italic_x start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) end_POSTSUBSCRIPT (7)
y(n+1)𝑦𝑛1\displaystyle y(n+1)italic_y ( italic_n + 1 ) =y(n)+py(n+1).absent𝑦𝑛subscript𝑝𝑦𝑛1\displaystyle=y(n)+p_{y}(n+1).= italic_y ( italic_n ) + italic_p start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ( italic_n + 1 ) . (8)

This area-preserving map is usually referred to as the kick and drift map [1].

Refer to caption
Figure 2: Visible branch detection with the manifold tracking algorithm. Trajectories in a random correlated potential with correlation length lc=0.1subscript𝑙𝑐0.1l_{c}=0.1italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 0.1 are launched from a planar front wave represented in the right panel (i). Rays propagate downward. Panels in the left column are plots of the position y(t)𝑦𝑡y(t)italic_y ( italic_t ) of the rays at time t𝑡titalic_t ordered according to their starting position y(0)𝑦0y(0)italic_y ( 0 ). Panels in the central column represent an estimation of the rays density. The horizontal axis ρ(t,y)𝜌𝑡𝑦\rho(t,y)italic_ρ ( italic_t , italic_y ) holds for the density and the vertical axis is the coordinate y(t)𝑦𝑡y(t)italic_y ( italic_t ). The representation has been tilted 90ºitalic-º\textordmasculineitalic_º to match the manifold coordinates. The figure has been divided into 4 time frames to explain the cusps and branches formation. (a)-(b) At time t=0.1𝑡0.1t=0.1italic_t = 0.1 the manifold is nearly flat and the density of rays is almost uniform. (c)-(d) As the wavefront begins to propagate a time t=0.2𝑡0.2t=0.2italic_t = 0.2, the shape of the manifold in panel (c) begins to wiggle as rays move influenced by the potential. The density begins to peak (panel (d)) although it is not a visible branch. (e)-(f) At time t=0.27𝑡0.27t=0.27italic_t = 0.27 the cusp is formed. Its manifestation in the manifold is a null derivative marked with a red dot in (e) and a peak in ray density in (f). The part of the manifold painted in red corresponds to |y(t)/y(0)|<1𝑦𝑡𝑦01|\partial y(t)/\partial y(0)|<1| ∂ italic_y ( italic_t ) / ∂ italic_y ( 0 ) | < 1. Along with the zero derivative, it is our criterion to detect visible branches in the physical space. (g)-(h) The last time frame t=0.4𝑡0.4t=0.4italic_t = 0.4 shows the two branches born from the cusp. The rays’ density in (h) perfectly corresponds with the red portion of the manifold in (g). The two branches are clearly visible in panel (i).

II.2 Counting branches

Intuitively, branches could be defined as unusual accumulation of flow in certain areas of space. Nevertheless, it can be hard to find a precise way to quantify that intuition. Among the different measures proposed in the literature, we have chosen an improved version of the manifold tracking algorithm [12] because it requires the smallest number of arbitrary parameters to be adjusted and produces the most robust results. Next, we explain its functioning.

The manifold tracking algorithm comprises two stages. For each time step, the algorithm looks at the evolution of the transverse coordinate y(t)𝑦𝑡y(t)italic_y ( italic_t ) as a function of the corresponding initial condition y(0)𝑦0y(0)italic_y ( 0 ), as shown in the left column of Fig. 2. Manifolds evolving in a weak potential may undergo cusp catastrophes, which suppose an accumulation of rays in a single point. The density of the rays ρ(t,y)𝜌𝑡𝑦\rho(t,y)italic_ρ ( italic_t , italic_y ) shows an acute peak at the cusps, as shown in panel (f) of Fig. 2. These cusps could be viewed as the origin of the branches. In two dimensions, a cusp results in two caustics, or focal lines, corresponding to two peaks in the density ρ(t,y)𝜌𝑡𝑦\rho(t,y)italic_ρ ( italic_t , italic_y ), as depicted in Fig. 2-(h). The following condition is met for the caustics

(y(t)y(0))y=ycaustic=0.subscript𝑦𝑡𝑦0𝑦subscript𝑦caustic0\left(\frac{\partial y(t)}{\partial y(0)}\right)_{y=y_{\text{caustic}}}=0.( divide start_ARG ∂ italic_y ( italic_t ) end_ARG start_ARG ∂ italic_y ( 0 ) end_ARG ) start_POSTSUBSCRIPT italic_y = italic_y start_POSTSUBSCRIPT caustic end_POSTSUBSCRIPT end_POSTSUBSCRIPT = 0 . (9)

Hence, we examine the local maxima and minima in the plot of y(t)𝑦𝑡y(t)italic_y ( italic_t ) versus y(0)𝑦0y(0)italic_y ( 0 ) to determine the number of branches, marking the initial phase of the manifold tracking algorithm. The underlying concept is that the ray density ρ(t,y)𝜌𝑡𝑦\rho(t,y)italic_ρ ( italic_t , italic_y ) corresponds to the projection of the y(t)y(0)𝑦𝑡𝑦0y(t)-y(0)italic_y ( italic_t ) - italic_y ( 0 ) plot onto the vertical axis, with extremes yielding density peaks. We can see that for the cusps, besides Eq. 9, the following condition also applies

(2y(t)y(0)2)y=ycusp=0.subscriptsuperscript2𝑦𝑡𝑦superscript02𝑦subscript𝑦cusp0\left(\frac{\partial^{2}y(t)}{\partial y(0)^{2}}\right)_{y=y_{\text{cusp}}}=0.( divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_y ( italic_t ) end_ARG start_ARG ∂ italic_y ( 0 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) start_POSTSUBSCRIPT italic_y = italic_y start_POSTSUBSCRIPT cusp end_POSTSUBSCRIPT end_POSTSUBSCRIPT = 0 . (10)

The process of cusp formation continues indefinitely in the ray-tracing approximation; however, beyond a certain threshold, new cusps cease to generate visible branches. When the ray density accumulated by these cusps falls below the background density, they are engulfed by the diffusion process. Consequently, relying solely on cusp formation to identify branches becomes insufficient, necessitating additional criteria to account for the branches.

The second stage of the manifold tracking algorithm relies on the fact that the rays concentrate in a small region around the caustics. More precisely, we define a visible branch as a continuous piece of the manifold that has stretched less than a certain amount:

|δy(t)δy(0)||sy(0)(t)|<ε,𝛿𝑦𝑡𝛿𝑦0subscript𝑠𝑦0𝑡𝜀\left|\frac{\delta y(t)}{\delta y(0)}\right|\equiv|s_{y(0)}(t)|<\varepsilon,| divide start_ARG italic_δ italic_y ( italic_t ) end_ARG start_ARG italic_δ italic_y ( 0 ) end_ARG | ≡ | italic_s start_POSTSUBSCRIPT italic_y ( 0 ) end_POSTSUBSCRIPT ( italic_t ) | < italic_ε ,

where sy(0)(t)subscript𝑠𝑦0𝑡s_{y(0)}(t)italic_s start_POSTSUBSCRIPT italic_y ( 0 ) end_POSTSUBSCRIPT ( italic_t ) is the stretching experienced by a small piece of the manifold initially at y(0)𝑦0y(0)italic_y ( 0 ) after a time t𝑡titalic_t.

If the slope of a continuous piece of the manifold is small enough and contains a minimum of Nrsubscript𝑁𝑟N_{r}italic_N start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT rays, then the branch becomes visible.

Figure 2 illustrates the whole detection process. Panels (a), (c), (e), and (g) show the transverse position y(t)𝑦𝑡y(t)italic_y ( italic_t ) of the particle at four different times t𝑡titalic_t as a function of their initial position y(0)𝑦0y(0)italic_y ( 0 ). The cusps are located at the stationary points with null derivative and marked with red dots. The pieces of the manifold highlighted in red are the regions with |sy(0)(t)|<1subscript𝑠𝑦0𝑡1|s_{y(0)}(t)|<1| italic_s start_POSTSUBSCRIPT italic_y ( 0 ) end_POSTSUBSCRIPT ( italic_t ) | < 1. The pieces of manifold in red are the visible branches as long as they contain at least Nrsubscript𝑁𝑟N_{r}italic_N start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT rays. In Fig. 2 (b), (d), (f) and (h), we represent the ray density ρ(y,t)𝜌𝑦𝑡\rho(y,t)italic_ρ ( italic_y , italic_t ) as a function of the y𝑦yitalic_y axis. The dashed line corresponds to the position of the branches. Finally, panel Fig. 2-(i) displays the ray density in the physical space xy𝑥𝑦x-yitalic_x - italic_y. The branch detection depends now on two parameters: the threshold ε𝜀\varepsilonitalic_ε for the stretching of the manifold and the minimal number Nrsubscript𝑁𝑟N_{r}italic_N start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT of rays contained in the piece of the manifold. The number of branches detected is stable for reasonable variations of these parameters. We have chosen ε=1𝜀1\varepsilon=1italic_ε = 1 and Nr=3subscript𝑁𝑟3N_{r}=3italic_N start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = 3 for the rest of the study.

II.3 Potentials

As already mentioned in the Introduction, the main goal of this paper is to investigate the laws governing the birth and death of the branches in periodic potentials. Instead of focusing in a particular physical system or model, we aim to find general results applicable to any two-dimensional periodic system exhibiting branched flow. Therefore, the potentials here play the role of mere test subjects that allow us to explore the behavior of the branches.

Perhaps the most straightforward two-dimensional periodic potential that comes to mind is

V(x,y)=sinx+siny.𝑉𝑥𝑦𝑥𝑦V(x,y)=\sin x+\sin y.italic_V ( italic_x , italic_y ) = roman_sin italic_x + roman_sin italic_y . (11)

However, as discussed in [7], this integrable potential cannot produce branched flow. Nevertheless, this sine potential serves as a reference to compare and study the particular features of branched flow, we will refer to it as periodic integrable potential.

A more interesting periodic potential can be built through the so-called Fermi potential [13], which can be written as

V(r)=j=1NV0/[1+exp(|rr0j|/σ)],𝑉𝑟superscriptsubscript𝑗1𝑁subscript𝑉0delimited-[]1𝑟subscript𝑟0𝑗𝜎V(\vec{r})=\sum_{j=1}^{N}V_{0}/\left[1+\exp(|\vec{r}-\vec{r_{0j}}|/\sigma)% \right],italic_V ( over→ start_ARG italic_r end_ARG ) = ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / [ 1 + roman_exp ( | over→ start_ARG italic_r end_ARG - over→ start_ARG italic_r start_POSTSUBSCRIPT 0 italic_j end_POSTSUBSCRIPT end_ARG | / italic_σ ) ] , (12)

where the parameters V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and σ𝜎\sigmaitalic_σ determine the height and steepness of the potential respectively, and r0jsubscript𝑟0𝑗r_{0j}italic_r start_POSTSUBSCRIPT 0 italic_j end_POSTSUBSCRIPT represents the positions in a square lattice (although other periodic structures could be considered). This potential is non-integrable and previous works have already used it to investigate branched flow and other diffusive phenomena [7, 13]. In this case, chaotic and periodic dynamics are present in phase space depending on the initial conditions.

Finally, a random Gaussian correlated potential with correlation length lcsubscript𝑙𝑐l_{c}italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is also used as a benchmark:

V(r)=0delimited-⟨⟩𝑉𝑟0\langle V(\vec{r})\rangle=0⟨ italic_V ( over→ start_ARG italic_r end_ARG ) ⟩ = 0 (13)
V(r1)V(r2)=c(|r1r2|).delimited-⟨⟩𝑉subscript𝑟1𝑉subscript𝑟2𝑐subscript𝑟1subscript𝑟2\langle{V(\vec{r_{1}})V(\vec{r_{2}})}\rangle=c(|\vec{r_{1}}-\vec{r_{2}}|).⟨ italic_V ( over→ start_ARG italic_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ) italic_V ( over→ start_ARG italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ) ⟩ = italic_c ( | over→ start_ARG italic_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG - over→ start_ARG italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG | ) . (14)

Here c𝑐citalic_c denotes a correlation function c(r)=v02exp(r2/lc)𝑐𝑟superscriptsubscript𝑣02superscript𝑟2subscript𝑙𝑐c(r)=v_{0}^{2}\exp(-r^{2}/l_{c})italic_c ( italic_r ) = italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_exp ( - italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ). The branch statistics for this random potential have been well established [10, 11] and provide a reference point for our research.

III Results

Now, we present the main numerical and theoretical results concerning the laws of the birth and death of the branches in a periodic potential.

III.1 Birth of the branches

As a wave or bundle of rays propagates in a weak potential, cusps arise and branches follow. Thus, we must account for the apparition of cusps to understand the creation of the branches. In [10], it is argued that cusps grow exponentially with time in random potentials,

Neαt,proportional-to𝑁superscript𝑒𝛼𝑡N\propto e^{\alpha t},italic_N ∝ italic_e start_POSTSUPERSCRIPT italic_α italic_t end_POSTSUPERSCRIPT , (15)

where the parameter α𝛼\alphaitalic_α is related to the Lyapunov exponent and the inhomogeneity of the stretching.

We can try to understand this exponential growth in the context of periodic potentials by means of a simplified argument. Looking at Fig. 2, we can see how the weak potential concentrates the rays in the cusp, and then two branches emerge as a maximum and minimum in the y(t)𝑦𝑡y(t)italic_y ( italic_t ) versus y(0)𝑦0y(0)italic_y ( 0 ) plot. The total effect is that the set of initial conditions y(0)𝑦0y(0)italic_y ( 0 ) get transformed into something that resembles a sine function sin(y(0))𝑦0\sin(y(0))roman_sin ( italic_y ( 0 ) ). Given the periodicity of the potential, we can loosely assume that this action is repeated indefinitely, so that yn+1=μsin(yn)subscript𝑦𝑛1𝜇subscript𝑦𝑛y_{n+1}=\mu\sin(y_{n})italic_y start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT = italic_μ roman_sin ( italic_y start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ), where the index n𝑛nitalic_n refers to the unit cell of the periodic potential and μ𝜇\muitalic_μ to the characteristics of the potential. This one-dimensional map folds the initial conditions y0subscript𝑦0y_{0}italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT at each step and, for sufficiently large μ𝜇\muitalic_μ, the number of folds increases exponentially with n𝑛nitalic_n. In other words, the number of branches increases exponentially with time, since the local maxima and minima of ynsubscript𝑦𝑛y_{n}italic_y start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT versus y0subscript𝑦0y_{0}italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT correspond with the branches, as explained in Sec. II.2. This argument captures the exponential growth of the branches, although it fails in some aspects. In particular, there is a degeneracy in the location of the branches, since in the sine map all the maxima/minima have the same value of y𝑦yitalic_y, meaning that all the branches are concentrated in a single point. This happens because the dynamics is actually more complicated than what a one-dimensional map can capture. At least a two-dimensional map such as the kick and drift map of Eq. 8 is needed to faithfully reproduce the main features of branched flow. The dynamics of the kick and drift map stretches and folds the initial manifold in phase space producing several vertical segments after some time. At these points, we have that (pyy)=subscript𝑝𝑦𝑦\left(\frac{\partial p_{y}}{\partial y}\right)=\infty( divide start_ARG ∂ italic_p start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_y end_ARG ) = ∞, which is the cusp condition. If the dynamics is periodic, the cusps will get undone and then reappear again in the same positions after some time. Thus, for periodic motion, the number of cusps remains constant on average. However, if the dynamics is chaotic, the cusps are formed at new unpredictable positions produced by the kick and drift dynamics. This process continues as time goes by, creating a fractal structure which is ultimately responsible for the exponential growth of the cusps.

Refer to caption
Figure 3: Number of caustics over time for various potentials. Both the random potential described by Eq. 14 and the periodic Fermi potential defined by Eq. 12 exhibit comparable exponential growth patterns. Conversely, the integrable potential displays distinct behavior, appearing to reach an upper limit. Initially, the number of visible branches equals the number of caustics; however, this correlation diminishes over time. The caustics are detected with the manifold tracking algorithm [12].

We conducted numerical computations to observe the evolution of caustics over time using various potentials, aiming to validate our theoretical insights. We can observe in Fig. 3 that the number of caustics exhibits exponential growth for random potentials, as anticipated. Similarly, for non-integrable periodic potentials like the Fermi potential of Eq. (12), exponential growth is evident. However, for integrable periodic potentials, growth is constrained by an upper limit. This underscores that spatial randomness in the potential is not a prerequisite for branched flow. Chaotic dynamics is the necessary ingredient for branched flow, achievable with both periodic and random potentials.

III.2 Death of the branches

Branched flow is a transient regime that precedes homogeneous diffusion. Sufficiently far from their origin, branches fade away and eventually die. Adopting a semiclassical perspective and thinking in terms of phase space, we can understand this effect as follows. The initial manifold gets stretched and folded creating the cusps that we studied in the previous section. These cusps give birth to the branches, which continue to live as long as the dynamics do not smear them too much. For random potentials, this means that the number of visible branches grows exponentially, peaks, and then decays exponentially with time [10],

NbeΩt,proportional-tosubscript𝑁𝑏superscript𝑒Ω𝑡N_{b}\propto e^{-\Omega t},italic_N start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ∝ italic_e start_POSTSUPERSCRIPT - roman_Ω italic_t end_POSTSUPERSCRIPT , (16)

where ΩΩ\Omegaroman_Ω is also related to the stretching factor distribution of the transverse direction. We must recall that in the ray-tracing approximation, the number of caustics keeps increasing indefinitely, since the dynamics keep stretching and twisting the initial manifold and the fractal structure of cusps goes on at all scales. However, as the whirls and curls diminish in size, the caustics eventually merge with the background density, becoming indistinguishable from it. Furthermore, if we want to make the connection with the quantum mechanics counterpart, the phase space has some finite resolution given by Planck-constant-over-2-pi\hbarroman_ℏ, blurring any details finer than this scale. Therefore, after the initial rise, cusps and visible branches become uncorrelated. Although for long times the number of caustics is no longer useful, we can still count branches directly from the data using the manifold tracking algorithm explained in Sec. II.2.

In non-integrable periodic potentials, an exponential decay is also to be expected, given that the chaotic dynamics is similar to the random case. Although algebraic decay of branches is plausible for longer times, it appears unlikely for the relatively short time scales of branched flow, as discussed later. In any case, besides the chaotic part of the flow, we must consider superwires, which are indefinitely stable branches associated with KAM islands in phase space. In random potentials, these islands occupy a negligible portion of phase space and do not significantly affect trajectory flow. However, in periodic potentials, the KAM islands fill a considerable amount of phase space and therefore carry an appreciable fraction of the flow. This means that Eq. 16 must be modified to take superwires into account, so for periodic potentials we have

NbeΩt+C,proportional-tosubscript𝑁𝑏superscript𝑒Ω𝑡𝐶N_{b}\propto e^{-\Omega t}+C,italic_N start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ∝ italic_e start_POSTSUPERSCRIPT - roman_Ω italic_t end_POSTSUPERSCRIPT + italic_C , (17)

where the constant C𝐶Citalic_C is related to the fraction of phase space occupied by the superwires.

Refer to caption
Figure 4: Number of branches over time, measured by the manifold tracking method (see Sec. II.2). After the initial growth, the number of branches decays exponentially. For the periodic potential of Eq. (11), we observe that the number of branches does not go to zero, instead it reaches a horizontal asymptote. This phenomenon is attributed to the presence of superwires, which persist indefinitely. The plots have been averaged for 40 different potential orientations from 0 to π/4𝜋4\pi/4italic_π / 4 with the entire manifold considered as it expands.

We have carried out numerical experiments to verify the previous theoretical arguments. First, we have computed the evolution of the number of branches Nbsubscript𝑁𝑏N_{b}italic_N start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT as a function of the longitudinal direction x𝑥xitalic_x for the three potentials proposed in Sec. II.3. The results are depicted in Fig. 4. In the random potential, the flow tends to a uniform ray density, and the number of branches Nbsubscript𝑁𝑏N_{b}italic_N start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT drops to zero exponentially, as expected.

However, in the case of the periodic integrable potential, we observe that the number of branches stabilizes after a transient period. Due to the presence of superwires associated with stable motion, the number of branches does not tend to zero as time progresses to infinity; instead, it converges to a nonzero value C𝐶Citalic_C, as predicted by Eq. (17). The Fermi periodic potential exhibits mixed behavior, displaying both exponential decay and a constant floor NbCsubscript𝑁𝑏𝐶N_{b}\rightarrow Citalic_N start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT → italic_C after the transient. To better understand the interaction between chaotic and periodic dynamics in non-integrable periodic potentials, we focus on the Fermi potential in this last case.

Refer to caption
Figure 5: (a) Number of branches Nbsubscript𝑁𝑏N_{b}italic_N start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT over time for the Fermi periodic potential. The branches have been classified depending on the stability of the rays measured with the maximum Lyapunov exponent λ𝜆\lambdaitalic_λ. Only the part of the manifold within the bound of the original wavefront is taken into account, which explains the difference of scale of Fig. 4. (b) Fraction of phase space occupied by the chaotic and stable flow, classified according to the Lyapunov exponents.

In Fig. 5-(a), the number of branches in the periodic Fermi potential is separated according to the Lyapunov exponents λ𝜆\lambdaitalic_λ of the trajectories. The number of total branches is the sum of contributions from the stable (λ0𝜆0\lambda\approx 0italic_λ ≈ 0) and chaotic rays (λ>0𝜆0\lambda>0italic_λ > 0). The comparison with Fig. 4 reveals that the curve corresponding to the chaotic trajectories matches the behavior of the random potential, while the periodic part of the flow tends to a constant number of superwires.

In our preceding theoretical analysis, we referred to the fraction of phase space occupied by the branches as the appropriate interpretation of the constant C𝐶Citalic_C in Eq. (17). Due to the potential overlap of multiple branches in phase space, a histogram-based approach is necessary to quantify how the branches populate the transverse space. Specifically, we identify bins where branches appear within the histogram and calculate the ratio fareasubscript𝑓𝑎𝑟𝑒𝑎f_{area}italic_f start_POSTSUBSCRIPT italic_a italic_r italic_e italic_a end_POSTSUBSCRIPT as the number of bins with branches divided by the total number of bins examined over a fixed transverse interval. As depicted in Fig. 5-(b), we observe a clear exponential decay in the chaotic dynamics contribution, while the superwires quickly stabilize around a floor value C𝐶Citalic_C.

Some results in existing literature suggest that the decay of the branches could be algebraic instead of exponential, due to the stickiness of the KAM islands [14, 15]. While this may hold true for extended periods, our simulations have not yet demonstrated clear indications of such behavior. We conjecture that the branched flow regime occurs before stickiness significantly influences the dynamics.

IV Discussion

This work helps to clarify the relation between periodic potentials and branched flow. First, we shed some light into the definition of branched flow. In many papers, branched flow is considered to be a phenomenon restricted just to the so-called random potentials. Nevertheless, as far as for the growth of the number of branches is concerned, the behavior in random and non-integrable periodic potentials is exactly the same. This underscores the importance of the chaotic dynamics for the occurrence of branched flow, rather than the spatial disorder of the potential.

Nonetheless, branched flow in periodic potentials also introduces intriguing novelties, particularly concerning superwires. The potential for stable motion in periodic systems engenders fresh behaviors, exemplified by the manner in which branches diminish. While branches associated with chaotic dynamics exhibit behaviors akin to those in random potentials, superwires persistently concentrate flow, creating a plateau in the decay of the branches. Superwires possess significant inherent interest [16], and their application across various fields holds a promising potential. Nonetheless, it is conceivable to envision a periodic potential with minimal KAM islands in phase space, producing branch behaviors practically indistinguishable from those in random potentials, thus underscoring the pivotal role of chaotic dynamics in branched flow.

We hope our work will stimulate further research into branched flow in periodic potentials, an area ripe with theoretical and practical connections waiting to be explored.

Acknowledgments

This work has been supported by the Spanish State Research Agency (AEI), the European Regional Development Fund (ERDF, EU) under Project No. PID2019-105554GB-I00, and the Research Council of Finland under Project No. 349956 (ManyBody2D).

References

  • Heller et al. [2021] E. J. Heller, R. Fleischmann, and T. Kramer, Branched flow, Physics Today 74, 44 (2021).
  • Topinka et al. [2001] M. Topinka, B. J. LeRoy, R. Westervelt, S. Shaw, R. Fleischmann, E. Heller, K. Maranowski, and A. Gossard, Coherent branched flow in a two-dimensional electron gas, Nature 410, 183 (2001).
  • Patsyk et al. [2020] A. Patsyk, U. Sivan, M. Segev, and M. A. Bandres, Observation of branched flow of light, Nature 583, 60 (2020).
  • Degueldre et al. [2016] H. Degueldre, J. J. Metzger, T. Geisel, and R. Fleischmann, Random focusing of tsunami waves, Nature Physics 12, 259 (2016).
  • Mok and Fleischmann [2023] K. H. Mok and R. Fleischmann, Branched flows in active random walks and the formation of ant trail patterns, Physical Review Research 5, 043299 (2023).
  • Heller [2018] E. J. Heller, The Semiclassical Way to Dynamics and Spectroscopy (Princeton University Press, 2018).
  • Daza et al. [2021] A. Daza, E. J. Heller, A. M. Graf, and E. Räsänen, Propagation of waves in high brillouin zones: Chaotic branched flow and stable superwires, Proceedings of the National Academy of Sciences 118, e2110285118 (2021).
  • Cao et al. [2018] Y. Cao, V. Fatemi, S. Fang, K. Watanabe, T. Taniguchi, E. Kaxiras, and P. Jarillo-Herrero, Unconventional superconductivity in magic-angle graphene superlattices, Nature 556, 43 (2018).
  • Joannopoulos et al. [1997] J. D. Joannopoulos, P. R. Villeneuve, and S. Fan, Photonic crystals, Solid State Communications 102, 165 (1997).
  • Kaplan [2002] L. Kaplan, Statistics of branched flow in a weak correlated random potential, Physical Review Letters 89, 184103 (2002).
  • Metzger et al. [2010] J. J. Metzger, R. Fleischmann, and T. Geisel, Universal statistics of branched flows, Physical Review Letters 105, 020601 (2010).
  • Metzger [2010] J. J. Metzger, Branched flow and caustics in two-dimensional random potentials and magnetic fields, Ph.D. thesis, University of Göttingen (2010).
  • Klages et al. [2019] R. Klages, S. S. G. Gallegos, J. Solanpää, M. Sarvilahti, and E. Räsänen, Normal and anomalous diffusion in soft Lorentz gases, Physical Review Letters 122, 064102 (2019).
  • Lai et al. [1992] Y.-C. Lai, M. Ding, C. Grebogi, and R. Blümel, Algebraic decay and fluctuations of the decay exponent in hamiltonian systems, Physical Review A 46, 4661 (1992).
  • Motter and Lai [2001] A. E. Motter and Y.-C. Lai, Dissipative chaotic scattering, Physical Review E 65, 015205 (2001).
  • Graf et al. [2024] A. M. Graf, K. Lin, M. Kim, J. Keski-Rahkonen, A. Daza, and E. J. Heller, Chaos-assisted dynamical tunneling in flat band superwires, Entropy 26, 492 (2024).