Estimating Lyapunov exponents in billiards

George Datseris,∗ Lukas Hupe,† and Ragnar Fleischmann Max Planck Institute for Dynamics and Self-Organization and Faculty of Physics, Georg-August-Universit¨atG¨ottingen (Dated: July 18, 2019) Dynamical billiards are paradigmatic examples of chaotic Hamiltonian dynamical systems with widespread applications in physics. We study how well their , characterizing the chaotic dynamics, and its dependence on external parameters can be estimated from volume arguments, with emphasis on billiards with mixed regular and chaotic phase spaces. We show that in the very diverse billiards considered here the leading contribution to the Lyapunov exponent is inversely proportional to the chaotic phase space volume, and subsequently discuss the generality of this relationship. We also extend the well established formalism by Dellago, Posch, and Hoover to calculate the Lyapunov exponents of billiards to include external magnetic fields and provide a software implementation of it.

From the foundations of statistical physics to transport properties of electronic devices, in many areas of physics billiard models are an im- portant tool for understanding complex dynam- ics. In a billiard model a point particle is mov- ing freely (and frictionless) on a flat (or con- stantly curved) surface until it hits the bound- ary of the billiard where it is specularly reflected. Chaotic dynamics in the billiard are characterized by a positive Lyapunov exponent, measuring how initially close trajectories separate exponentially fast. Obtaining its value so far usually requires detailed numerical simulations of the chaotic dy- FIG. 1. (a) A regular (blue) and chaotic (red) in the mushroom billiard (MB), whose cap radius is a constant set namics. In our paper we assess how well the Lya- to 1 and the stem has width w and height h. The cap and punov exponent can be estimated from quite gen- the stem are separated with different background colors (or- eral considerations. Especially we study how pa- ange, green). (b) chaotic orbits without (orange) and with rameter changes that vary the phase space struc- (red) magnetic field and regular orbits (blue, purple) in the ture of the billiard get reflected in the Lyapunov magnetic periodic Sinai billiard (MPSB) with disk radius r. exponent. For example the application of an ex- ternal magnetic field can force some trajectories in the billiard on closed cyclotron orbits. We lasers [6,7] and room acoustics [8]. Billiard-models have show how the mere existence of such orbits varies also been particularly successful in helping to understand the Lyapunov exponent of the chaotic dynamics transport properties of electronic nanostructures such as through the phase space volume they occupy . quantum dots and antidot super-lattices [9–18]. The knowledge of this connection will be help- A billiard consists of a finite (or periodic) domain in ful to understand physical mechanisms in many which a point particle performs free flight with unit ve- systems like the magneto-transport in graphene locity. Upon collision with the boundary of the domain nanostructures. the particle (typically) is specularly reflected. In Fig.1 arXiv:1904.05108v2 [nlin.CD] 17 Jul 2019 we are showing the two example billiards we will be con- sidering in this paper: the mushroom billiard (MB) [19] I. INTRODUCTION and the periodic Sinai billiard [1] without (PSB) and with magnetic field (MPSB). Dynamical billiards are a well-studied class of dynam- An essential characterization of the chaotic dynamics ical systems, having applications in many different fields of a billiard is of course provided by its Lyapunov expo- of physics. Besides playing a prominent role in ergodic nents. (In this article we will study the Lyapunov expo- theory [1–3], billiards are important example systems for nents of the billiard flow in the physical, continuous time, understanding [4,5], with practical ap- in contrast to those of the billiard map in a discrete time plications e.g. in modelling optical microresonators for that counts the number of collisions with the boundary). For a two dimensional billiard the Lyapunov exponents are four numbers λ1−4 that measure how “chaotic” the ∗ [email protected] billiard is, in terms of the average exponential rates of ex- † GD and LH contributed equally to this work pansion (and contraction) of the phase space along cer- 2 tain characteristic directions. Due to the Hamiltonian sions with curved boundaries [34]. Assuming an aver- nature of the dynamics, the Lyapunov exponents fulfil age perturbation growth increase of C between collisions λ1 = −λ4, λ2 = λ3 = 0. Therefore in the remainder with curved boundaries and an average time κ between of the text we will be only considering the largest expo- such collisions, one would expect a perturbation growth t/κ nent λ ≡ λ1. The fundamental mathematical properties of |δΓ(t)| ≈ C |δΓ(0)|. This means for the Lyapunov of the Lyapunov exponents in billiards, including rigor- exponent of the ergodic component of phase space (see ous proofs of their existence, have been studied in litera- definition in Eq. (4) below) we expect ture and can e.g. be found in Ref. [20–22] and references therein. log(C) λ ≈ . (2) Quantitative studies of the Lyapunov exponent in ac- κ tual physical billiards are surprisingly rare however. A computational framework for calculating λ in billiard sys- In general, billiards have non-curved boundaries as tems was formulated by Dellago, Posch, and Hoover in well as curved ones. The mean collision time τ between Refs. [23, 24] (to which we refer to as DPH framework two consecutive collisions with any parts of the billiard in the following, and which we will extend to the dy- boundary is known analytically for any billiard and is namics in magnetic fields). Alternative approaches are given by presented e.g in Refs. [25, 26]. In the literature, Lya- π|Q| punov exponents have been computed for the PSB on τ = (3) square [27] and hexagonal [23, 27] lattices, as well as for |∂Q| for the stadium billiard [23, 28], which is related to the mushroom billiard. Furthermore, there are results for the where |Q| is the area of the billiard and |∂Q| the total magnetic elliptical billiard [29] and the inverse magnetic length of its boundary [20]. Since eq. (3) is averaged over stadium [30]. the entire boundary of the billiard it includes contribu- All these quantitative calculations rely on detailed nu- tions from both chaotic and regular components (if any). merical simulations of the complex billiard dynamics. Also notice that a formula similar to eq. (3) exists for In this paper, however, we want to follow a different billiards of any dimension, see [20]. approach exploring approximate expressions for the pa- The starting point our work is the observation that in rameter dependence of the Lyapunov exponents in some billiards the mean collision time between curved bound- paradigmatic cases, especially of billiards with mixed aries κ (a more precise definition will be given in sec- phase space, where regions of regular and chaotic dy- tionIIC) is fundamentally linked to the chaotic phase namics coexist. Our work is motivated by a recent study space volume VCH by Kac’s lemma [31–33]. Therefore that has shown that magnetoresistance measurements also the Lyapunov exponent is linked to the chaotic phase in graphene and semiconductor nanostructures directly space volume, and the aim of this paper is to explore reflect the parameter dependence of the chaotic phase how far considerations like Eq. (1) allow us to estimate space volume [9]. This is due to the fact that character- the parameter dependence of the Lyapunov exponent in istic transport times in the chaotic sea are fundamentally billiard systems. To this end we will analyze the contri- linked to the respective volume of the chaotic phase space butions to the approximate expression (2) and compare in the corresponding billiards. Namely, for the magnetic it with detailed numerical simulations. In sec.II we pro- periodic Sinai billiard (MPSB) in Fig.1b it was analyt- vide the basic framework we will use for computing λ, ically shown that the mean collision time κ(B) between as well as apply the aforementioned back-of-the-envelope successive collisions with the discs (of radius r) as a func- calculation to realistic perturbation growth. Following in tion of an applied external magnetic field B is equal to sec.III we present our results about the periodic Sinai the varying chaotic phase space portion gc(B) times the billiard and the mushroom billiard. We conclude by dis- value of κ at zero magnetic field [9], i.e. cussing the generality of our approach, while presenting one additional billiard with mixed phase space, the in- 1 − πr2 κ(B) = g (B) × κ(0) = g (B) × . (1) verse stadium billiard, which has been studied by V¨or¨os c c 2r et al. [30]. (The value of κ(0) is obtained from Eq. (3) below.) For the convenience of the reader we replicate the proof of eq. (1) from [9], which uses Kac’s lemma [31–33], in Ap- II. LYAPUNOV EXPONENTS IN BILLIARDS pendixD. The Lyapunov exponent in billiards is also linked In this section we first give a brief overview of the DPH to mean collision times as the following back-of-the- framework [23, 24] for numerically computing λ, reciting envelope calculation motivates. Let us study the pertur- the equations that will be relevant for our study. We will bation growth, i.e. the exponential growth of the phase then extend the framework to motion in an external mag- space distance |δΓ(t)| of two initially infinitesimally close netic field. In the following we will assume that the point by trajectories. The origin of the exponential pertur- particle in the billiard has unit mass and momentum and bation growth and thus of chaos in billiards are colli- velocity are the same. 3

The (maximum) Lyapunov exponent is defined based the perturbation vector δΓ(t) in the presence of a per- on the evolution of the four-dimensional perturbation pendicular magnetic field is given by vector δΓ = (δq, δp)T as δΓ(t) = B · δΓ(t0), (7) 1 |δΓ(t)|   λΓ(0),δΓ(0) = lim log (4) ρ sin(ωt) ρ (cos(ωt) − 1) t→∞ t |δΓ(0)|  I2×2   −ρ (cos(ωt) − 1) ρ sin(ωt)  B =   with δΓ evolving according to the evolution equations in  cos(ωt) − sin(ωt)  02×2 tangent space, δΓ˙ = J(Γ(t))·δΓ where J is the Jacobian sin(ωt) cos(ωt) matrix of the . For all Γ(0) inside an ergodic component of phase space the value of λ does not with the cyclotron frequency ω = 2B and the cyclotron depend on the initial condition. radius ρ = 1/ω. As already mentioned in the introduc- tion, in the billiard the particle always moves with unit velocity by convention. The expressions that give the A. Without magnetic field discontinuous change of the perturbation vector at a col- lision with a wall or disc are The time evolution in tangent space for a particle mov- ! δq − 2 (δq · n) n ing in a straight line is 0 δΓ = hδq,ei 0 − δp − 2 (δp · n) n − 2γr cos φ e  t ·  ! δΓ(t) = I2×2 I2×2 δΓ(0). (5) (δq · n) 0 02×2 I2×2 ω (8) (p · n) S · p At discrete time points Eq. (5) is interrupted by collisions 2 2! ! −2n1n2 n − n n1 with the boundary and the perturbation vector changes S =2 1 2 , n = n2 − n2 2n n n discontinuously. The perturbation vector just after the 1 2 1 2 2 collision (we’ll use 0 to label quantities right after the collision) is derived from the one just before the collision where γr again is the of the wall segment (0 for a straight wall, ± 1 for a disc). as [24] r ! δq − 2 (δq · n) n 0 C. The “toy model” δΓ = (δq·e) 0 (6) δp − 2 (δp · n) n − 2γr cos φ e Before finding an approximate expression for the value for a collision with a boundary segment of curvature γr. of λ in our model systems, it is worthwhile to get an The two types of boundaries we will encounter in this impression of how the norm of the perturbation vector work are straight walls and the boundaries of circular evolves with time. In Fig.2 we show typical plots for discs. For a straight wall section γr = 0 and for a disc the three different billiards. We computed the pertur- of radius r we have γ = ± 1 , with − for collisions hap- r r bation growth using the DPH framework, sampling the pening from the inside of the disc (as in the MB) and perturbation vector immediately before and after every + otherwise (as in the PSB). Here n denotes the normal collision to resolve the instantaneous jumps. As the DPH vector of the boundary segment at the collision point q, evolution is linear, in the actual numerical simulations and φ is the angle of incidence (measured with respect 0 we renormalized the perturbation vector after sampling to the normal to the surface). The vectors e, e are unit to prevent numerical error due to the rapid perturbation vectors orthogonal to the incident and reflected momenta 0 growth. p and p respectively (for more details see [24]). Let us first examine Fig.2(a, b). We see that the norm Notice that a collision with a straight wall does not of the perturbation vector changes in two ways. Let the change the norm of δΓ because both the coordinate and j−1 j-th collision with a disc happen at time t = P ∆t = the velocity components are reflected specularly. j i=0 i tj−1 + ∆tj−1. There a discontinuous change of the per- 0 turbation vector norm happens, so that |δΓj| = aj|δΓj| (in general aj is a function of δΓj). The collision event is B. With magnetic field followed by a time-interval ∆tj, in which the perturba- tion norm changes continuously because there are no col- We now extend the DPH formalism for a particle ex- lisions with curved boundaries. Just before the next col- periencing a magnetic field perpendicular to the billiard lision with a disc the perturbation norm takes the value plane. In this section we present only the final expres- |δΓj+1|. In the following we will refer to these repeated sions. The full calculations are presented in appendixA. segments of the growth curve as elementary growth seg- The magnetic field is uniform, with value B (positive ments, starting with one collision event with a disc and means counter-clockwise rotation). The free evolution of ending just before the next one. In general it is the 4

FIG. 2. Typical time evolution of the of the norm of the perturbation vector |δΓ(t)| for the periodic Sinai billiard without (a) and with (c) magnetic field (B = 1) and for the mushroom billiard (e) (without magnetic field). Zoom-ins are below each panel. For (a-d) red markers mean collision with the disc while blue markers mean “collision” with the periodic walls (not a true collision, but a way of recording the value of |δΓ|). For (e-f) red is collision with cap head, orange with cap walls, blue with stem sides and green with stem bottom. The coloured background stripes denote the elementary growth segments discussed in sec.IIC (random colors are used) with hatched orange color used for the laminar episodes (see sec.IIIC). segment of the perturbation growth curve between suc- We then recursively apply Eq. (10) to get cessive dispersing or defocussing collisions which are the n−1 origin of chaos in billiards [34]. An elementary growth Y |δΓ | = a z(∆t )|δΓ(0)| ⇒ segment thus reflects the perturbation growth during an n i i “effective free path” as defined by Bunimovich [34]. i=0 n−1 X A crucial simplification we do in deriving an approx- log (|δΓn|) = log(ai) + log(z(∆ti)) (11) imate expression for the Lyapunov exponent will be to i=0 express the perturbation growth in the time-interval ∆tj (using |δΓ(0)| = 1) and with T = Pn−1 ∆t we use as a function z(∆tj) of the interval length. The actual n i=0 i 0 Eq. (4) to write λ = limn→∞ log (|δΓn|) /Tn. The quo- precise value |δΓj+1/δΓj| is not a simple scalar function of the time interval since for example in the case of the tient of the infinite sums is the same as the ratio of the PSB we can obtain from Eq. (5) average over all unit cells (denoted by h·i), i.e. 1 λ ≈ (hlog(a)i + hlog(z(∆t))i) (12) q κ |δΓ (∆t)| = (δq0 + δp0 ∆t)2 + (δp0 )2 (9) j+1 j j j κ ≡ h∆ti . (13) Averaging over the unit cells implicitly assumes the er- (due to the linearity of the equations of motion of the godicity. Notice also that in some billiards there could 0 tangent space we can assume a norm of |δΓj| = 1 in be several ergodic chaotic components that are sepa- Eq. (9)). Eq. (9) depends on the initial orientations of rated from each other. In such cases the above process both the momentum and position deviation vectors and has to be applied to each component separately, since thus is not a function of just ∆t. each component has its own exponent λ. For the bil- We will show, however, by analysing numerical data, liards considered here we have found that furthermore that a reasonable approximation can be obtained by as- hlog(z(∆t))i = log(z(h∆ti)) = log(z(κ)) is a good ap- suming that such a function z(∆t) exists. Notice that proximation that we will use. This approximation is valid this assumption regards only the existence of z. The when the standard deviation of ∆t is small (compared to functional form and its can be completely ar- its mean). bitrary (and in fact in the following we have three dif- In the remainder of the text we will call Eq. (12) the ferent versions of z for the different billiards). For each “toy model”. It is the more detailed version of the back- elementary growth segment we thus write of-the-envelope calculation given in the introduction. In the following sections we will apply this toy model to specific billiards and see how well it approximates the 0 |δΓj+1| = z(∆tj) × aj × |δΓj| = z(∆tj) × |δΓj|. (10) Lyapunov exponent and its parameter dependence. 5

D. Software

All numerical computations presented in this paper were performed with an open source software to simulate billiards, DynamicalBilliards.jl [35]. In this paper we extend the DPH framework to magnetic fields. We also implemented this extension in the software (which pre- viously only included the non-magnetic case). All code we used for this paper, including all code to replicate the figures we show here, is publicly available on GitHub: https://github.com/Datseris/arXiv_1904.05108.

III. RESULTS

A. Periodic Sinai billiard

We start our analysis with the PSB because it is the simplest case and we are able to give a fully analytical expression for the Lyapunov exponent in the simplified toy model. We note that in the absence of a magnetic field the PSB is ergodic and its phase space is not mixed. Nevertheless, it will serve as a pedagogic example of how FIG. 3. (a) Lyapunov exponent of the periodic Sinai billiard the toy model approximates the Lyapunov exponent. for different radii, compared with the toy model. The dashed For the PSB τ = κ and the value of τ is known from curve obtains hlog(a)i by numeric average while the red curve uses Eq. (19). The blue curve is using the DPH framework. Eq. (3) (b) Average value of hlog(a)i used in panel (a). (c) Perturba- 1 − πr2 tion norm increase during the√ free flight part. In the zoom in κ = . (14) (d) we also plot the curve 1 + ∆t2 as a dashed line. PSB 2r An approximation for z(∆t) is easily found as well from Eq. (9), namely We now need to average hlog(a)i. We start our ap- proximation by replacing the inner product (δq·e) by an p 2 averaged quantity b(r) (we show below how b depends on zPSB(t) ≈ 1 + t (15) r). It is expected that perturbations will grow and ori- which uses the assumption that after a collision with a ent themselves perpendicular to the particle’s direction disc the momentum contribution to the perturbation vec- of motion. Since e is a unit vector perpendicular to the tor is much larger than the position contribution, i.e. particle’s momentum and thus parallel to δq, this means |δp0|  |δq0|. Numerical calculations show that this ap- that h | (δq · e) | / | δΓ | i = h | δq | / | δΓ | i = b(r). proximation is valid for small enough radii (see Fig.3). Which portion of |δΓ| is contained in |δq| is answered The instantaneous change factor hlog(a)i is rather large based on how perturbations evolve during the free flight for all but very large disc radii. Also as seen in panel part. Starting from the j-th collision the perturbations (c) and its inset (d), Eq. (15) almost perfectly approx- evolve for time ∆tj. Using Eq. (9) (and assuming that imates the perturbation norm increase during the free the cross-terms δp · δq will drop out in the averaging) we flight part. obtain

We still need an approximate expression for a, the r 0 2 |δqj | 2 instantaneous change factor, which we can derive from |δp0 |2 + ∆tj |δq | j Eq. (6). Since the norm of the position deviation does not j+1 = . (17) |δΓ | r 0 2 change at the collision, we focus on the momentum devia- j+1 |δqj | 2 δq·e 0 2 + 1 + ∆tj 0 (r) 2 0 (r) |δpj | tion δp = δp − r cos(φ) e with δp = δp−2 (δp · n) n (if not explicitly written otherwise all quantities in this To analytically resolve eq. (17) we use the same as- paragraph have a collision-time index j, which we sup- sumption as in eq. (15), |δp0|  |δq0|. We then average, press to make the symbols simpler). We define A = replacing ∆t by κ, which leads to 2 δq·e r cos(φ) and carrying out the calculations leads to s 2 κPSB 0 p 2 0 b(r) = 2 . (18) |δΓ |/|δΓ| = a = 1 + A − 2A(e · δp) (16) 1 + κPSB 0 again using the assumption that |δΓj| = 1 (and recall We discard the term 2A(e · δp) in Eq. (16) again that |δp(r)| = |δp|, |δq0| = |δq|). assuming that the inner product averages to 0. Now 6 the only variable left to average over is φ, the angle This more complex behaviour is of course hidden in the with respect to the normal vector. This is distributed more complicated formulas of our extension to the DPH in [−π/2, π/2] with probability distribution of cos(φ)/2. framework for magnetic fields. For example, explicitly Therefore writing out Eq. (7) gives   π s 2 h 2 2 Z 2   2b cos(φ) |δΓ(t)| = (δpx + δpy)+ hlog(a)i = log  1 +  dφ − π r cos(φ) 2 2 2  cos(ωt) − 1 sin(ωt) √ δqx + δpy + δpx + −1 2b  2 2   ω ω csch r 4b + r b = + log (19) #1/2 r r  cos(ωt) − 1 sin(ωt)2 δq − δp + δp y x ω y ω with csch−1 the inverse hyperbolic cosecant. We then put eqs. (14), (15), (18) and (19) into the toy (21) model of Eq. (12) and obtain an analytic approximation (where again we assumed |δΓ(0)| = 1). The consequences for the Lyapunov exponent of Eq. (21) can be seen in Fig.4(d, e). Using a uni-variate    scalar function z(∆t) to approximate these distributions csch−1 2b(r) p4b(r)2 + r2 2r r appears to be a bold move, but in the end it will turn λ (r) = + PSB 1 − πr2  r out to give a good approximation. To obtain z(∆t) we simplify Eq. (21) to s 2 b(r) 1 − πr2  s log + log  1 +  . 1 − cos(ωt)2 sin(ωt)2 r 2r z (t) = 1 + + . (22) MPSB ω ω (20) which is also plotted in Fig.4(d, e). The result is shown in Fig.3(a), compared with the nu- In the next step we compute hlog(a)i numerically and merical value of λ using the DPH framework as well as use its value in the toy model along with zMPSB(κ(B)). with the result of computing the term hlog(a)i in the We remind that the value of κ, the mean collision time be- toy model numerically from the evolution of the pertur- tween discs in MPSB, is not known analytically but it is bation vector norm. All three curves are in excellent connected with the chaotic phase space portion through agreement for small and intermediate r, only for large r Eq. (1). The results of the toy model are presented in does Eq. (20) slightly deviate from the numerical values Fig.4. because the approximation for b(r) in Eq. (18) and thus Besides the fact that our toy model approximates λ for hlog(a)i becomes less accurate. very accurately, Fig.4 nicely shows the impact of phase space restriction on λ. In our toy model the value of λ is composed of five contributions, the first being the B. Magnetic periodic Sinai billiard denominator κ. The value of hlog(a)i itself has two con- tributions, one again stemming from κ (as shown in the We now want to apply the same process to the MPSB, previous section) and the other from B. The function which, however, has a mixed phase space: there exist zMPSB also has two contributions, one from B and one collision-less orbits like those seen in Fig.1(b) that con- from κ. Therefore three out of five contributions to λ are stitute the regular part of phase space (other unstable inherently linked to the restriction of the chaotic phase periodic orbits of zero measure are not relevant here). We space by regular orbits. are of course only considering the Lyapunov exponent of the chaotic part of the phase space, which means that we initialise particles only in the chaotic phase space region. C. Mushroom billiard The mean collision time κMPSB between successive colli- sions with discs is also only defined for the chaotic phase Because the volume fractions of the regular and the space part (as the regular trajectories do not collide with chaotic phase space regions are not known analytically the discs). in the MPSB we have turned to a billiard that also has a The free flight evolution in the MPSB is fundamen- mixed phase space but allows to calculate these fractions, tally different from the PSB. Not only are the functional and, as we will show, the relevant average time scales forms different but in addition due to magnetic focusing analytically: the mushroom billiard (MB). The regular it is possible (and in fact quite frequent) for the pertur- orbits in the MB are orbits forever staying in the cap, bation norm to decrease during the evolution, as can be evolving exactly like they would as if they were in a cir- seen in Fig.4(e,f). In addition, as visible in Fig.2(d), cular billiard [19]. The tangential circle to these orbits it is also possible for the norm to decrease during the has a radius ≥ w/2 as shown in Fig.1(a). The rest of instantaneous change as well. the orbits, which do not satisfy this criterion, eventually 7

FIG. 5. (a) A laminar episode (orange, dashed) and two chaotic episodes (blue, purple) in the mushroom billiard. Starts and ends of each episode are denoted with closed and FIG. 4. (a) Lyapunov exponent of the magnetic periodic Sinai open circles and the blue episode starts directly after the or- billiard (MPSB) for different radii versus the magnetic field, ange. The cap head is plotted in dark color to differentiate. compared with the toy model. The solid curves are the nu- (b) Mean return time to the stem bottom, which is equivalent meric result λ, the dashed curves are the toy model using the with the average elementary growth segment time, eqs. (27), numeric average of hlog(a)i. (b) Numeric average of hlog(a)i (28). (c,d) Volume of chaotic phase space for the mushroom versus the magnetic field. (c) Chaotic phase space portion billiard. (gc = κ/κ(0)) of the MPSB ((a-c) share legend). (d, e) Per- turbation norm change during the free flight in the MPSB (calculated with the DPH framework). Plotted with dashed butions to the increase, each seemingly approximated lines are Eq. (22) (data for r = 0.2). as linear function of ∆t. By analysing the dynamics in more detail, it turns out that the different contribu- tions of Fig.6(c,d) come from the trapping of the chaotic enter the stem and are chaotic. The tangential circle ar- orbits in the regular phase space. In coordinate space gument was used in [5] to obtain an analytic expression this means that the particles get trapped in the cap and for the regular phase space volume VREG of the MB as a mimic the motion of the regular phase space there until function of the billiard parameters eventually escaping after some time. This effect is often called intermittency, and is known to occur in mushroom  w  w  w2  VREG = 2π arccos − 1 − (23) billiards [36, 37]. Intermittent behaviour in the MB can 2 2 4 happen in the stem as well, where orbits stay trapped VTOT = 2π(hw + π/2), (24) bouncing between the stem walls.

VCH = VTOT − VREG (25) We therefore have to separate the elementary growth segment into two different “episodes”: the chaotic where VTOT,VCH is the total and chaotic phase space episode c and the laminar episode `, where the particle is volume (all lengths are scaled to the cap radius r which trapped in the cap. Notice that the second intermittent is fixed to r = 1). The parameter dependence of VCH is behaviour, trapping in the stem, does not lead to a new illustraded in Fig.5 c and d. Interestingly, VCH does not type of dynamics, but is just prolonged chaotic episodes vanish for small h, although it is obvious that there are (similarly to a large free flight in the PSB). We show the no chaotic orbits for h = 0. This discontinuity is due to two episodes in Fig.5(a). Numeric calculations shown in the fact that the volume of chaotic phase space in the cap Fig.6(e, f) show that each episode has a different average is independent of stem height for nonzero h, but drops time, τc, τ` respectively. to zero for h = 0. During the chaotic episodes the picture is very simi- In Fig.6(c,d) we present a scatter plot of various pos- lar to the PSB. A collision with the cap head gives an sible increases of the perturbation norm during the unit instantaneous increase to |δΓ|, followed by an approx- cells. We found that there are clearly distinct contri- imately linear increase until the next collision with the 8

FIG. 6. (a, b) Lyapunov exponents in the mushroom billiard (MB) versus the width w or height h of the stem. Solid lines are numeric results using the DPH framework, dashed lines are using the toy model. (c, d) Perturbation norm increase during the chaotic c and laminar ` episodes. (e, f) Parameters of the toy model versus w or h (for constant h = 1 and w = 1 respectively); legend is shared. cap head. Here the linear increase approximation is valid The instantaneous change factor a is the same be- because for the chaotic episodes ∆t & 2h + 2 − w. Af- tween the laminar and chaotic episodes so we do not ter colliding with the cap head the particle may return need to separate it. Notice that for the laminar episode to the stem immediately which initialises another chaotic we only consider the first jump as the instantaneous in- episode. Occasionally, after ending a chaotic episode, the crease. Subsequent jumps that decrease rapidly are en- particle will get trapped in the cap (see Fig.5(a) orange), coded in the linear growth approximation. After comput- starting a laminar episode. Even though there are suc- ing hlog(a)i numerically, we still need a value for κMB, cessive collisions with the cap head in this episode, the the elementary growth segment average time, to apply perturbations do not increase exponentially. The succes- our toy model λ = (hlog(a)i + log(zMB(τc, τ`))/κMB. Nu- sive instantaneous increases are very quickly becoming merically we can estimate insignificant (see Fig.2(e,f)) due to the fact that cap collisions have an initially focusing effect which only be- κMB = τc + f`τ`. (27) comes de-focusing if the consecutive free motion is long enough, which is not the case in the laminar episodes. However we can estimate κMB analytically as well, using Therefore the overall perturbation growth inside the cap Kac’s lemma [31–33]. The key to this is understanding trapping episodes is linear. that the mean elementary growth segment time is equiva- Let nc and n` be the counts of chaotic and laminar lent with the mean return time to the stem bottom, since episodes up to time T . Notice that n` is strictly less all phases in the end of the day have to go there, since than nc since a chaotic episode always follows a laminar all phases are part of the chaotic phase space. episode, but the inverse is only occasionally true. In the We present the full proof in appendixB. The final ex- limit T → ∞ we define f` = n`/(n` + nc) to be the pression is given by frequency of the laminar episodes. We then write the V (h, w) function z as κ (h, w) = CH (28) MB 2w zMB(∆t) = (oc + sc∆tc) + f`(o` + s`∆t`) (26) We compare the analytic formula with the numeric result with oi the offset and si the slope of the linear approx- in Fig.5(b) and find the expected perfect agreement, imation (we obtain these values with least squares fit since Eq. (28) is exact. In appendixC we also present an to Fig.6(c, d)). For the chaotic episodes s, o are con- analytic approximation for τc. Since we know κMB and stant versus h, w while for the laminar episodes o depends τc analytically, we also know the product f`τ` (but we strongly on w. Also, for the chaotic episodes o has nega- don’t have an expression for f` or τ` individually). tive value (of around -0.8) which is expected due to the We can now use our toy model to compare with λ, focusing effect. We once again compute hlog(z(∆t))i sim- which we do in Fig.6(a, b). Again we find good agree- ply by replacing ∆t by its average values τc, τ` in Eq. (26). ment between toy model and the numerical simulation 9 using the DPH framework. The model mildly diverges for very small w, probably because the mean laminar time τ` diverges as seen in Fig.6(e). As was the case in the MPSB, the average elemen- tary growth segment time κ is inversely proportional to λ and directly proportional to the chaotic phase space volume VCH. This shows that phase space restrictions have an immediate impact in the value of the Lyapunov exponent even for billiards with intermittent dynamics. Notice that in the MB both VCH and λ increase as w in- creases. This is simply due to the dependence of VCH on w, as well as the direct dependence of κMB on 1/w (this for example was not the case in Eq. (1) for the MPSB).

IV. DISCUSSION

To summarise, we have examined the value of the Lya- punov exponent λ in chaotic billiards. We were able to create a conceptually simple model that approximates λ very well. The model is based on how perturbations evolve in billiards on average and helps to understand how each part of the dynamics contributes to the pertur- FIG. 7. The inverse stadium billiard (ISB). (a) An example bation increase. The simple model is written as eq. (12), orbit in the ISB (stadium width and length are 0.5). Outside which is the stadium the particle undergoes circular motion with ra- 1 dius 1/ω. (b) Boundary map (see AppendixB) of the ISB, λ = (hlog (a)i + hlog (z(∆t))i) κ computed for ω = 10. In the middle one can see stability islands, which seem to have a boundary. (c) Lyapunov where a the instantaneous change of |δΓ| at a collision exponent λ and chaotic volume VtextCH versus ω, both nor- with a curved boundary and z(t) the continuous change malized to their maximums for comparison. λ is obtained of |δΓ| in between collisions with curved boundaries. κ with a modified version of the DPH framework using tangent is the average elementary growth segment time equal to space evolution matrices derived by V¨or¨oset al. in [30]. VCH the mean collision time between curved boundaries. The is the volume of the billiard flow of the chaotic orbits and is approximations that lead to the toy model were the fol- calculated by weighting the area of the ergodic region (ob- lowing. First we assumed that the chaotic phase space tained numerically) on the boundary map with its mean time is ergodic and time averages can be replaced by phase to next collision. We only consider orbits that do intersect the billiard boundary of the ISB. space averages and that for ∆t, log(a), log(z(∆t)), their averages are finite and greater than 0. We then made the simplifying assumption that the norm of the perturba- tion vector increases continuously in between successive in billiard is given by the inverse chaotic phase space chaos-inducing collisions (i.e. in each elementary growth volume, we present one final billiard, called the inverse segment) as a univariate function of the time interval stadium billiard (ISB) shown in Fig.7(a), originally stud- z(∆t). ied by V¨or¨oset al. in Ref. [30]. In this billiard a particle We used Eq. (12) to find an analytic expression for λ is propagating inside the stadium on straight lines, but in the periodic Sinai billiard (PSB). We have also shown after passing the boundary of the stadium it is subjected that Eq. (12) can be used to analyse the Lyapunov ex- to a constant magnetic field, which brings the particle ponent in the magnetic periodic Sinai billiard (MPSB), back inside the stadium, as depicted in Fig.7(a). In the and by approximating the numerical curves identified the limit of infinite magnetic field the ISB recovers the fully main contributions. We could follow the same approach chaotic Bunimovich stadium, for finite magnetic fields it for the mushroom billiard (MB), even though the process has a mixed phase space. Here, we don’t want to analyze is complicated in this case by intermittent dynamics. In the ISB in any detail but only point out that also in this both billiards with mixed phase space we connected the billiard the parameter dependence of its Lyapunov expo- chaotic phase space volume with λ through κ and showed nent is closely following the inverse of the chaotic phase that λ has a leading contribution given by the inverse of space volume as shown in Fig.7(b,c). the chaotic phase space volume (for the MPSB we used We want stress how different the mechanisms are that the chaotic portion instead of volume, because the total lead to chaos in the three different billiards. In the MPSB phase space volume does not depend on B). it is dispersing as well as the magnetic field. In MB it is To strengthen our point that a prominent contribution defocusing and in the ISB it is even more involved. Yet in to the parameter dependence of the Lyapunov exponent all three cases we find what is suggested by our toy model: 10 the Lyapunov exponent has a leading contribution that Here n is the normal vector of the obstacle at the collision is inverse to the chaotic phase space volume. point, I2×2 and 02×2 are the 2×2 identity and zero matri- Because generally chaos in billiards arises via dispers- ces respectively, and (a ⊗ b)jk = ajbk is a second-order ing and defocusing collisions with curved boundary seg- tensor. ments [34], the Lyapunov exponent is necessarily in- versely linked with the mean return time to these bound- ary segments. Furthermore, Kac’s lemma dictates that 1. Propagation the mean return time of the chaotic trajectories to these boundaries is directly proportional to the chaotic phase In magnetic billiards, particles propagate in circular space volume. For this reason we hypothesise that for arcs. To get the simplest possible set of equations of most chaotic billiards with mixed phase space the Lya- motion describing this mode of propagation, it is useful to punov exponent has a leading contribution inverse to the introduce a phase angle θ = arctan(py/px). For uniform chaotic phase space volume. circular motion, this phase angle grows linearly in time This should straight-forwardly carry over to higher di- as the system rotates with constant angular velocity ω. mensions as well, since Kac’s lemma, the DPH frame- This can be exploited to determine p˙ using the chain rule work, as well as our toy model, do not depend in any dp dθ dp dp way on the dimensionality of the billiard. So far we didn’t p˙ = = · = ω · . (A5) find significant differences between billiards with sharply dt dt dθ dθ divided phase space (like the MB and the MPSB) and a fractal phase space structure (like the ISB). To con- Using kpk = 1, one can easily calculate the explicit rela- clude whether there are fundamental differences between tion between p and θ sharply-divided and fractal phase spaces one will have to dpx do more research. What we want to point out is that = −py px = cos(θ) dθ for fractal phase spaces it is much harder to estimate the ⇒ (A6) p = sin(θ) dp volume of the chaotic set. y y = p dθ x and combine the results of eqs. (A5) and (A6) to receive Appendix A: Evolution of perturbation vector in a the equations of motion magnetic field

! In their paper, Dellago, Posch and Hoover give two p F(Γ) = where R = 0 −1 . (A7) main equations to compute the evolution of perturba- ω · R p 1 0 tions in tangent space along a piece-wise smooth flow defined by the autonomous ODE system Using Eq. (A2), we can now state the equations of evo- lution for a perturbation vector δΓ using the Jacobian J Γ˙ = F(Γ). (A1) of F. During smooth propagation, the perturbation vector δΓ ! 0 along the trajectory Γ evolves according to δ˙Γ = J · δΓ = 2×2 I2×2 δΓ. (A8) 02×2 −ω · R

∂F δ˙Γ = · δΓ. (A2) ∂Γ Using an exponential ansatz, we can compute the general Γ solution of this system and receive a final result of If smooth propagation is interrupted at discrete times δΓ(t) = · δΓ(t ), (A9) tj(Γ) by a discontinuous jump, represented here by a B 0   differentiable map Mj(Γ), the perturbation vector after ρ sin(ωt) ρ (cos(ωt) − 1) I2×2 the jump is given by  −ρ (cos(ωt) − 1) ρ sin(ωt)  =   B   ∂Mj  cos(ωt) − sin(ωt)  0 02×2 δΓ = · δΓi + sin(ωt) cos(ωt) ∂Γ Γ   (A3) ∂Mj  · F(Γi) − F M(Γ) δτc where ρ = 1/ω is the cyclotron radius. ∂Γ Γ where δτc = tj(Γ + δΓ) − tj(Γ). For the case of elastic 2. Collisions reflection with an obstacle, Mj can be written as ! The derivation of the collision map for δΓ is largely 0 I2×2 02×2 analogous to the process used by DPH to derive their re- Γ = Mj(Γ) = · Γ. (A4) 02×2 I2×2 − 2(n ⊗ n) sult for non-magnetic billiards. The equations of motion 11 are assumed as stated in Eq. (A7). For equation (A3), we require the Jacobian matrix of Mj, which is ! ∂M 0 j = I2×2 (A10) ∂Γ AB

∂n where A = 2 ((n ⊗ p) + hp, ni ) I2×2 ∂q B = I2×2 − 2 (n ⊗ n).

By inserting eqs. (A7), (A4) and (A10) into (A3), we FIG. 8. (a) Geometric derivation of δτc. As all spatial pertur- find bations are small, it is sufficient to approximate the obstacle ! " ! as a straight line. In this example, the satellite particle col- 0 0 lides after the reference particle. (b) Decomposition of mo- δΓ0 = I2×2 δΓ + I2×2 · AB AB mentum into normal and tangential components ! !# (A11) p p − 2 (n ⊗ n) p − δτc. ω · R p ω · R B p It can be computed by determining the signed distance from the satellite to its collision point measured along It is now helpful to continue calculations for δq0 and δp0 its trajectory at the time tj(Γ) of the collision of the separately. For the position component δq0, equation reference particle. As we are considering the linearized (A11) can be written as dynamics of the perturbation and Eq. (A3) is valid only to the first order of δΓ, we will ignore all higher orders 0 δq = δq + [p − p + 2 (n ⊗ n) p] δτc of δΓ in the subsequent calculations.

= δq + 2 δτc (n ⊗ n) p. (A12) Furthermore, we will denote vector components in normal direction of the obstacle by a subscript n, i.e. 0 For the momentum component δp , we get an = ha, ni. Similarly, the tangent component of a vec- tor will be denoted by a subscript t. 0 δp = A δq + B δp + δτc [pA + ωS] (A13) Geometrically, one can immediately derive the follow- 2 2! ing two relations from the two triangles highlighted in −2n1n2 n1 − n2 where S := BR − RB = 2 2 2 . Fig.8a n1 − n2 2n1n2 ρ sin φ = hδq, ni − hc, ni (blue triangle) (A16) This can be simplified by using the fact that (b ⊗ a)c = ρ sin θ = hc, ni (red triangle) (A17) hc, ai b and introducing the quantity δqc = δq + δτc p, which represents the real space difference vector between the collision points of satellite and reference trajectories. where c is a vector between the obstacle and the cyclotron centre, as shown in fig.8a. Eliminating the factor hc, ni ∂n yields δp0 =δp − 2 hδp, ni n − 2 (hp, δq i n+ (A14) ∂q c hδq, ni hp, ni δqc) + δτc ω Sp. sin θ = − sin φ. (A18) ρ Using geometric considerations outlined by DPH we can π now rewrite the penultimate term to get By construction, θ cannot exceed 2 in absolute value. Therefore, we can safely apply the arcsine function to hδq, ei eq. (A18), receiving an expression for θ. δp0 = δp − 2 hδp, ni n − 2γ e0 + δτ ω p (A15) R cos φ c S hδq, ni  θ = arcsin − sin φ . (A19) where φ is the angle of incidence, γR is the local curvature r of the obstacle and e and e0 are unit vectors orthogonal 0 to p and p respectively. We can now use this to compute the angle α correspond- ing to the arc the particle has to travel during δτc. Using Fig.8a, we can determine an expression for α 3. Collision delay time α = φ + θ The quantity δτc in eqs. (A3) and (A11) can be inter- Eq. (A19) hδq, ni  preted as the time delay in between the collisions of the = φ + arcsin − sin φ . (A20) reference trajectory Γ and its satellite Γ + δΓ. r 12

This can be further simplified by expressing φ in terms billiard. Using Kac’s lemma [31–33], which states that for of the satellite particle’s momentum (compare fig.8b). volume-preserving maps, the mean number of iterations n required to return to a compact subset S of phase  p + δp  S α = arcsin t t + (A21) space is given by kp + δpk | {z } µ(A) K nS = (B1) hδq, ni p + δp  µ(S) arcsin − t t . r kp + δpk where µ(·) is the volume of a set and A is the subset of | {z } L phase space accessible to orbits originating in S. To transform the billiard flow into a map, we discretize We can now linearize Eq. (A21) to receive our final re- time in small steps ∆t, implicitly considering the limit of sult for α. As the individual terms are somewhat com- ∆t → 0. We now choose the set S of momenta and plicated, but very similar, we will treat them separately. positions defined by The first-order Taylor expansion of leftmost term is   S = {q ∈ Y, p > 0} (B2) pt y K ≈ arcsin + ζ(δpt) (A22) kpk where Y is a box of width w and height  at the bot- where ζ is of the form a·δpt +b·δpn with a, b given by the tom of the stem. One should be careful about the choice respective partial derivatives. Expanding the rightmost of S. The simplistic approach of choosing the cap semi- term in (A21) yields the similar result of circle as the returning set (since the elementary growth segments are delimited by collisions with curved bound-   pt 1 hδqi, ni aries) will not yield the correct result. That is because L ≈ arcsin − − ζ(δpt) + · . kpk p 2 r 1 − pt the mean return time to the cap semicircle inherently (A23) includes contributions from both the periodic orbits of This simplifies the final result for α significantly. Using the MB as well as the laminar episodes of the elemen- the anti-symmetry of the arcsine function, we can see tary growth segments, which we have already shown to that most of (A22) and (A23) cancel out, leaving only correspond to “free flight”-like motion. The phase space volume of S is µ(S) = πw in the 1 hδq, ni α = · . (A24) limit of ∆t → 0. As the chaotic phase space component p 2 r 1 − pt of mushroom billiards is ergodic [19], we know that the measure of the subset A of phase space accessible from Finally, we have to multiply equation (A24) with the cy- S is given by the volume of chaotic phase space V = clotron radius r, then divide the resulting arclength by CH VTOT − VREG. kpk to get δτc. Applying Kac’s lemma to get the mean iterations to As kpk = 1 per convention, we can substitute return to the stem and multiplying by ∆t, we get a result p1 − p2 = |p |. However, we know that p < 0 as the t n t for the mean return time κS(h, w, r) to the set S. reference particle must have been moving towards the obstacle to collide with it. Therefore, we can further V (h, w) κ (h, w) = CH ∆t. (B3) simplify our result to obtain S πw hδq, ni δτ = − . (A25) To eliminate ∆t, we divide by κS for w = 2 to get c hp, ni VCH(h, w) 2 This is the same result as for linear propagation in [24, κS(h, w) = κS(h, 2). (B4) VTOT(h, 2) w Eq. 18], since higher orders of δΓ were neglected. Combining eqs. (A12), (A15) and (A25), we receive As this equation no longer depends on , we can now take the final result the limit  → 0. The reason to use w = 2 here is because ! the MB becomes the stadium billiard for w = 2 (and thus 0 δq − 2 (δq · n) n is fully chaotic with no regular components with measure δΓ = hδq,ei 0 − δp − 2 (δp · n) n − 2γr cos φ e > 0). ! We can find κ (h, 2) because of the of the (δq · n) 0 S ω . (A26) MB for w = 2. Specifically, it holds that κS(h, 2) = (p · n) · p S nS(h)×τ with nS the mean amount of iterations to return to the stem and τ the mean collision time in the MB given by Eq. (3). We consider the boundary map of the billiard Appendix B: Mean return time to stem (Birkhoff coordinates), (ξ, sin φn), with ξ the coordinate along the boundary (i.e. the arc-length) and φn the angle In this section, we will derive an analytic expression for of incidence with respect to the normal vector at ξ. This the mean return time to the stem bottom in a mushroom coordinate system is a discrete mapping and Kac’s lemma 13 applies directly. Therefore the mean iterations to return Determining the cap transit length is more difficult as to the stem bottom are it depends both on α and δ. Using the law of sines, we can derive 2|∂Q| n = (B5) r  δ   S 2 · 2 c = cos asin cos α − α . (C3) cos(α) r To get an average result, this expression has to be in- where |∂Q| is the perimeter the boundary (the explicit tegrated both over δ and α, again using the fact that factor of 2 represents the contribution of sin φ ). Using n α is cosine-distributed. Unfortunately, we were unable Eq. (3), we find to solve the integrals analytically. Therefore, we decided to approximate R c dδ by a polynomial before performing π · (2h + π/2) τ = . (B6) the second integration, yielding |∂Q| w2 w4 hci ≈ 1 − − − · · · . (C4) Combining eqs. (B5) and (B6), we receive an expression 36 1200 for the mean stem return time in the fully ergodic case

π Appendix D: Proof of Eq. (1) κ (h, 2) = (2h + π/2). (B7) S 2 We are interested in the flow of the PSB but to apply This expression can be simplified by substituting the to- Kac’s lemma we need a discrete system. Thus, we obtain tal phase space volume as defined in eq. (25), yielding a map of the flow Φt by discretizing in time (similarly ∆tε with appendixB), f := Φ , with ∆tε ∼ ε. To prove VTOT(h, 2) eq. (1) we will apply Kac’s lemma to a set S. κS(h, 2) = . (B8) 4 Let TS (B) = ∆tε × nS (B) (where B is the magnetic field) denote physical recurrence time instead of map it- Inserting this result into eq. (B4), we receive erations. Let W be a circle of radius r + ε concentric to the disk of the PSB and define the phase space subset Sε such that VCH(h, w) κMB(h, w) = . (B9) 2w Sε = {x, v : x ∈ W and v · η(x) < 0} (D1) where η(x) is the vector normal to W. The mean collision time κ of the PBS is exactly T in the limit ε → 0. Appendix C: Mean duration of chaotic episodes S To find µ(S; B), the measure of set S, we first realise that it does not depend on the magnetic field, µ(S; B) = A chaotic episode in the MB as defined above consists µ(S; 0) = µ(S). This is due to the infinitesimal width of the particle travelling from the cap head directly into of S, over which motion can always be approximated by to the stem and back up to the cap head, without any a straight line for all finite magnetic fields values (i.e. other collisions inside the cap. To determine the mean equalling the magnetic field free case). Then, using (B1) duration τc of these episodes, we can geometrically de- at B = 0, we have termine the average lengths of the trajectories, exploiting     that all trajectories are uniquely defined by the angle of ∆tε ∆tε TS (B = 0) = µ(A; B = 0) = , (D2) incidence α and the distance δ from the cap center at µ(Sε) µ(Sε) which they enter the stem. because the PSB without magnetic field is fully ergodic To simplify the calculation, it is useful to split the tra- and thus µ(A; B = 0) = µ(M) = 1 (here M denotes the jectory into the mean length of cap transit hci and the entire phase space whose measure we set for simplicity mean length of stem transit hsi where τc = 2 hci + 2 hsi. to 1). By substitution we get TS (B) = µ(A; B) × TS (0). The stem transit length can be easily computed using For small enough magnetic fields all chaotic orbits (and simple trigonometry and depends only on the angle α up to measure 0 only those) collide with the disks. In the limit ε → 0 both ∆tε and µ(Sε) go to 0 linearly with h s = . (C1) ε, therefore their ratio converges, i.e. TS → κ. Since by cos(α) definition µ(A) = gc, the portion of chaotic orbits in the PSB (because we set the entire volume to have measure As the directions of particle momenta are equidis- 1), we find that the mean collision time is given by the tributed, we know that α must be cosine-distributed. We fraction of chaotic orbits as a function of the magnetic can now integrate over α to obtain the mean s, finding field B, times the mean collision time at B = 0

Z h 1 κ(B) = gc(B) × κ(0). (D3) hsi = cos(α)dα = πh. (C2) cos α 2 14

[1] Y. G. Sinai, Russ. Math. Surv. 25, 137 (1970). 131–174. [2] L. A. Bunimovich, Funct Anal Its Appl 8, 254 (1974). [18] L. A. Ponomarenko, F. Schedin, M. I. Katsnelson, [3] L. A. Bunimovich, Commun.Math. Phys. 65, 295 (1979). R. Yang, E. W. Hill, K. S. Novoselov, and A. K. Geim, [4] H.-J. St¨ockmann and J. Stein, Phys. Rev. Lett. 64, 2215 Science 320, 356 (2008). (1990). [19] L. A. Bunimovich, Chaos Interdiscip. J. Nonlinear Sci. [5] A. H. Barnett and T. Betcke, Chaos Interdiscip. J. Non- 11, 802 (2001). linear Sci. 17, 043125 (2007). [20] N. Chernov, J. Stat. Phys. 88, 1 (1997). [6] A. D. Stone, Nature 465, 696 (2010). [21] N. Chernov and L. S. Young, in Hard Ball Systems and [7] C. Gmachl, F. Capasso, E. E. Narimanov, J. U. N¨ockel, the Lorentz Gas, Encyclopaedia of Mathematical Sci- A. D. Stone, J. Faist, D. L. Sivco, and A. Y. Cho, Science ences (Springer, Berlin, Heidelberg, 2000) pp. 89–120. 280, 1556 (1998). [22] N. Chernov and R. Markarian, Chaotic Billiards, mathe- [8] S. Koyanagi, T. Nakano, and T. Kawabe, J. Acoust. Soc. matic ed. (American Mathematical Society, 2006) p. 328. Am. 124, 719 (2008). [23] C. Dellago and H. A. Posch, Phys. Rev. E 52, 2401 [9] G. Datseris, T. Geisel, and R. Fleischmann, New Journal (1995). of Physics 21, 043051 (2019). [24] C. Dellago, H. A. Posch, and W. G. Hoover, Phys. Rev. [10] R. Fleischmann, T. Geisel, and R. Ketzmerick, Physical E 53, 1485 (1996). Review Letters 68, 1367 (1992). [25] N. I. Chernov, Funct Anal Its Appl 25, 204 (1991). [11] D. Weiss, M. L. Roukes, A. Menschig, P. Grambow, [26] P. L. Garrido, Journal of Statistical Physics 88, 807 K. von Klitzing, and G. Weimann, Phys. Rev. Lett. 66, (1997). 2790 (1991). [27] P. Gaspard and F. Baras, Phys. Rev. E 51, 5332 (1995). [12] N. I. Chernov, G. L. Eyink, J. L. Lebowitz, and Y. G. [28] H. A. Posch and R. Hirschl, in Hard Ball Systems and the Sinai, Commun.Math. Phys. 154, 569 (1993). Lorentz Gas, Encyclopaedia of Mathematical Sciences [13] R. Yagi, R. Sakakibara, R. Ebisuoka, J. Onishi, (Springer, Berlin, Heidelberg, 2000) pp. 279–314. K. Watanabe, T. Taniguchi, and Y. Iye, Physical Re- [29] O. Meplan, F. Brut, and C. Gignoux, J. Phys. Math. view B 92, 195406 (2015). Gen. 26, 237 (1993). [14] H. Maier, J. Ziegler, R. Fischer, D. Kozlov, Z. D. Kvon, [30] Z. V¨or¨os,T. Tasn´adi,J. Cserti, and P. Pollner, Phys. N. Mikhailov, S. A. Dvoretsky, and D. Weiss, Nature Rev. E 67 (2003), 10.1103/PhysRevE.67.065202. Communications 8, 1 (2017), arXiv:1708.07766. [31] M. Kac, Bull. Amer. Math. Soc. 53, 1002 (1947). [15] S. Chen, Z. Han, M. M. Elahi, K. M. M. Habib, L. Wang, [32] R. S. MacKay, J Nonlinear Sci 4, 329 (1994). B. Wen, Y. Gao, T. Taniguchi, K. Watanabe, J. Hone, [33] J. D. Meiss, Chaos 7, 139 (1997). A. W. Ghosh, and C. R. Dean, Science 353 (2016), [34] L. A. Bunimovich, Nonlinearity 31, R78 (2018). 10.1126/science.aaf5481, arXiv:1602.08182. [35] G. Datseris, J. Open Source Softw. 2, 458 (2017). [16] J. P. Bird, ed., Electron Transport in Quantum Dots [36] E. G. Altmann, A. E. Motter, and H. Kantz, Chaos 15 (Springer US, 2003). (2005), 10.1063/1.1979211. [17] C. Beenakker, in Quantum Dots: a Doorway to [37] E. G. Altmann, A. E. Motter, and H. Kantz, Physi- Nanoscale Physics (Springer Berlin Heidelberg, 2005) pp. cal Review E - Statistical, Nonlinear, and Soft Matter Physics 73, 1 (2006).