This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Follow-the-Regularized-Leader Routes to Chaos in Routing Games

Jakub Bielawski Department of Mathematics, Cracow University of Economics, Rakowicka 27, 31-510 Kraków, Poland [email protected] Thiparat Chotibut Chula Intelligent and Complex Systems, Department of Physics, Faculty of Science, Chulalongkorn University, Bangkok 10330, Thailand. [email protected], [email protected] Fryderyk Falniowski Department of Mathematics, Cracow University of Economics, Rakowicka 27, 31-510 Kraków, Poland [email protected] Grzegorz Kosiorowski Department of Mathematics, Cracow University of Economics, Rakowicka 27, 31-510 Kraków, Poland [email protected] Michał Misiurewicz Department of Mathematical Sciences, Indiana University-Purdue University Indianapolis, 402 N. Blackford Street, Indianapolis, IN 46202, USA [email protected]  and  Georgios Piliouras Engineering Systems and Design, Singapore University of Technology and Design, 8 Somapah Road, Singapore 487372 [email protected]
Abstract.

We study the emergence of chaotic behavior of Follow-the-Regularized Leader (FoReL) dynamics in games. We focus on the effects of increasing the population size or the scale of costs in congestion games, and generalize recent results on unstable, chaotic behaviors in the Multiplicative Weights Update dynamics [42, 10, 11] to a much larger class of FoReL dynamics. We establish that, even in simple linear non-atomic congestion games with two parallel links and any fixed learning rate, unless the game is fully symmetric, increasing the population size or the scale of costs causes learning dynamics to become unstable and eventually chaotic, in the sense of Li-Yorke and positive topological entropy. Furthermore, we show the existence of novel non-standard phenomena such as the coexistence of stable Nash equilibria and chaos in the same game. We also observe the simultaneous creation of a chaotic attractor as another chaotic attractor gets destroyed. Lastly, although FoReL dynamics can be strange and non-equilibrating, we prove that the time average still converges to an exact equilibrium for any choice of learning rate and any scale of costs.

\markleft

BIELAWSKI, CHOTIBUT, FALNIOWSKI, KOSIOROWSKI, MISIUREWICZ, PILIOURAS

1. Introduction

We study the dynamics of online learning in a non-atomic repeated congestion game. Namely, every iteration of the game presents a population (i.e., a continuum of players) with a choice between two strategies, and imposes on them a cost which increases with the fraction of population adopting the same strategy. In each iteration, the players update their strategy accommodating for the outcomes of previous iterations. The structure of cost function here concerns that of the congestion games, which are introduced by Rosenthal [45] and are amongst the most studied classes of games. A seminal result of [40] shows that congestion games are isomorphic to potential games; as such, numerous learning dynamics are known to converge to Nash equilibria [14, 16, 15, 28, 29, 35, 5].

A prototypical class of online learning dynamics is Follow the Regularized Leader (FoReL) [48, 22]. FoReL algorithm includes as special cases ubiquitous meta-algorithms, such as the Multiplicative Weights Update (MWU) algorithm  [2]. Under FoReL, the strategy in each iteration is chosen by minimizing the weighted (by the learning rate) sum of the total cost of all actions chosen by the players and the regularization term. FoReL dynamics are known to achieve optimal regret guarantees (i.e., be competitive with the best fixed action with hindsight), as long as they are executed with a highly optimized learning rate; i.e., one that is decreasing with the steepness of the online costs (inverse to the Lipschitz constant of the online cost functions) as well as decreasing with time TT at a rate 1/T1/\sqrt{T} [48].

Although precise parameter tuning is perfectly reasonable from the perspective of algorithmic design, it seems implausible from the perspective of behavioral game theory and modeling. For example, experimental and econometric studies based on a behavioral game theoretic learning model known as Experienced Weighted Attraction (EWA), which includes MWU as a special case, have shown that agents can use much larger learning rates than those required for the standard regret bounds to be meaningfully applicable [24, 23, 25, 9]. In some sense, such a tension is to be expected because small and optimized learning rates are designed with system stability and asymptotic optimality in mind, whereas selfish agents care more about short-term rewards which result in larger learning rates and more aggressive behavioral adaptation. Interestingly, recent work on learning in games study exactly such settings of FoReL dynamics with large, fixed step-sizes, showing that vanishing and even constant regret is possible in some game settings [4, 3].

For congestion games, it is reasonable to expect that increased demands and thus larger daily costs should result in steeper behavioral responses, as agents become increasingly agitated at the mounting delays. However, to capture this behavior we need to move away from the standard assumption of effective scaling down of the learning rate. Then, the costs increase and allow more general models that can incorporate non-vanishing regret. Thus, in this regime, FoReL dynamics in congestion games cannot be reduced to standard regret based analysis [7], or even Lyapunov function arguments (e.g., [43]), and more refined and tailored arguments will be needed.

In a series of work  [42, 10, 11], the special case of MWU was analyzed under arbitrary population, demands. In a nutshell, for any fixed learning rate, MWU becomes unstable/chaotic even in small congestion games with just two strategies/paths as long as the total demand exceeds some critical threshold, whereas for small population sizes it is always convergent. Can we extend our understanding from MWU to more general FoReL dynamics? Moreover, are the results qualitatively similar showing that the dynamic is either convergent for all initial conditions or non-convergent for almost all initial conditions, or can there be more complicated behaviors depending on the choice of the regularizer of FoReL dynamics?

Refer to caption
Figure 1. Coexistence of locally attracting Nash equilibrium (green), limit cycles, and chaos in the same congestion game. Since congestion game has an associated convex potential (cost) function Φa,b(x)=a22((1b)x2+b(1x)2)\Phi_{a,b}(x)=\frac{a^{2}}{2}\left((1-b)x^{2}+b(1-x)^{2}\right) with a unique global minimum at the Nash equilibrium bb, standard learning algorithms such as gradient-like update with a small step size will converge to the equilibrium. However, here we highlight the unusual coexistence of the attracting Nash equilibrium, limit cycles, and chaos for FoReL dynamics with log-barrier regularizer r(x)=(1x)log(1x)+xlog(x)512log(x2+x+0.11)r(x)=(1-x)\log(1-x)+x\log(x)-\frac{5}{12}\log(-x^{2}+x+0.11). The right column shows that FoReL dynamics xnx_{n} depends on the initial conditions (cyan and orange colors.) Red color encodes the dynamics initialized near the left critical point xlx_{l}, which converges to the Nash equilibrium bb. Blue color encodes the dynamics initialized near the right critical point xrx_{r}, which converge to the limit cycle of period 2 (top), and to chaotic attractors (bottom). Convergence to the Nash equilibrium arises through dynamics that lower the cost function at every successive steps (left column), while convergence to a limit cycle or a chaotic attractor incur large cost, bouncing around in the cost landscape away from the Nash equilibrium. Remarkably, despite being periodic or chaotic, we prove that the time-average of the dynamics converges exactly to the Nash equilibrium bb, independent of the interior initial conditions. The bifurcation diagram associated with b=0.61b=0.61 that demonstrates coexistence of multiple attractors in the same game is shown in Fig. 2

Our model & results. We analyze FoReL-based dynamics with steep regularizers111Steepness of the regularizer guarantees that the dynamics will be well-defined as a function of the current state of the congestion game. For details, see Section 2. in non-atomic linear congestion games with two strategies. This seemingly simple setting will suffice for the emergence of highly elaborate and unpredictable behavioral patterns. For any such game GG and an arbitrarily small but fixed learning rate ϵ\epsilon, we show that there exists a system capacity N0(G,ϵ)N_{0}(G,\epsilon) such that the system is unstable when the total demand exceeds this threshold. In such case, we observe complex non-equilibrating dynamics: periodic orbits of any period and chaotic behavior of trajectories (Section 7). A core technical result is that almost all such congestion games (i.e. unless they are fully symmetric), given sufficiently large demand, will exhibit Li-Yorke chaos and positive topological entropy (Section 7.1). In the case of games with asymmetric equilibrium flow, the bifurcation diagram is very complex (see Section 8). Li-Yorke chaos implies that there exists an uncountable set of initial conditions that gets scrambled by the dynamics. Formally, given any two initial conditions x(0),y(0)x(0),y(0) in this set, lim inftdist(x(t),y(t))=0\liminf\limits_{t\rightarrow\infty}dist(x(t),y(t))=0 while lim suptdist(x(t),y(t))>0\limsup\limits_{t\rightarrow\infty}dist(x(t),y(t))>0, meaning trajectories come arbitrarily close together infinitely often but also then move away again. In the special case where the two edges have symmetric costs (equilibrium flow is the 5050%50-50\% split), the system will still become unstable given large enough demand, but chaos is not possible. Instead, in the unstable regime, all but a measure zero set of initial conditions gets attracted by periodic orbits of period two which are symmetric around the equilibrium. Furthermore, we construct formal criteria for when the Nash equilibrium flow is globally attracting. For such systems we can prove their equilibration and thus social optimality even when standard regret bounds are not applicable (Section 6.1). Also, remarkably, whether the system is equilibrating or chaotic, we prove that the time-average flows of FoReL dynamics exhibit regularity and always converge exactly to the Nash equilibrium (Section 4).

In Section 8, for the first time, to our knowledge, we report strange dynamics arising from FoReL in congestion games. Firstly, we numerically show that for FoReL dynamics a locally attracting Nash equilibrium and chaos can coexist, see Figure 2. This is also formally proven in Section 6.2. Given the prominence of local stability analysis to equilibria for numerous game theoretic settings which are widely used in Artificial Intelligence, such as Generative Adversarial Networks (GANs), e.g., [18, 37, 33, 41, 53], we believe that this result is rather important as it reveals that local stability analysis is not sufficient to guard against chaotic behaviors even in a trivial game with one (locally stable) Nash equilibrium! Secondly, Figure 4 reveals that chaotic attractors can be non-robust. Specifically, we show that mild perturbations in the parameter can lead to the destruction of one complex attractor while another totally distinct complex attractor is born! To the best of our knowledge, these phenomena have never been reported, and thus expanding our understanding of the range of possible behaviors in game dynamics. Several more examples of complex phenomena are provided in Section 8. Finally, further calculations for entropic regularizers can be found in Appendix A.

Our findings suggest that the chaotic behavior of players using Multiplicative Weights Update algorithm in congestion games (see results from [42, 10, 11]) is not an exception but the rule. Chaos is robust and can be seen for a vast subclass of online learning algorithms. In particular, our results apply to an important subclass of regularizers, of generalized entropies, which are widely used concepts in information theory, complexity theory, and statistical mechanics [13, 51, 50]. Steep functions [36, 34, 35] and generalized entropies are also often used as regularizers in game-theoretic setting [12, 35, 8]. In particular, Havrda-Charvát-Tsallis entropy-based dynamics was studied, for instance, in [20, 26]. Lastly, the emergence of chaos is clearly a hardness type of result. Such results only increase in strength the simpler the class of examples is. Complicated games are harder to learn and it is harder for players to coordinate on an equilibrium. Thus, in more complicated games one should expect even more complicated, unpredictable behaviors.

2. Model

We consider a two-strategy congestion game (see [45]) with a continuum of players (agents), where all of the players apply the Follow the Regularized Leader (FoReL) algorithm to update their strategies [48]. Each of the players controls an infinitesimally small fraction of the flow. We assume that the total flow of all the players is equal to NN. We denote the fraction of the players adopting the first strategy at time nn as xnx_{n}. The second strategy is then chosen by 1xn1-x_{n} fraction of the players. This model encapsulates how a large population of commuters selects between the two alternative paths that connect the initial point to the end point. When a large fraction of the players adopt the same strategy, congestion arises, and the cost of choosing the same strategy increases.

Linear congestion games: We focus on linear cost functions. Specifically, the cost of each path (link, route, or strategy) is proportional to the load. By denoting cjc_{j} the cost of selecting the strategy jj (when xx relative fraction of the agents choose the first strategy),

(1) c1(x)=αNx,c2(1x)=βN(1x),c_{1}(x)=\alpha Nx,\hskip 50.0ptc_{2}(1-x)=\beta N(1-x),

where α,β>0\alpha,\beta>0 are the coefficients of proportionality. Without loss of generality we will assume throughout the paper that α+β=1\alpha+\beta=1. Therefore, the values of α\alpha and β=1α\beta=1-\alpha indicate how different the path costs are from each other.

A quantity of interest is the value of the equilibrium split; i.e., the relative fraction of players using the first strategy at equilibrium. The first benefit of this formulation is that the fraction of agents using each strategy at equilibrium is independent of the flow NN. The second benefit is that, independent of α\alpha, β\beta and NN, playing Nash equilibrium results in the optimal social cost, which is the point of contact with the Price of Anarchy research [30, 11].

2.1. Learning in congestion games with FoReL algorithms

We assume that the players at time n+1n+1 know the cost of the strategies at time nn (equivalently, the realized flow (split) (xn,1xn)(x_{n},1-x_{n})) and update their choices according to the Follow the Regularized Leader (FoReL) algorithm. Namely, in the period n+1n+1 the players choose the first strategy with probability xn+1x_{n+1} such that:

(2) xn+1=argminx(0,1)(εjn[c1(xj)x+c2(1xj)(1x)]+R(x,1x))=argminx(0,1)(εjn[αNxjx+βN(1xj)(1x)]+R(x,1x)),\displaystyle\begin{split}x_{n+1}&=\arg\min_{x\in(0,1)}\left(\varepsilon\sum_{j\leq n}\left[c_{1}(x_{j})\cdot x+c_{2}(1-x_{j})\cdot(1-x)\right]+R(x,1-x)\right)\\ &=\arg\min_{x\in(0,1)}\left(\varepsilon\sum_{j\leq n}\left[\alpha N\cdot x_{j}\cdot x+\beta N\cdot(1-x_{j})\cdot(1-x)\right]+R(x,1-x)\right),\end{split}

where c1(xj)x+c2(1xj)(1x)c_{1}(x_{j})\cdot x+c_{2}(1-x_{j})\cdot(1-x) is a total cost that is inflicted on the population of agents playing against the mix (x,1x)(x,1-x) in period jj, while R:(0,1)2R\colon(0,1)^{2}\mapsto\mathbb{R} is a regularizer which represents a “risk penalty”: namely, that term would penalize abrupt changes of strategy based on a small amount of data from previous iterations of the game. The existence of a regularizer rules out strategies that focus too much on optimizing with respect to the history of our game. A weight coefficient ε>0\varepsilon>0 of our choosing is used to balance these two terms and may be perceived as a propensity to learn and try new strategies based on new information: the larger ε\varepsilon is, the faster the players learn and the more eager they are to update their strategies. Commonly adopted as a standard assumption, the learning rate ε\varepsilon can be regarded as a small, fixed constant in the following analysis but its exact value is not of particular interest. Our analysis/results holds for any fixed choice of ε\varepsilon.

Note that FoReL can also be regarded as an instance of an exploration-exploitation dynamics under the multi-armed bandits framework in online learning [54]. In the limit ε1\varepsilon\gg 1 such that (2) is well approximated by the minimization of the cumulative expected cost

jn[c1(xj)x+c2(1xj)(1x)]=jnc2(1xj)+(jn[c1(xj)c2(1xj)])x,\sum_{j\leq n}\left[c_{1}(x_{j})\cdot x+c_{2}(1-x_{j})\cdot(1-x)\right]=\sum_{j\leq n}c_{2}(1-x_{j})+\left(\sum_{j\leq n}\left[c_{1}(x_{j})-c_{2}(1-x_{j})\right]\right)x,

the minimization yields

xn+1={0,jn[c1(xj)c2(1xj)]>0,1,jn[c1(xj)c2(1xj)]<0.x_{n+1}=\begin{cases}0,&\sum_{j\leq n}\left[c_{1}(x_{j})-c_{2}(1-x_{j})\right]>0,\\ 1,&\sum_{j\leq n}\left[c_{1}(x_{j})-c_{2}(1-x_{j})\right]<0.\end{cases}

Namely, the strategy that incurs the least cumulative cost in the past time horizon is selected with probability 1. This term thus represents exploitation dynamics in reinforcement learning and multi-armed bandits framework. In the opposite limit when ϵ1\epsilon\ll 1, (2) is well approximated by the minimization of the regularizer R(x,1x).R(x,1-x). For the Shannon entropy regularizer R(x,1x)=HS(x,1x)=xlogx+(1x)log(1x)R(x,1-x)=-H_{S}(x,1-x)=x\log x+(1-x)\log(1-x) that results in the Multiplicative Weight Update algorithm (see the details in Appendix A and Sec. 3), its minimization yields

xn+1=(1xn+1)=1/2.x_{n+1}=(1-x_{n+1})=1/2.

The entropic regularization term tends to explore every strategy with equal probabilities, neglecting the information of the past cumulative cost. Thus, this regularization term corresponds to exploration dynamics. Therefore, ε\varepsilon adjusts the tradeoff between exploration and exploitation. The continuous time variant of (2) with the Shannon entropy regularizer has been studied as models of collective adaption [47, 46], also known as Boltzmann QQ learning [27], in which the exploitation term is interpreted as behavioral adaptation whereas the exploration term represents memory loss. More recent continuous-time variants study generalized entropies as regularizers, leading to a larger class of dynamics called Escort Replicator Dynamics [20] which was analyzed extensively in [34, 35].

Motivated by the continuous-time dynamics with generalized entropies, we extend FoReL discrete-time dynamics (2) to a larger class of regularizers. For a given regularizer RR, we define an auxiliary function:

(3) r:(0,1)xR(x,1x).r\colon(0,1)\ni x\mapsto R(x,1-x)\in\mathbb{R}.

We restrict the analysis to a FoReL class of regularizers for which the dynamics implied by the algorithm is well-defined. Henceforth, we assume that RR is a steep symmetric convex regularizer, namely R𝐒𝐒𝐂R\in\operatorname{\mathbf{SSC}}, where:

𝐒𝐒𝐂={R𝒞2((0,1)2):(x,y)(0,1)2R(y,x)=R(x,y);x(0,1)r′′(x)>0;limx0+r(x)=}.\operatorname{\mathbf{SSC}}=\left\{R\in\mathcal{C}^{2}((0,1)^{2}):\ \forall_{(x,y)\in(0,1)^{2}}R(y,x)=R(x,y);\ \forall_{x\in(0,1)}r^{\prime\prime}(x)>0;\ \lim_{x\to 0^{+}}r^{\prime}(x)=-\infty\right\}.

These conditions on regularizers are not overly restrictive: the assumptions for convexity and symmetry of the regularizer are natural, and if limx0r(x)\lim_{x\to 0}r^{\prime}(x) is finite, then the dynamics of xnx_{n} from (2) will not be well-defined.

Many well-known and widely used regularizers like (negative) Arimoto entropies (Shannon entropy, Havrda-Charvát-Tsallis (HCT) entropies and log-barrier being most famous ones) and (negative) Rényi entropies, under mild assumptions, belong to 𝐒𝐒𝐂\operatorname{\mathbf{SSC}} (see Appendix A). A standard non-example is the square of the Euclidean norm R(x,1x)=x2+(1x)2R(x,1-x)=x^{2}+(1-x)^{2}.

3. The dynamics introduced by FoReL

Let R𝐒𝐒𝐂R\in\operatorname{\mathbf{SSC}}. Assume that up to the iteration n>0n>0 the trajectory (x0,x1,,xn1)(x_{0},x_{1},\ldots,x_{n-1}) was established by (2). Then

xn=argminx(0,1)(Nεjn1[αxjx+β(1xj)(1x)]+r(x)).x_{n}=\arg\min_{x\in(0,1)}\left(N\varepsilon\sum_{j\leq n-1}\left[\alpha\cdot x_{j}\cdot x+\beta\cdot(1-x_{j})\cdot(1-x)\right]+r(x)\right).

First order condition yields

r(xn)=Nεjn1[αxjβ(1xj)].r^{\prime}(x_{n})=-N\varepsilon\sum_{j\leq n-1}\left[\alpha\cdot x_{j}-\beta\cdot(1-x_{j})\right].

We know that rr is convex, therefore the sufficient and necessary condition for xn+1x_{n+1} to satisfy (2) takes form:

(4) r(xn+1)=Nεjn[αxjβ(1xj)]=r(xn)Nε[αxnβ(1xn)]=r(xn)Nε[xnβ].\displaystyle\begin{split}r^{\prime}(x_{n+1})&=-N\varepsilon\sum_{j\leq n}\left[\alpha\cdot x_{j}-\beta\cdot(1-x_{j})\right]=r^{\prime}(x_{n})-N\varepsilon\left[\alpha\cdot x_{n}-\beta\cdot(1-x_{n})\right]\\ &=r^{\prime}(x_{n})-N\varepsilon\left[x_{n}-\beta\right].\end{split}

We define Ψ:(0,1)xr(x)\Psi\colon(0,1)\ni x\mapsto-r^{\prime}(x)\in\mathbb{R}. Table 1 depicts functions Ψ\Psi for different entropic regularizers222By substituting the negative Shannon entropy as rr in (4), that is r(x)=R(x,1x)=xlogx+(1x)log(1x)r(x)=R(x,1-x)=x\log x+(1-x)\log(1-x), we obtain the Multiplicative Weights Update algorithm.. Before proceeding any further, we need to establish crucial properties of the function Ψ\Psi.

Table 1. Homeomorphisms Ψ\Psi for regularizers from 𝐒𝐒𝐂\operatorname{\mathbf{SSC}}.
regularizer r(x)r(x) Ψ(x)\Psi(x)
Shannon xlogx+(1x)log(1x)x\log x+(1-x)\log(1-x) log1xx\log\frac{1-x}{x}
Havrda-Charvát-Tsallis, q(0,1)q\in(0,1) 11q(1xq(1x)q)\frac{1}{1-q}(1-x^{q}-(1-x)^{q}) q1q(xq1(1x)q1)\frac{q}{1-q}\left(x^{q-1}-(1-x)^{q-1}\right)
Rényi, q(0,1)q\in(0,1) 1q1log(xq+(1x)q)\frac{1}{q-1}\log(x^{q}+(1-x)^{q}) q1qxq1(1x)q1xq+(1x)q\frac{q}{1-q}\cdot\frac{x^{q-1}-(1-x)^{q-1}}{x^{q}+(1-x)^{q}}
log-barrier logxlog(1x)-\log x-\log(1-x) 1x11x\frac{1}{x}-\frac{1}{1-x}
Proposition 3.1.

Let Ψ\Psi be a function derived from a regularizer from 𝐒𝐒𝐂\operatorname{\mathbf{SSC}}. Then

  1. i)

    Ψ(1x)=Ψ(x)\Psi(1-x)=-\Psi(x) for x(0,1)x\in(0,1).

  2. ii)

    Ψ\Psi is a homeomorphism, limx0+Ψ(x)=\lim\limits_{x\to 0^{+}}\Psi(x)=\infty, and limx1Ψ(x)=\lim\limits_{x\to 1^{-}}\Psi(x)=-\infty.

Proof.

Due to the condition R(x,y)=R(y,x)R(x,y)=R(y,x), we have that Rx(x,1x)=Ry(1x,x)\frac{\partial R}{\partial x}(x,1-x)=\frac{\partial R}{\partial y}(1-x,x). Thus, if φ(x)=1x\varphi(x)=1-x, then:

Ψ(1x)\displaystyle\Psi(1-x) =[r(φ(x))]=Rx(1x,x)+Ry(1x,x)\displaystyle=-[r(\varphi(x))]^{\prime}=-\frac{\partial R}{\partial x}(1-x,x)+\frac{\partial R}{\partial y}(1-x,x)
=Ry(x,1x)+Rx(x,1x)=r(x)=Ψ(x).\displaystyle=-\frac{\partial R}{\partial y}(x,1-x)+\frac{\partial R}{\partial x}(x,1-x)=r^{\prime}(x)=-\Psi(x).

This implies (i). Moreover Ψ(x)=r′′(x)<0\Psi^{\prime}(x)=-r^{\prime\prime}(x)<0. Thus, Ψ\Psi is decreasing.

limx0+Ψ(x)=limx0+r(x)=.\lim_{x\to 0^{+}}\Psi(x)=-\lim_{x\to 0^{+}}r^{\prime}(x)=\infty.

From (i) we obtain that limx1Ψ(x)=\lim_{x\to 1^{-}}\Psi(x)=-\infty. ∎

By Proposition 3.1.ii, Ψ\Psi is a homeomorphism between (0,1)(0,1) and \mathbb{R}.

After substituting

(5) a=Nε,b=βa=N\varepsilon,\;\;b=\beta

we obtain from (4) a general formula for the dynamics

(6) xn+1=Ψ1(Ψ(xn)+a(xnb)),x_{n+1}=\Psi^{-1}\left(\Psi(x_{n})+a(x_{n}-b)\right),

where a>0,b(0,1)a>0,b\in(0,1). Thus, we introduce fa,b:[0,1][0,1]f_{a,b}\colon[0,1]\mapsto[0,1] as

(7) fa,b(x)={0,x=0Ψ1(Ψ(x)+a(xb)),x(0,1)1,x=1.f_{a,b}(x)=\begin{cases}0,\ \ x=0\\ \Psi^{-1}\left(\Psi(x)+a(x-b)\right),\ \ x\in(0,1)\\ 1,\ \ x=1\end{cases}.

By the properties of Ψ\Psi, fa,b:[0,1][0,1]f_{a,b}\colon[0,1]\mapsto[0,1] is continuous, and (7) defines a discrete dynamical system emerging from the FoReL algorithm for the pair of parameters (a,b)(a,b).

Lemma 3.2.

The following properties hold:

  1. i)

    fa,b(x)>xf_{a,b}(x)>x if and only if x<bx<b and fa,b(x)<xf_{a,b}(x)<x if and only if x>bx>b.

  2. ii)

    If φ:(0,1)(0,1)\varphi\colon(0,1)\mapsto(0,1) is given by φ(x)=1x\varphi(x)=1-x, then φfa,b=fa,1bφ.\varphi\circ f_{a,b}=f_{a,1-b}\circ\varphi.

  3. iii)

    Under the dynamics defined by (7), there exists a closed invariant and globally attracting interval I(0,1)I\subset(0,1).

Proof.

We obtain (i) directly from (7) and the fact that Ψ\Psi is decreasing.

Ψ\Psi is a homeomorphism, thus if y=Ψ(x)y=\Psi(x) for some x(0,1)x\in(0,1), then

y=Ψ(Ψ1(y))=Ψ(1Ψ1(y)).y=\Psi(\Psi^{-1}(y))=-\Psi(1-\Psi^{-1}(y)).

Hence,

Ψ1(y)=1Ψ1(y).\Psi^{-1}(-y)=1-\Psi^{-1}(y).

Now let x(0,1)x\in(0,1). Then

(φfa,b)(x)\displaystyle(\varphi\circ f_{a,b})(x) =1fa,b(x)=1Ψ1(Ψ(x)+a(xb))=Ψ1(Ψ(x)a(xb))\displaystyle=1-f_{a,b}(x)=1-\Psi^{-1}\left(\Psi(x)+a(x-b)\right)=\Psi^{-1}\left(-\Psi(x)-a(x-b)\right)
=Ψ1(Ψ(1x)+a((1x)(1b)))=(fa,1bφ)(x),\displaystyle=\Psi^{-1}\left(\Psi(1-x)+a\big{(}(1-x)-(1-b)\big{)}\right)=(f_{a,1-b}\circ\varphi)(x),

and (ii) follows.

By (i), fa,b(x)>xf_{a,b}(x)>x for x(0,b)x\in(0,b) and fa,b(x)<xf_{a,b}(x)<x for x(b,1)x\in(b,1). Therefore, there exists 0<δ1<min{b,1b}0<\delta_{1}<\min\{b,1-b\} such that |12x|>|12fa,b(x)||\frac{1}{2}-x|>|\frac{1}{2}-f_{a,b}(x)| for x(0,1)(δ1,1δ1)x\in(0,1)\setminus(\delta_{1},1-\delta_{1}). There exists also δ2>0\delta_{2}>0 such that fa,b([δ1,1δ1])(δ2,1δ2)f_{a,b}([\delta_{1},1-\delta_{1}])\subset(\delta_{2},1-\delta_{2}). Set δ=min{δ1,δ2}\delta=\min\{\delta_{1},\delta_{2}\}. Then, the interval I=[δ,1δ]I=[\delta,1-\delta] is invariant.

To complete the proof of (iii) we need to show that II is attracting. Assume that x(0,1)Ix\in(0,1)\setminus I is such that its fa,bf_{a,b}-trajectory never enters II. Since δδ1\delta\leq\delta_{1}, the distance between fa,bn(x)f_{a,b}^{n}(x) and II (that is, dI(fa,bn(x))d_{I}(f_{a,b}^{n}(x)), where dI(z)=δzd_{I}(z)=\delta-z for z[0,δ]z\in[0,\delta] and dI(z)=z(1δ)d_{I}(z)=z-(1-\delta) for z[1δ,1]z\in[1-\delta,1]) is decreasing and δ<f(δ)<1δ\delta<f(\delta)<1-\delta. Sequence dI(fa,bn(x))d_{I}(f_{a,b}^{n}(x)) is decreasing and bounded from below by 0, so it is convergent to some ϵ0\epsilon\geq 0. Therefore, the ω\omega-set of the trajectory of xx is a non-empty subset of dI1({ϵ})=Iϵ={δϵ,1δ+ϵ}d_{I}^{-1}(\{\epsilon\})=I_{\epsilon}=\{\delta-\epsilon,1-\delta+\epsilon\}. However, no non-empty subset of IϵI_{\epsilon} can be invariant (and thus, can be an ω\omega-set of a trajectory), because δϵδ1\delta-\epsilon\leq\delta_{1} and thus fa,b(Iϵ)(δϵ,1δ+ϵ)f_{a,b}(I_{\epsilon})\subset(\delta-\epsilon,1-\delta+\epsilon), and fa,b(Iϵ)Iϵ=f_{a,b}(I_{\epsilon})\cap I_{\epsilon}=\emptyset. By this contradiction, such xx does not exist, thus II is globally attracting.

4. Average behavior — Nash equilibrium is Cesáro attracting

We start by studying asymptotic behavior by looking on the average behavior of orbits. We will show that the orbits of our dynamics exhibit regular average behavior known as Cesáro attraction to the Nash equilibrium bb.

Definition 4.1.

For an interval map ff a point pp is Cesáro attracting if there is a neighborhood UU of pp such that for every xUx\in U the averages 1nk=0n1fk(x)\frac{1}{n}\sum_{k=0}^{n-1}f^{k}(x) converge to pp.

Theorem 4.2 (Cesáro attracting).

For every a>0a>0, b(0,1)b\in(0,1) and x0(0,1)x_{0}\in(0,1) we have

(8) limn1nk=0n1fa,bk(x0)=b.\lim\limits_{{n\to\infty}}\frac{1}{n}\sum_{k=0}^{n-1}f_{a,b}^{k}(x_{0})=b.
Proof.

Fix x0(0,1)x_{0}\in(0,1) and let xk=fa,bk(x0)x_{k}=f_{a,b}^{k}(x_{0}).

From (7) we get by induction that

(9) xn=fa,b(xn1)=Ψ1(Ψ(x0)+a(k=0n1(xkb))).x_{n}=f_{a,b}(x_{n-1})=\Psi^{-1}\left(\Psi(x_{0})+a\left(\sum_{k=0}^{n-1}(x_{k}-b)\right)\right).

By Lemma 3.2.iii there is δ>0\delta>0 such that there exists a closed, globally absorbing and invariant interval I(δ,1δ)I\subset(\delta,1-\delta). Thus, for sufficiently large nn

δ<xn=Ψ1(Ψ(x0)+a(k=0n1(xkb)))<1δ.\delta<x_{n}=\Psi^{-1}\left(\Psi(x_{0})+a\left(\sum_{k=0}^{n-1}(x_{k}-b)\right)\right)<1-\delta.

Ψ\Psi is decreasing, thus

Ψ(δ)>Ψ(x0)+a(k=0n1(xkb))>Ψ(1δ).\Psi(\delta)>\Psi(x_{0})+a\left(\sum_{k=0}^{n-1}(x_{k}-b)\right)>\Psi(1-\delta).

Therefore

1an(Ψ(δ)Ψ(x0))>1nk=0n1xkb>1an(Ψ(1δ)Ψ(x0)),\frac{1}{an}\left(\Psi(\delta)-\Psi(x_{0})\right)>\frac{1}{n}\sum_{k=0}^{n-1}x_{k}-b>\frac{1}{an}\left(\Psi(1-\delta)-\Psi(x_{0})\right),

so

|1nk=0n1xkb|<1anmax{|Ψ(δ)Ψ(x0)|,|Ψ(1δ)Ψ(x0)|}.\left|\frac{1}{n}\sum_{k=0}^{n-1}x_{k}-b\right|<\frac{1}{an}\max\{\left|\Psi(\delta)-\Psi(x_{0})\right|,\left|\Psi(1-\delta)-\Psi(x_{0})\right|\}.

Thus, (8) follows. ∎

Corollary 4.3.

The center of mass of any periodic orbit {x0,x1,,xn1}\{x_{0},x_{1},\ldots,x_{n-1}\} of fa,bf_{a,b} in (0,1)(0,1), namely x0+x1++xn1n\frac{x_{0}+x_{1}+\ldots+x_{n-1}}{n}, is equal to bb.

Applying the Birkhoff Ergodic Theorem, we obtain:

Corollary 4.4.

For every probability measure μ\mu, invariant for fa,bf_{a,b} and such that μ({0,1})=0\mu(\{0,1\})=0, we have

[0,1]x𝑑μ=b.\int_{[0,1]}x\;d\mu=b.

In the following sections, we will show that, despite the regularity of the average trajectories which converge to the Nash equilibrium bb, the trajectories themselves typically exhibit complex and diverse behaviors.

5. Two definitions of chaos

In this section we introduce two notions of chaotic behavior: Li-Yorke chaos and (positive) topological entropy. Most definitions of chaos focus on complex behavior of trajectories, such as Li-Yorke chaos or fast growth of the number of distinguishable orbits of length nn, detected by positivity of the topological entropy.

Definition 5.1 (Li-Yorke chaos).

Let (X,f)(X,f) be a dynamical system and (x,y)X×X(x,y)\in X\times X. We say that (x,y)(x,y) is a Li-Yorke pair if

lim infndist(fn(x),fn(y))=0,andlim supndist(fn(x),fn(y))>0.\liminf_{n\to\infty}dist(f^{n}(x),f^{n}(y))=0,\;\;\text{and}\;\;\limsup_{n\to\infty}dist(f^{n}(x),f^{n}(y))>0.

A dynamical system (X,f)(X,f) is Li-Yorke chaotic if there is an uncountable set SXS\subset X (called scrambled set) such that every pair (x,y)(x,y) with x,ySx,y\in S and xyx\neq y is a Li-Yorke pair.

Intuitively orbits of two points from the scrambled set have to gather themselves arbitrarily close and spring aside infinitely many times but (if XX is compact) it cannot happen simultaneously for each pair of points. Obviously the existence of a large scrambled set implies that orbits of points behave in unpredictable, complex way.

A crucial feature of the chaotic behavior of a dynamical system is also exponential growth of the number of distinguishable orbits. This happens if and only if the topological entropy of the system is positive. In fact positivity of topological entropy turned out to be an essential criterion of chaos [17]. This choice comes from the fact that the future of a deterministic (zero entropy) dynamical system can be predicted if its past is known (see [52, Chapter 7]) and positive entropy is related to randomness and chaos. For every dynamical system over a compact phase space, we can define a number h(f)[0,]h(f)\in[0,\infty] called the topological entropy of transformation ff. This quantity was first introduced by Adler, Konheim and McAndrew [1] as the topological counterpart of a metric (and Shannon) entropy. In general, computing topological entropy is not an easy task. However, in the context of piecewise monotone interval maps, topological entropy is equal to the exponential growth rate of the minimal number of monotone subintervals for fnf^{n}.

Theorem 5.2 ([39]).

Let ff be a piecewise monotone interval map and, for all n1n\geq 1, let mnm_{n} be the minimal cardinality of a monotone partition for fnf^{n}. Then

h(f)=limn1nlogmn=infn11nlogmn.h(f)=\lim_{n\to\infty}\frac{1}{n}\log m_{n}=\inf_{n\geq 1}\frac{1}{n}\log m_{n}.

6. Asymptotic stability of Nash equilibria

6.1. Asymptotic stability of Nash equilibria

The dynamics induced by (7) admits three fixed points: 0, 11 and bb. By Lemma 3.2.iii we know that all orbits starting from (0,1)(0,1) eventually fall into a globally attracting interval II. Thus, the points 0 and 11 are repelling. When does the Nash equilibrium bb attract all point from (0,1)(0,1)? First, we look when bb is an attracting and when it is a repelling fixed point. With this aim, we study the derivative of fa,bf_{a,b}:

fa,b(x)=(Ψ1)(Ψ(x)+a(xb))(Ψ(x)+a).f^{\prime}_{a,b}(x)=\left(\Psi^{-1}\right)^{\prime}\left(\Psi(x)+a(x-b)\right)\cdot\left(\Psi^{\prime}(x)+a\right).

Then,

(10) fa,b(b)=(Ψ1)(Ψ(b))(Ψ(b)+a)=Ψ(b)+aΨ(b).f^{\prime}_{a,b}(b)=\left(\Psi^{-1}\right)^{\prime}(\Psi(b))\cdot\left(\Psi^{\prime}(b)+a\right)=\frac{\Psi^{\prime}(b)+a}{\Psi^{\prime}(b)}.

The fixed point bb is attracting if and only if |fa,b(b)|<1\left|f^{\prime}_{a,b}(b)\right|<1, which is equivalent to the condition:

(11) |Ψ(b)+a|<Ψ(b).\left|\Psi^{\prime}(b)+a\right|<-\Psi^{\prime}(b).

Thus, the fixed point bb is attracting if and only if a(0,2Ψ(b))a\in\left(0,-2\cdot\Psi^{\prime}(b)\right) and repelling otherwise. We will answer when bb is globally attracting on (0,1)(0,1). First we will show the following auxiliary lemma.

Lemma 6.1.

Let a function g:Ig\colon I\mapsto\mathbb{R} be such that g′′′<0g^{\prime\prime\prime}<0. Then

g(x+y2)>g(x)g(y)xyg^{\prime}\left(\frac{x+y}{2}\right)>\frac{g(x)-g(y)}{x-y}

for every x,yIx,y\in I.

Proof.

Without loss of generality we can assume that x<yx<y. Then

yx2g(x+y2)xx+y2g(t)dt=xx+y2tx+y2g′′(s)dsdt\displaystyle\frac{y-x}{2}g^{\prime}\left(\frac{x+y}{2}\right)-\int_{x}^{\frac{x+y}{2}}g^{\prime}(t)\operatorname{\text{dt}}=\int_{x}^{\frac{x+y}{2}}\int_{t}^{\frac{x+y}{2}}g^{\prime\prime}(s)\operatorname{\text{ds}}\operatorname{\text{dt}}
>x+y2yx+y2tg′′(s)dsdt=x+y2yg(t)dtyx2g(x+y2),\displaystyle>\int_{\frac{x+y}{2}}^{y}\int_{\frac{x+y}{2}}^{t}g^{\prime\prime}(s)\operatorname{\text{ds}}\operatorname{\text{dt}}=\int_{\frac{x+y}{2}}^{y}g^{\prime}(t)\operatorname{\text{dt}}-\frac{y-x}{2}g^{\prime}\left(\frac{x+y}{2}\right),

where the inequality follows from the fact that g′′(s)g^{\prime\prime}(s) is smaller in the latter region while the integration is over the set of the same size. Therefore,

g(x+y2)>1yxxyg(t)dt=g(y)g(x)yx,g^{\prime}\left(\frac{x+y}{2}\right)>\frac{1}{y-x}\int_{x}^{y}g^{\prime}(t)\operatorname{\text{dt}}=\frac{g(y)-g(x)}{y-x},

which completes the proof of lemma. ∎

The following theorem answers, whether bb is globally attracting on (0,1)(0,1).

Theorem 6.2.

Let Ψ\Psi be a homeomorphism derived from a regularizer from 𝐒𝐒𝐂\operatorname{\mathbf{SSC}}. Suppose that bb is an attracting fixed point of fa,bf_{a,b}. If Ψ′′′<0\Psi^{\prime\prime\prime}<0, then trajectories of all points from (0,1)(0,1) converge to bb.

Proof.

In order to prove this theorem it is sufficient to show that fa,bf_{a,b} doesn’t have periodic orbits of period 2.

Suppose that {x0,x1}(0,1)\{x_{0},x_{1}\}\in(0,1) is a periodic orbit of fa,bf_{a,b} of period 2.

x0+x12=b.\frac{x_{0}+x_{1}}{2}=b.

We have that x1=Ψ1(Ψ(x0)+a(x0b))x_{1}=\Psi^{-1}\left(\Psi(x_{0})+a(x_{0}-b)\right), and therefore,

Ψ(x1)=Ψ(x0)+a(x0b).\Psi(x_{1})=\Psi(x_{0})+a(x_{0}-b).

Thus, Ψ(x1)Ψ(x0)=a2(x1x0),\Psi(x_{1})-\Psi(x_{0})=-\frac{a}{2}(x_{1}-x_{0}), or equivalently

(12) a=2Ψ(x1)Ψ(x0)x1x0.a=-2\cdot\frac{\Psi(x_{1})-\Psi(x_{0})}{x_{1}-x_{0}}.

By Lemma 6.1

(13) Ψ(b)>Ψ(x1)Ψ(x0)x1x0=a2,\Psi^{\prime}(b)>\frac{\Psi(x_{1})-\Psi(x_{0})}{x_{1}-x_{0}}=-\frac{a}{2},

but the point bb is attracting if and only if Ψ(b)<a2\Psi^{\prime}(b)<-\frac{a}{2}, which contradicts the inequality (13). Therefore, ff has no periodic point of period 2.

Now, by [6], Chapter VI, Proposition 1, every trajectory of ff converges to a fixed point.

Corollary 6.3.

Let Ψ′′′<0\Psi^{\prime\prime\prime}<0. Then the Nash equilibrium bb attracts all points from the open interval (0,1)(0,1) if and only if a(0,2Ψ(b))a\in(0,-2\Psi^{\prime}(b)).

Functions Ψ\Psi derived from Shannon entropy, HCT entropy or log-barrier satisfy the inequality Ψ′′′<0\Psi^{\prime\prime\prime}<0. Nevertheless, this additional condition is needed, because for an arbitrary Ψ\Psi derived from 𝐒𝐒𝐂\operatorname{\mathbf{SSC}} attracting orbits of any period may exist together with the attracting Nash equilibrium bb. In the next section we will discuss thoroughly an example of such behavior. This shows that even for the well-known class of FoReL algorithms knowledge of local behavior (even attraction) of the Nash equilibrium may not be enough to properly describe behavior of agents.

Refer to caption
Refer to caption
Figure 2. Coexistence of the attracting Nash equilibrium and chaos. The bifurcation diagrams for fa,bf_{a,b} where the dynamics is induced by the regularizer r(x)=(1x)log(1x)+xlogx0.4167log(x2+x+0.11)r(x)=(1-x)\log(1-x)+x\log x-0.4167\log(-x^{2}+x+0.11) for b=0.61b=0.61. On the horizontal axis the parameter aa is between 2.62.6 and 3.43.4, and on the vertical axis values of fa,bf_{a,b} are shown. As starting points for bifurcation diagrams two critical points of fa,bf_{a,b} are taken — red refers to the critical point in (0,0.5)(0,0.5) and blue the critical point in (0.5,1)(0.5,1). Each critical point is iterated 4000 times, visualizing the last 200 iterates. On the top picture first red and then blue trajectories are drawn, and on the bottom one the order is reversed. Function Ψ(x)=r(x)=log(1x)logx+0.4167[11.1x1x+0.1]\Psi(x)=-r^{\prime}(x)=\log(1-x)-\log x+0.4167[\frac{1}{1.1-x}-\frac{1}{x+0.1}] fulfills all assumptions of Theorem 6.2 excluding Ψ′′′<0\Psi^{\prime\prime\prime}<0. Although for a<2Ψ(b)3.28a<-2\Psi^{\prime}(b)\approx 3.28 the unique Nash equilibrium is attracting we can observe chaotic behavior already for a>3.15a>3.15. The picture suggests that in the coexistence region we have an interval which is invariant for fa,b2f_{a,b}^{2}, and in it we see the usual evolution of unimodal maps. This means that sometimes we see an attracting periodic orbit, sometimes a chaotic attractor.

6.2. Coexistence of attracting Nash equilibrium and chaos

In this section we will describe an example of the regularizer from 𝐒𝐒𝐂\operatorname{\mathbf{SSC}}, which introduces game dynamics in which attracting Nash equilibrium coexist with chaos, see example in Figure 2. This phenomenon is observed by replacing the Shannon entropic regularizer by the log-barrier regularizer. Namely, we take

Ψ(x)=log(1x)logx+0.4167(11.1x1x+0.1).\Psi(x)=\log(1-x)-\log x+0.4167\cdot\left(\frac{1}{1.1-x}-\frac{1}{x+0.1}\right).

We will show that there exist a>0a>0, b(0,1)b\in(0,1) such that fa,bf_{a,b} has an attracting fixed point (which is the Nash equilibrium) yet the map can be chaotic!

Proposition 6.4.

There exist a>0a>0, b(0,1)b\in(0,1) such that fa,bf_{a,b} has an attracting fixed point (Nash equilibirum), positive topological entropy and is Li-Yorke chaotic.

Proof.

Let us take b=0.61b=0.61 and a=3.25a=3.25 (see the graph of fa,bf_{a,b} in Figure 3).

Refer to caption
Figure 3. Graph of fa,bf_{a,b} for a=3.25a=3.25, b=0.61b=0.61 generated by the regularizer r(x)=(1x)log(1x)+xlogx0.4167log(x2+x+0.11)r(x)=(1-x)\log(1-x)+x\log x-0.4167\log(-x^{2}+x+0.11).

Set

ξ(x):=Ψ(x)+a(xb)=log(1x)logx+0.4167(11.1x1x+0.1)+3.25(x0.61).\xi(x):=\Psi(x)+a(x-b)=\log(1-x)-\log x+0.4167\cdot\left(\frac{1}{1.1-x}-\frac{1}{x+0.1}\right)+3.25\cdot(x-0.61).

Since a<3.2825965210953082Ψ(b)a<3.282596521095308\approx-2\Psi^{\prime}(b), the fixed point bb is attracting.

To show that fa,bf_{a,b} is chaotic, we will prove that fa,bf_{a,b} has a periodic point of period 6. With this aim, we will show that (f2)3(x)<x<f2(x)(f^{2})^{3}(x)<x<f^{2}(x) for any x[0.9559,0.956]x\in[0.9559,0.956]. We start by showing that ξ(x)\xi(x) is monotone on [0.9559,0.956][0.9559,0.956]. Formula for the derivative of ξ(x)\xi(x) is

ξ(x)=11x1x+0.4167(1(1.1x)2+1(x+0.1)2)+3.25.\xi^{\prime}(x)=-\frac{1}{1-x}-\frac{1}{x}+0.4167\cdot\left(\frac{1}{(1.1-x)^{2}}+\frac{1}{(x+0.1)^{2}}\right)+3.25.

Set z=(x0.5)2z=(x-0.5)^{2}. Then ξ(x)=0\xi^{\prime}(x)=0 if and only if g(z)=0g(z)=0, where

g(z)=3.25z31.3191z2+0.377874z0.050706.g(z)=3.25\cdot z^{3}-1.3191\cdot z^{2}+0.377874\cdot z-0.050706.

We have

g(z)=9.75z22.6382z+0.377874,g^{\prime}(z)=9.75\cdot z^{2}-2.6382\cdot z+0.377874,

and the discriminant of this quadratic polynomial is negative. Therefore, gg has only one zero (approximately 0.20772597686456770.2077259768645677), so ξ\xi^{\prime} has only two zeros, symmetric with respect to 0.5. Thus, as

ξ(0.9559)0.03051955745677404,ξ(0.956)0.05413532613604133\xi^{\prime}(0.9559)\approx-0.03051955745677404,\;\;\xi^{\prime}(0.956)\approx-0.05413532613604133

there is no zero of ξ\xi^{\prime} between these two points. Moreover, those computations give us an approximation to both zeros of ξ\xi^{\prime}: 0.04423034670508420.0442303467050842 and 0.95576965329491580.9557696532949158.

Now we look at the first six images of [0.9559,0.956][0.9559,0.956]:

Ψ(0.063)\Psi(0.063)\tabto51pt 0.5449390463486314\approx 0.5449390463486314
ξ(0.956)\xi(0.956)\tabto51pt 0.5450794481395858\approx 0.5450794481395858
ξ(0.9559)\xi(0.9559)\tabto51pt 0.5450836794177281\approx 0.5450836794177281
Ψ(0.062)\Psi(0.062)\tabto51pt 0.5458384284441133\approx 0.5458384284441133

Ψ(0.991)\Psi(0.991)\tabto48pt 1.26049734857964\approx-1.26049734857964
ξ(0.062)\xi(0.062)\tabto48pt 1.235161571555887\approx-1.235161571555887
ξ(0.063)\xi(0.063)\tabto48pt 1.232810953651368\approx-1.232810953651368
Ψ(0.99)\Psi(0.99)\tabto48pt 1.189231609934426\approx-1.189231609934426

Ψ(0.52)\Psi(0.52)\tabto47pt 0.03369120600501601\approx-0.03369120600501601
ξ(0.991)\xi(0.991)\tabto47pt 0.02224734857964017\approx-0.02224734857964017
ξ(0.99)\xi(0.99)\tabto47pt 0.04576839006557365\approx 0.04576839006557365
Ψ(0.47)\Psi(0.47)\tabto47pt 0.05052025169168718\approx 0.05052025169168718

Ψ(0.76)\Psi(0.76)\tabto44pt 0.4116261583651984\approx-0.4116261583651984
ξ(0.47)\xi(0.47)\tabto44pt 0.4044797483083129\approx-0.4044797483083129
ξ(0.52)\xi(0.52)\tabto44pt 0.3261912060050159\approx-0.3261912060050159
Ψ(0.69)\Psi(0.69)\tabto44pt 0.3112461911278587\approx-0.3112461911278587

Ψ(0.54)\Psi(0.54)\tabto44pt 0.06732925721803665\approx-0.06732925721803665
ξ(0.69)\xi(0.69)\tabto44pt 0.05124619112785883\approx-0.05124619112785883
ξ(0.76)\xi(0.76)\tabto44pt 0.07587384163480165\approx 0.07587384163480165
Ψ(0.45)\Psi(0.45)\tabto44pt 0.08411125490271053\approx 0.08411125490271053

Ψ(0.8)\Psi(0.8)\tabto44pt 0.4602943611198909\approx-0.4602943611198909
ξ(0.45)\xi(0.45)\tabto44pt 0.4358887450972894\approx-0.4358887450972894
ξ(0.54)\xi(0.54)\tabto44pt 0.2948292572180365\approx-0.2948292572180365
Ψ(0.65)\Psi(0.65)\tabto44pt 0.2486392084062237\approx-0.2486392084062237.

We have fa,b(x)=Ψ1(ξ(x))f_{a,b}(x)=\Psi^{-1}(\xi(x)), so Ψ(fa,b(x))=ξ(x)\Psi(f_{a,b}(x))=\xi(x). Write x,y\langle x,y\rangle for [x,y][x,y] or [y,x][y,x]. If ξ(x),ξ(y)Ψ(z),Ψ(w)\langle\xi(x),\xi(y)\rangle\subset\langle\Psi(z),\Psi(w)\rangle and ξ\xi is monotone on x,y\langle x,y\rangle, then fa,b(x),fa,b(y)z,w\langle f_{a,b}(x),f_{a,b}(y)\rangle\subset\langle z,w\rangle. Thus, the computations show that

fa,b2([0.9559,0.956])[0.99,0.991]f_{a,b}^{2}([0.9559,0.956])\subset[0.99,0.991]

and

fa,b6([0.9559,0.956])[0.65,0.8].f_{a,b}^{6}([0.9559,0.956])\subset[0.65,0.8].

Therefore, for any x[0.9559,0.956]x\in[0.9559,0.956] we have

(fa,b2)3(x)<x<(fa,b2)(x),(f_{a,b}^{2})^{3}(x)<x<(f_{a,b}^{2})(x),

so by theorem from [31], fa,b2f_{a,b}^{2} has a periodic point of period 3 and fa,bf_{a,b} has a periodic point of period 6. Thus, because fa,bf_{a,b} has a periodic point of period that is not a power of 2, the topological entropy h(fa,b)h(f_{a,b}) is positive (see [38]) and it is Li-Yorke chaotic. ∎

Corollary 6.5.

There exist FoReL dynamics such that when applied to symmetric linear congestion games with only two strategies/paths the resulting dynamics have

  • a set of positive measure of initial conditions that converge to the unique and socially optimum Nash equilibrium and

  • an uncountable scrambled set for which trajectories exhibit Li-Yorke chaos,

  • periodic orbits of all possible even periods.

Thus, the (long-term) social cost depends critically on the initial condition.

FoReL dynamics induced by this regularizer manifests drastically different behaviors that depend on the initial condition.

7. Behavior for sufficiently large aa

7.1. Non-convergence for sufficiently large aa

In this subsection, we study what happens as we fix bb and let aa be arbitrarily large333By (5) it reflects the case when we fix cost functions (and learning rate ε\varepsilon) and increase the total demand NN.. First, we study the asymmetric case, namely b1/2b\neq 1/2. We show chaotic behavior of our dynamical system for aa sufficiently large, that is we will show that if aa is sufficiently large then fa,bf_{a,b} is Li-Yorke chaotic, has periodic orbits of all periods and positive topological entropy.

The crucial ingredient of our analysis is the existence of periodic orbit of period 3.

Theorem 7.1.

If b(0,1)\{12}b\in(0,1)\backslash\{\frac{1}{2}\}, then there exists aba_{b} such that if a>aba>a_{b} then fa,bf_{a,b} has a periodic point of period 3.

Proof.

By Lemma 3.2.ii, without loss of generality, we may assume b(0,12)b\in(0,\frac{1}{2}). We will show that there exists x0(0,1)x_{0}\in(0,1) such that fa,b3(x0)<x0<fa,b(x0)f_{a,b}^{3}(x_{0})<x_{0}<f_{a,b}(x_{0}).

Fix a>0a>0 and b,x(0,1)b,x\in(0,1). We set xn=fa,bn(x)x_{n}=f_{a,b}^{n}(x), then formula (9) holds. Hence fa,b(x)>xf_{a,b}(x)>x if and only if x<bx<b and, because Ψ1\Psi^{-1} is decreasing, fa,b3(x)<xf_{a,b}^{3}(x)<x is equivalent to x+fa,b(x)+fa,b2(x)>3bx+f_{a,b}(x)+f_{a,b}^{2}(x)>3b.

From the fact that b(0,12)b\in(0,\frac{1}{2}) we have that 3b1<b3b-1<b. So we can take x0>0x_{0}>0 such that 3b1<x0<b3b-1<x_{0}<b. Then fa,b(x0)>x0f_{a,b}(x_{0})>x_{0}. Moreover

limafa,b(x0)=limaΨ1(Ψ(x0)+a(x0b))=1.\lim_{a\to\infty}f_{a,b}(x_{0})=\lim_{a\to\infty}\Psi^{-1}\left(\Psi(x_{0})+a(x_{0}-b)\right)=1.

Thus, since 3bx0<13b-x_{0}<1, there exists ab>0a_{b}>0 such that if a>aba>a_{b}, then fa,b(x0)>3bx0f_{a,b}(x_{0})>3b-x_{0}, so x0+fa,b(x0)+fa,b2(x0)>3bx_{0}+f_{a,b}(x_{0})+f_{a,b}^{2}(x_{0})>3b. Hence, if a>aba>a_{b}, then fa,b3(x0)<x0f_{a,b}^{3}(x_{0})<x_{0}.

Now we conclude that fa,bf_{a,b} has a periodic point of period 3 for a>aba>a_{b}, from theorem from [31], which implies that if fn(x)<x<f(x)f^{n}(x)<x<f(x) for some odd n>1n>1, then ff has a periodic point of period nn.

By the Sharkovsky Theorem ([49]), existence of a periodic orbit of period 3 implies existence of periodic orbits of all periods, and by the result of [32], period 3 implies Li-Yorke chaos. Moreover, because fa,bf_{a,b} has a periodic point of period that is not a power of 2, the topological entropy h(fa,b)h(f_{a,b}) is positive (see [38]). Thus:

Corollary 7.2.

If b(0,1){1/2}b\in(0,1)\setminus\{1/2\}, then there exists aba_{b} such that if a>aba>a_{b} then fa,bf_{a,b} has periodic orbits of all periods, has positive topological entropy and is Li-Yorke chaotic.

This result has an implication in non-atomic routing games. Recall that the parameter aa expresses the normalized total demand. Thus, Corollary 7.2 implies that when the costs (cost functions) of paths are different, then increasing the total demand of the system will inevitably lead to chaotic behavior.

Now we consider the symmetric case, when b=12b=\frac{1}{2}, which corresponds to equal coefficients of the cost functions, α=β\alpha=\beta. To simplify the notation we denote fa=fa,1/2f_{a}=f_{a,1/2}.

Theorem 7.3.

If the parameter aa is small enough, then all trajectories of faf_{a} starting from (0,1)(0,1) converge to the attracting fixed point 1/21/2. There exists aba_{b} such that if a>aba>a_{b}, then all points from (0,1)(0,1) (except countably many points, whose trajectories eventually fall into the repelling point 1/21/2) are attracted by periodic attracting orbits of the form {σa,1σa}\{\sigma_{a},1-\sigma_{a}\}, where 0<σa<1/20<\sigma_{a}<1/2. Moreover, if there exists δ>0\delta>0 such that Ψ\Psi is convex on (0,δ)(0,\delta), then there exists a unique attracting orbit {σa,1σa}\{\sigma_{a},1-\sigma_{a}\}, which attracts trajectories of all points from (0,1)(0,1), except countably many points, whose trajectories eventually fall into the repelling fixed point 1/21/2.

Proof.

By Lemma 3.2.ii the maps faf_{a} and φ\varphi commute. Set ga=φfa=faφg_{a}=\varphi\circ f_{a}=f_{a}\circ\varphi. Since φ\varphi is an involution, we have ga2=fa2g_{a}^{2}=f_{a}^{2}. We show that the dynamics of faf_{a} is simple, no matter how large aa is.

We aim to find fixed points and points of period 2 of faf_{a} and gag_{a}. Clearly,

fa(0)=0,fa(1)=1,ga(0)=1,ga(1)=0.f_{a}(0)=0,\ \ \ f_{a}(1)=1,\ \ \ g_{a}(0)=1,\ \ \ g_{a}(1)=0.

By (9) we have

fa2(x)=Ψ1(Ψ(x)+a(x+fa(x)1)),f_{a}^{2}(x)=\Psi^{-1}(\Psi(x)+a(x+f_{a}(x)-1)),

so the fixed points of fa2f_{a}^{2} are 0, 11 and the solutions to x+fa(x)1=0x+f_{a}(x)-1=0, that is, to ga(x)=xg_{a}(x)=x. Thus, the fixed points of ga2g_{a}^{2} (which, as we noticed, is equal to fa2f_{a}^{2}) are the fixed points of gag_{a} and 0 and 11.

We can choose the invariant interval Ia=Ia,1/2I_{a}=I_{a,1/2} symmetric, so that φ(Ia)=Ia\varphi(I_{a})=I_{a}. Let us look at Ga=ga|Ia:IaIaG_{a}=g_{a}|_{I_{a}}:I_{a}\to I_{a}. All fixed points of Ga2G_{a}^{2} are also fixed points of GaG_{a}, so GaG_{a} has no periodic points of period 2. By the Sharkovsky Theorem, GaG_{a} has no periodic points other than fixed points. For such maps it is known (see, e.g., [6]) that the ω\omega-limit set of every trajectory is a singleton of a fixed point, that is, every trajectory converges to a fixed point. If x(0,1)Iax\in(0,1)\setminus I_{a}, then the gag_{a}-trajectory of xx after a finite time enters IaI_{a}, so gag_{a}-trajectories of all points of (0,1)(0,1) converge to a fixed point of gag_{a} in IaI_{a}. Observe that a fixed point of gag_{a} can be a fixed point of faf_{a} (other that 0, 1) or a periodic point of faf_{a} of period 2. Thus, the faf_{a}-trajectory of every point of (0,1)(0,1) converges to a fixed point or a periodic orbit of period 22 of faf_{a}, other than 0 and 11.

Observe now that 1/21/2 is a fixed point of both faf_{a} and gag_{a}. The fixed points of gag_{a} in [0,1/2][0,1/2] are the solutions of the equation ga(x)=xg_{a}(x)=x, which is equivalent to fa(x)=1xf_{a}(x)=1-x, further to

Ψ(x)+a(x1/2)=Ψ(1x)\Psi(x)+a(x-1/2)=\Psi(1-x)

and finally, by Proposition 3.1.i, to

2Ψ(x)=a(x1/2).2\Psi(x)=-a(x-1/2).

Define γa(x)=a/2(x1/2)\gamma_{a}(x)=-a/2(x-1/2). We look for σa(0,1/2)\sigma_{a}\in(0,1/2) such that

(14) Ψ(σa)=γa(σa).\Psi(\sigma_{a})=\gamma_{a}(\sigma_{a}).

We know that Ψ(1/2)=γa(1/2)=0\Psi(1/2)=\gamma_{a}(1/2)=0 and γa(1/2)=a2\gamma_{a}^{\prime}(1/2)=-\frac{a}{2}. As Ψ(1/2)<0\Psi^{\prime}(1/2)<0 (Ψ\Psi is strictly decreasing) there is no solution of (14) in (0,1/2)(0,1/2) for sufficiently small aa. Then, 1/21/2 is the only fixed point of gag_{a} in (0,1)(0,1). Thus, 1/21/2 will attract all points from (0,1)(0,1).

If γa(1/2)<Ψ(1/2)\gamma_{a}^{\prime}(1/2)<\Psi^{\prime}(1/2), then there exists x(0,1/2)x\in(0,1/2) such that γa(x)>Ψ(x)\gamma_{a}(x)>\Psi(x). Because limx0+Ψ(x)=+\lim_{x\to 0^{+}}\Psi(x)=+\infty and limx0+γa(x)=a/4\lim_{x\to 0^{+}}\gamma_{a}(x)=a/4 and both functions are continuous, there exists σa(0,1/2)\sigma_{a}\in(0,1/2) such that Ψ(σa)=γa(σa)\Psi(\sigma_{a})=\gamma_{a}(\sigma_{a}). Finally, γa(1/2)<Ψ(1/2)\gamma_{a}^{\prime}(1/2)<\Psi^{\prime}(1/2) if and only if a>2Ψ(1/2)a>-2\Psi^{\prime}(1/2).

Lastly, if Ψ\Psi is convex on some neighborhood of zero, that is (0,δ)(0,\delta), then we can choose aa (sufficiently large) such that all solutions of (14) in (0,12)(0,\frac{1}{2}) lay in (0,δ)(0,\delta). From the fact that Ψ\Psi is convex on this interval and γa\gamma_{a} is an affine function we obtain uniqueness of σa\sigma_{a}. ∎

Theorem 7.3 has a remarkable implication in non-atomic routing games. It implies that if cost of both paths is the same, then there is a threshold such that if the total demand will cross this threshold, then starting from almost any initial condition the system will oscillate, converging to the symmetric periodic orbit of period 2, never converging to the Nash equilibrium.

8. Experimental results

Refer to caption
Refer to caption
Figure 4. Simultaneous creation and destruction of different attractors. The bifurcation diagrams for fa,bf_{a,b} where the dynamics is determined by taking (negative) log-barrier regularizer with parameter b=0.61b=0.61. On the horizontal axis the parameter aa is between 146.97146.97 and 147147, and on the vertical axis values of fa,bf_{a,b} between 0.270.27 and 0.340.34 are shown. As starting points for bifurcation diagrams two critical points of fa,bf_{a,b} are taken (regularity of this map, see Appendix B, guarantees that their trajectories detect all attractors). — red refers to the critical point in (0,0.5)(0,0.5) and blue to the critical point in (0.5,1)(0.5,1). Each critical point is iterated 4000 times, visualizing the last 200 iterates. On the top picture first red and then blue trajectories are drawn and on the bottom one first blue and then red. We observe the collapse of the red attractor (built on the left critical point) with the simultaneous creation of the blue one (built on the right critical point).

In this section we report complex behaviors in bifurcation diagrams of FoReL dynamics. We investigate the structures of the attracting periodic orbits and chaotic attractors associated with the interval map fa,b:[0,1][0,1]f_{a,b}:[0,1]\mapsto[0,1] defined by (7). In the asymmetric case, that is when bb differs from 0.50.5, the standard equilibrium analysis applies when the fixed point bb is stable, which is when |fa,b(b)|1|f^{\prime}_{a,b}(b)|\leq 1, or equivalently when a2Ψ(b)a\leq-2\Psi^{\prime}(b). Therefore, as we argued in the previous section, in this case the dynamics will converge toward the fixed point bb whenever a<2Ψ(b)a<-2\Psi^{\prime}(b). However, when a2Ψ(b)a\geq-2\Psi^{\prime}(b) there is no attracting fixed point. Moreover, a chaotic behavior of trajectories emerges when aa is sufficiently large, as the period-doubling bifurcations route to chaos is guaranteed to arise.

In particular, we study the attractors of the map fa,bf_{a,b} generated by the log-barrier regularizer (see Example A.2 with η(x)=logx\eta(x)=\log x) and by the Havrda-Charvát-Tsallis regularizer for q=0.5q=0.5 (see Example A.3). Note that for both of these regularizers, we have that Ψ′′′<0\Psi^{\prime\prime\prime}<0. Note also that the functions Ψ\Psi for these regularizers can be found in Table 1.

We first focus on the log-barrier regularizer444From the regularity of the map fa,bf_{a,b} (see Appendix B), we know that every limit cycle of the dynamics generated by fa,bf_{a,b} can be found by studying the behavior of the critical points of fa,bf_{a,b}. Therefore, all attractors of this dynamics can be revealed by following the trajectories of these two critical points (as in Figure 4).. Figure 4 reveals an unusual bifurcation phenomenon, which, to our knowledge, is not known in other natural interval maps. We observe simultaneous evolution of two attractors in the opposite directions: one attractor, generated by the trajectory of the left critical point, is shrinking, while the other one, generated by the trajectory of the right critical point, is growing. Figure 5 shows another unusual bifurcation phenomenon: a chaotic attractor arises via period-doubling bifurcations and then collapses. After that, the trajectories of the critical points, one after the other, jump, and then they together follow a period-doubling route to chaos once more.

Finally we study the bifurcation diagrams generated from Havrda-Charvát-Tsallis regularizer with q=0.5q=0.5. In Figure 8 we observe a finite number of period-doubling (and period-halving) bifurcations, a behavior that does not lead to chaos. Nevertheless, as aa increases from 39.915 to 39.93, the trajectory of the right critical point leaves the attractor which it shared with the trajectory of the left critical point, and builds a separate chaotic attractor. When chaos arises, however, we observe that the induced dynamics of the log-barrier regularizer and of the Havrda-Charvát-Tsallis regularizer with q=0.5q=0.5 both exhibit period-doubling routes to chaos though the regularizers are starkly different, see Figure 6 and Figure 7 respectively.

Refer to caption
Refer to caption
Figure 5. Locally complex behavior. The bifurcation diagrams for fa,bf_{a,b} where the dynamics is determined by taking (negative) log-barrier as the regularizer for b=0.61b=0.61. On the horizontal axis the parameter aa is between 153153 and 156.5156.5, and on the vertical axis values of fa,bf_{a,b} are between 0.030.03 and 0.130.13. As starting points for bifurcation diagrams two critical points of fa,bf_{a,b} are taken — red refers to the critical point in (0,0.5)(0,0.5) and blue the critical point in (0.5,1)(0.5,1). Each critical point is iterated 4000 times, then visualizing the last 200 iterates. On the top picture first red and then blue trajectories are drawn, and on the bottom one the order is reversed. As aa increases chaotic behavior of orbits disappears (around 153.25153.25). Then, within the window [153.25,156.3][153.25,156.3], chaos emerges at [153.5,154][153.5,154] and vanishes. Then trajectories jump, one after the other, and then generate a chaotic attractor which then spreads, vanishes, and finally spreads onto the whole interval.
Refer to caption
Refer to caption
Figure 6. Period-doubling road to chaos. The bifurcation diagrams for fa,bf_{a,b} where the dynamics is determined by taking (negative) log-barrier as the regularizer: r(x)=logxlog(1x)r(x)=-\log x-\log(1-x) for b=0.61b=0.61. On the horizontal axis the parameter aa is between 1010 and 190190, and on the vertical axis the value of fa,bf_{a,b} ranges between 0 and 11. As starting points for bifurcation diagrams two critical points of fa,bf_{a,b} are taken (regularity of this map, see Appendix B, guarantees that by studying their trajectories we visit all attractors) — red refers to the left critical point (in (0,0.5)(0,0.5)) and blue to the right critical point (in (0.5,1)(0.5,1)). Each critical point is iterated 4000 times, visualizing the last 200 iterates. On the top picture first red and then blue trajectories are drawn, and on the bottom one the order is reversed. The first bifurcation takes place at the moment when the Nash equilibrium bb becomes repelling. Then we observe period-doubling route to chaos. In addition two different attractors are visible for a(92,96)a\in(92,96).
Refer to caption
Refer to caption
Figure 7. Period-doubling road to chaos with Havrda-Charvát-Tsallis regularizer. The bifurcation diagrams for fa,bf_{a,b} where the dynamics is determined by taking (negative) Havrda-Charvát-Tsallis entropy with q=0.5q=0.5 as the regularizer and b=0.61b=0.61. On the horizontal axis the parameter aa is between 44 and 4646, and on the vertical axis values of fa,bf_{a,b} ranges between 0 and 11. As starting points for bifurcation diagrams two critical points of fa,bf_{a,b} are taken — red refers to the critical point in (0,0.5)(0,0.5) and blue the critical point in (0.5,1)(0.5,1). Each critical point is iterated 4000 times, then visualizing the last 200 iterates. On the top picture first red and then blue trajectories are drawn, and on the bottom one the order is reversed. The first bifurcation takes place at the moment when the Nash equilibrium bb becomes repelling. Then we observe period-doubling route to chaos. In addition two different attractors are visible for a(22.5,24.5)a\in(22.5,24.5).
Refer to caption
Refer to caption
Figure 8. Period-doubling not always lead to chaos.The bifurcation diagrams for fa,bf_{a,b} where the dynamics is determined by taking (negative) Havrda-Charvát-Tsallis entropy with q=0.5q=0.5 as the regularizer, that is, r(x)=1x11xr(x)=\frac{1}{\sqrt{x}}-\frac{1}{\sqrt{1-x}}. We fix b=0.61b=0.61. On the horizontal axis the parameter aa is between 39.7539.75 and 4040, and on the vertical axis values of fa,bf_{a,b} are between 0.650.65 and 0.80.8. As starting points for bifurcation diagrams two critical points of fa,bf_{a,b} are taken — red refers to the critical point in (0,0.5)(0,0.5) and blue the critical point in (0.5,1)(0.5,1). Each critical point is iterated 4000 times, then visualizing the last 200 iterates. On the top picture first red and then blue trajectories are drawn, and on the bottom one the order is reversed. As aa increases both trajectories go through the same forward and backward period doubling steps. Then, as aa increases from 39.91539.915 to 39.9339.93, the trajectory of the right critical point escapes the attractor which she shared with the trajectory of the left critical point, and builds separate chaotic attractor. Then it jumps back to the red attractor.

9. Conclusion

We study FoReL dynamics in non-atomic congestion games with arbitrarily small but fixed step-sizes, rather than with decreasing and regret-optimizing step-sizes. Our model allows for agents that can learn over time (e.g., by tracking the cumulative performance of all actions to inform about their future decisions), while being driven by opportunities for short-term rewards, rather than only by long-term asymptotic guarantees. As a result, we can study the effects of increasing system demand and delays on agents’ responses, which can become steeper as they are increasingly agitated by the increasing costs. Such assumptions are well justified from a behavioral game theory perspective [24, 9]; however, FoReL dynamics are pushed outside of the standard parameter regime in which classic black-box regret bounds do not apply meaningfully. Using tools from dynamical systems, we show that, under sufficiently large demand, dynamics will unavoidably become chaotic and unpredictable. Thus, our work vastly generalizes previous results that hold in the special case of Multiplicative Weights Update [42, 10, 11]. We also report a variety of undocumented complex behaviors such as the co-existence of a locally attracting Nash equilibrium and of chaos in the same game. Despite this behavioral complexity of the day-to-day behavior, the time-average system behavior is always perfectly regular, converging to an exact equilibrium. Our analysis showcases that local stability in congestion games should not be considered as a foregone conclusion and paves the way toward further investigations at the intersection of optimization theory, (behavioral) game theory, and dynamical systems.

Acknowledgements

Georgios Piliouras acknowledge AcRF Tier 2 grant 2016-T2-1-170, grant PIE-SGP-AI-2018-01, NRF2019-NRF- ANR095 ALIAS grant and NRF 2018 Fellowship NRF-NRFF2018-07. Fryderyk Falniowski acknowledges the support of the National Science Centre, Poland, grant 2016/21/D/HS4/01798 and COST Action CA16228 “European Network for Game Theory”. Research of Michał Misiurewicz was partially supported by grant number 426602 from the Simons Foundation. Jakub Bielawski and Grzegorz Kosiorowski acknowledge support from a subsidy granted to Cracow University of Economics. Thiparat Chotibut acknowledges a fruitful discussion with Tanapat Deesuwan, and was partially supported by grants for development of new faculty staff, Ratchadaphiseksomphot endownment fund, and Sci-Super VI fund, Chulalongkorn University.

References

  • [1] R. L. Adler, A. G. Konheim, and M. H. McAndrew. Topological entropy. Transactions of American Mathematical Society, 114:309–319, 1965.
  • [2] S. Arora, E. Hazan, and S. Kale. The multiplicative weights update method: a meta-algorithm and applications. Theory of Computing, 8(1):121–164, 2012.
  • [3] J. P. Bailey, G. Gidel, and G. Piliouras. Finite regret and cycles with fixed step-size via alternating gradient descent-ascent. CoRR, abs/1907.04392, 2019.
  • [4] J. P. Bailey and G. Piliouras. Fast and furious learning in zero-sum games: Vanishing regret with non-vanishing step sizes. In Advances in Neural Information Processing Systems, volume 32, pages 12977–12987, 2019.
  • [5] P. Berenbrink, M. Hoefer, and T. Sauerwald. Distributed selfish load balancing on networks. In ACM Transactions on Algorithms (TALG), 2014.
  • [6] L. Block and W. A. Coppel. Dynamics in one dimension, volume 513 of Lecture Notes in Mathematics. Springer, Berlin New York, 2006.
  • [7] A. Blum, E. Even-Dar, and K. Ligett. Routing without regret: On convergence to nash equilibria of regret-minimizing algorithms in routing games. In Proceedings of the twenty-fifth annual ACM symposium on Principles of distributed computing, pages 45–52. ACM, 2006.
  • [8] I. M. Bomze, P. Mertikopoulos, W. Schachinger, and M. Staudigl. Hessian barrier algorithms for linearly constrained optimization problems. SIAM Journal on Optimization, 29(3):2100–2127, 2019.
  • [9] C. F. Camerer. Behavioral game theory: Experiments in strategic interaction. Princeton University Press, 2011.
  • [10] T. Chotibut, F. Falniowski, M. Misiurewicz, and G. Piliouras. Family of chaotic maps from game theory. Dynamical Systems, 2020. https://doi.org/10.1080/14689367.2020.1795624.
  • [11] T. Chotibut, F. Falniowski, M. Misiurewicz, and G. Piliouras. The route to chaos in routing games: When is price of anarchy too optimistic? Advances in Neural Information Processing Systems, 33, 2020.
  • [12] P. Coucheney, B. Gaujal, and P. Mertikopoulos. Penalty-regulated dynamics and robust learning procedures in games. Mathematics of Operations Research, 40(3):611–633, 2015.
  • [13] I. Csiszár. Axiomatic characterizations of information measures. Entropy, 10(3):261–273, 2008.
  • [14] E. Even-Dar and Y. Mansour. Fast convergence of selfish rerouting. In Proceedings of the Sixteenth Annual ACM-SIAM Symposium on Discrete Algorithms, SODA ’05, pages 772–781, Philadelphia, PA, USA, 2005. Society for Industrial and Applied Mathematics.
  • [15] S. Fischer, H. Räcke, and B. Vöcking. Fast convergence to wardrop equilibria by adaptive sampling methods. In Proceedings of the Thirty-eighth Annual ACM Symposium on Theory of Computing, STOC ’06, pages 653–662, New York, NY, USA, 2006. ACM.
  • [16] D. Fotakis, A. C. Kaporis, and P. G. Spirakis. Atomic congestion games: Fast, myopic and concurrent. In B. Monien and U.-P. Schroeder, editors, Algorithmic Game Theory, volume 4997 of Lecture Notes in Computer Science, pages 121–132. Springer Berlin Heidelberg, 2008.
  • [17] E. Glasner and B. Weiss. Sensitive dependence on initial conditions. Nonlinearity, 6(6):1067–1085, 1993.
  • [18] I. Goodfellow, J. Pouget-Abadie, M. Mirza, B. Xu, D. Warde-Farley, S. Ozair, A. Courville, and Y. Bengio. Generative adversarial nets. In Advances in neural information processing systems, pages 2672–2680, 2014.
  • [19] P. D. Grünwald and A. P. Dawid. Game theory, maximum entropy, minimum discrepancy and robust bayesian decision theory. The Annals of Statistics, 32(4):1367–1433, 2004.
  • [20] M. Harper. Escort evolutionary game theory. Physica D: Nonlinear Phenomena, 240(18):1411–1415, 2011.
  • [21] J. Havrda and F. Charvát. Quantification method of classification processes. concept of structural aa-entropy. Kybernetika, 3(1):30–35, 1967.
  • [22] E. Hazan et al. Introduction to online convex optimization. Foundations and Trends® in Optimization, 2(3-4):157–325, 2016.
  • [23] T.-H. Ho and C. Camerer. Experience-weighted attraction learning in coordination games: Probability rules, heterogeneity, and time-variation. Journal of Mathematical Psychology, 42:305–326, 1998.
  • [24] T.-H. Ho and C. Camerer. Experience-weighted attraction learning in normal form games. Econometrica, 67:827–874, 1999.
  • [25] T.-H. Ho, C. F. Camerer, and J.-K. Chong. Self-tuning experience weighted attraction learning in games. Journal of Economic Theory, 133:177–198, 2007.
  • [26] G. P. Karev and E. V. Koonin. Parabolic replicator dynamics and the principle of minimum Tsallis information gain. Biology Direct, 8:19 – 19, 2013.
  • [27] A. Kianercy and A. Galstyan. Dynamics of Boltzmann QQ learning in two-player two-action games. Phys. Rev. E, 85:041145, Apr 2012.
  • [28] R. Kleinberg, G. Piliouras, and É. Tardos. Multiplicative updates outperform generic no-regret learning in congestion games. In ACM Symposium on Theory of Computing (STOC), 2009.
  • [29] R. Kleinberg, G. Piliouras, and É. Tardos. Load balancing without regret in the bulletin board model. Distributed Computing, 24(1):21–29, 2011.
  • [30] E. Koutsoupias and C. Papadimitriou. Worst-case equilibria. In C. Meinel and S. Tison, editors, STACS 99, pages 404–413, Berlin, Heidelberg, 1999. Springer Berlin Heidelberg.
  • [31] T. Y. Li, M. Misiurewicz, G. Pianigiani, and J. A. Yorke. Odd chaos. Physics Letters A, 87(6):271–273, 1982.
  • [32] T. Y. Li and J. A. Yorke. Period three implies chaos. The American Mathematical Monthly, 82:985–992, 1975.
  • [33] T. Liang and J. Stokes. Interaction matters: A note on non-asymptotic local convergence of generative adversarial networks. In The 22nd International Conference on Artificial Intelligence and Statistics, pages 907–915. PMLR, 2019.
  • [34] P. Mertikopoulos and W. H. Sandholm. Learning in games via reinforcement and regularization. Mathematics of Operations Research, 41(4):1297–1324, 2016.
  • [35] P. Mertikopoulos and W. H. Sandholm. Riemannian game dynamics. Journal of Economic Theory, 177:315–364, 2018.
  • [36] P. Mertikopoulos and Z. Zhou. Learning in games with continuous action sets and unknown payoff functions. Mathematical Programming, 173(1-2):465–507, 2019.
  • [37] L. Mescheder, A. Geiger, and S. Nowozin. Which training methods for gans do actually converge? arXiv preprint arXiv:1801.04406, 2018.
  • [38] M. Misiurewicz. Horseshoes for mapping of the interval. Bull. Acad. Polon. Sci. Sér. Sci., 27:167–169, 1979.
  • [39] M. Misiurewicz and W. Szlenk. Entropy of piecewise monotone mappings. Studia Mathematica, 67(1):45–63, 1980.
  • [40] D. Monderer and L. S. Shapley. Fictitious play property for games with identical interests. Journal of Economic Theory, 68(1):258–265, 1996.
  • [41] V. Nagarajan and J. Z. Kolter. Gradient descent gan optimization is locally stable. In Advances in neural information processing systems, pages 5585–5595, 2017.
  • [42] G. Palaiopanos, I. Panageas, and G. Piliouras. Multiplicative weights update with constant step-size in congestion games: Convergence, limit cycles and chaos. In Advances in Neural Information Processing Systems, pages 5872–5882, 2017.
  • [43] I. Panageas, G. Piliouras, and X. Wang. Multiplicative weights updates as a distributed constrained optimization algorithm: Convergence to second-order stationary points almost always. In International Conference on Machine Learning, pages 4961–4969. PMLR, 2019.
  • [44] A. Rényi. On measures of entropy and information. Proceedings of the Fourth Berkeley Symposium on Mathematical Statistics and Probability, Volume 1: Contributions to the Theory of Statistics, 1961.
  • [45] R. Rosenthal. A class of games possessing pure-strategy Nash equilibria. International Journal of Game Theory, 2(1):65–67, 1973.
  • [46] Y. Sato, E. Akiyama, and J. P. Crutchfield. Stability and diversity in collective adaptation. Physica D: Nonlinear Phenomena, 210(1):21 – 57, 2005.
  • [47] Y. Sato and J. P. Crutchfield. Coupled replicator equations for the dynamics of learning in multiagent systems. Phys. Rev. E, 67:015206, Jan 2003.
  • [48] S. Shalev-Shwartz. Online learning and online covex optimization. Foundations and Trends in Machine Learning, 4(2):107–194, 2012.
  • [49] A. N. Sharkovsky. Coexistence of the cycles of a continuous mapping of the line into itself. Ukrain. Math. Zh., 16:61–71, 1964.
  • [50] W. Słomczyński, J. Kwapień, and K. Życzkowski. Entropy computing via integration over fractal measures. Chaos: An Interdisciplinary Journal of Nonlinear Science, 10(1):180–188, 2000.
  • [51] C. Tsallis. Possible generalization of Boltzmann-Gibbs statistics. Journal of Statistical Physics, 52(1-2):479–487, 1988.
  • [52] B. Weiss. Single orbit dynamics, volume 95 of CBMS Regional Conference Series in Mathematics. American Mathematical Society, Providence, RI, 2000.
  • [53] Y. Yaz, C.-S. Foo, S. Winkler, K.-H. Yap, G. Piliouras, V. Chandrasekhar, et al. The unusual effectiveness of averaging in gan training. In International Conference on Learning Representations, 2018.
  • [54] Q. Zhao. Multi-Armed Bandits: Theory and Applications to Online Learning in Networks. Morgan and Claypool Publishers, 2019.

Appendices

Appendix A Generalized entropies as regularizers

We present information measures which are often used as regularizers.

Example A.1 (Shannon entropy).

Let R(x,y)=HS(x,y)R(x,y)=-H_{S}(x,y), where

HS(x,y)=xlogxylogy.H_{S}(x,y)=-x\log x-y\log y.

Then HS(x,1x)H_{S}(x,1-x) is the Shannon entropy of a probability distribution (x,1x)(x,1-x) and

r(x)=R(x,1x)=HS(x,1x)=xlogx+(1x)log(1x).r(x)=R(x,1-x)=-H_{S}(x,1-x)=x\log x+(1-x)\log(1-x).

From

r(x)=logx1xr^{\prime}(x)=\log\frac{x}{1-x}

we observe that r𝐒𝐒𝐂r\in\operatorname{\mathbf{SSC}}.555By substituting the (negative) Shannon Entropy as RR into (4) we obtain the Multiplicative Weights Update algorithm.

Example A.2 (Arimoto entropies).

We consider the class of Arimoto entropies [13], that is functions defined as

Hη(x,y)=η(x)+η(y),H_{\eta}(x,y)=\eta(x)+\eta(y),

where η𝒞2((0,1))\eta\in\mathcal{C}^{2}((0,1)) is a concave function.666In the decision theory Arimoto entropies correspond to separable Bregman scores [19].

We define

R(x,y)=Hη(x,y).R(x,y)=-H_{\eta}(x,y).

Then, by its definition, R𝒞2((0,1)2)R\in\mathcal{C}^{2}((0,1)^{2}) and R(y,x)=R(x,y)R(y,x)=R(x,y). Moreover,

r(x)=R(x,1x)=Hη(x,1x)=η(x)η(1x),r(x)=R(x,1-x)=-H_{\eta}(x,1-x)=-\eta(x)-\eta(1-x),
r(x)=η(x)+η(1x) and r′′(x)=η′′(x)η′′(1x).r^{\prime}(x)=-\eta^{\prime}(x)+\eta^{\prime}(1-x)\text{ and }r^{\prime\prime}(x)=-\eta^{\prime\prime}(x)-\eta^{\prime\prime}(1-x).

Thus, rr is convex and the limit limx1η(x)\lim_{x\to 1^{-}}\eta^{\prime}(x) is finite. Therefore, the condition for R=Hη𝐒𝐒𝐂R=-H_{\eta}\in\operatorname{\mathbf{SSC}} is steepness of η\eta at zero:

(15) limx0+η(x)=.\lim_{x\to 0^{+}}\eta^{\prime}(x)=\infty.

Hence, R=Hη𝐒𝐒𝐂R=-H_{\eta}\in\operatorname{\mathbf{SSC}} if and only if η\eta satisfies (15).

Several well-known regularizers are given by (negative) Arimoto entropies satisfying (15). For instance, the Shannon entropy from Example A.1 is an Arimoto entropy for η(x)=xlogx\eta(x)=-x\log x, as well as log-barrier regularizer obtained from η(x)=logx\eta(x)=\log x. Another widely used (especially in statistical physics) example of Arimoto entropy is the Havrda-Charvát-Tsallis entropy.777This entropy (called also entropy of degree qq) was first introduced by Havrda and Charvát [21] and used to bound probability of error for testing multiple hypotheses. In statistical physics it is known as Tsallis entropy, referring to [51].

Example A.3 (Havrda-Charvát-Tsallis entropies).

The Havrda-Charvát-Tsallis entropy for q(0,)q\in(0,\infty) is defined as

(16) Hq(x,y)={11q(xq+yq1)for q1HS(x,y)for q=1.H_{q}(x,y)=\begin{cases}\frac{1}{1-q}(x^{q}+y^{q}-1)&\text{for }q\neq 1\\ H_{S}(x,y)&\text{for }q=1\end{cases}.

HqH_{q} is an Arimoto entropy for η(x)=11q(xq12)\eta(x)=\frac{1}{1-q}\left(x^{q}-\frac{1}{2}\right), satisfying (15) for 0<q<10<q<1. If R(x,y)=Hq(x,y)R(x,y)=-H_{q}(x,y) then

r(x)=R(x,1x)=1q1(xq+(1x)q1)andr(x)=qq1(xq1(1x)q1),r(x)=R(x,1-x)=\frac{1}{q-1}(x^{q}+(1-x)^{q}-1)\;\text{and}\;r^{\prime}(x)=\frac{q}{q-1}\left(x^{q-1}-(1-x)^{q-1}\right),

and r𝐒𝐒𝐂r\in\operatorname{\mathbf{SSC}} for q(0,1]q\in(0,1].

For q>1q>1 the Havrda-Charvát-Tsallis entropy does not satisfy (15) and, consequently, the regularizer RR emerging from the Havrda-Charvát-Tsallis entropy does not belong to 𝐒𝐒𝐂\operatorname{\mathbf{SSC}}. Standard non-example is Euclidean norm, which we get from (16) when q=2q=2. Then

r(x)=R(x,1x)=H2(x,1x)=x2+(1x)21r(x)=R(x,1-x)=-H_{2}(x,1-x)=x^{2}+(1-x)^{2}-1

and as limx0+r(x)=2\lim_{x\to 0^{+}}r^{\prime}(x)=-2, RR doesn’t belong to 𝐒𝐒𝐂\operatorname{\mathbf{SSC}}.

Evidently there exist functions which are not Arimoto entropies but also generate regularizers that belong to 𝐒𝐒𝐂\operatorname{\mathbf{SSC}}, one of them being the Rényi entropy of order q<1q<1.

Example A.4 (Rényi entropies).

The Shannon entropy represents an expected mean of individual informations of the form Ik=logpkI_{k}=-\log p_{k}. Rényi [44] introduced alternative information measures, namely generalized means g1(pkg(Ik))g^{-1}(\sum p_{k}g(I_{k})), where gg is a continuous, strictly monotone function.Then, the Rényi entropy of order q1q\neq 1 correspond to g(x)=exp((1q)x)g(x)=\exp((1-q)x), namely:

HqR(x,y)={11qlog(xq+yq),for q1HS(x,y),for q=1.H_{q}^{R}(x,y)=\begin{cases}\frac{1}{1-q}\log\left(x^{q}+y^{q}\right),&\text{for }q\neq 1\\ H_{S}(x,y),&\text{for }q=1\end{cases}.

As the variables xx and yy are not separable, this is not an Arimoto entropy. However, for R(x,y)=HqR(x,y)R(x,y)=-H_{q}^{R}(x,y), R𝒞2((0,1)2)R\in\mathcal{C}^{2}((0,1)^{2}) and R(y,x)=R(x,y)R(y,x)=R(x,y). Moreover,

r(x)=R(x,1x)=HqR(x,1x)=1q1log(xq+(1x)q)r(x)=R(x,1-x)=-H_{q}^{R}(x,1-x)=\frac{1}{q-1}\log\left(x^{q}+(1-x)^{q}\right)

and

r(x)=qq1xq1(1x)q1xq+(1x)q.r^{\prime}(x)=\frac{q}{q-1}\cdot\frac{x^{q-1}-(1-x)^{q-1}}{x^{q}+(1-x)^{q}}.

Thus, for q(0,1)q\in(0,1) we know that r′′(x)>0r^{\prime\prime}(x)>0 on (0,1)(0,1) and limx0+r(x)=\lim_{x\to 0^{+}}r^{\prime}(x)=-\infty. Because H1R=HSH_{1}^{R}=H_{S} we infer that R𝐒𝐒𝐂R\in\operatorname{\mathbf{SSC}} for q(0,1]q\in(0,1].

Appendix B Regularity of log-barrier dynamics

To understand better the phenomenon discussed in Section 8, let us investigate regularity of fa,bf_{a,b}. Nice properties of interval maps are guaranteed by the negative Schwarzian derivative. Let us recall that the Schwarzian derivative of ff is given by the formula

Sf=f′′′f32(f′′f)2.Sf=\frac{f^{\prime\prime\prime}}{f^{\prime}}-\frac{3}{2}\left(\frac{f^{\prime\prime}}{f^{\prime}}\right)^{2}.

A “metatheorem” states that almost all natural noninvertible interval maps have negative Schwarzian derivative. Note that, by Lemma 3.2.ii, if aΨ(b)a\leq-\Psi^{\prime}(b) then fa,bf_{a,b} is a homeomorphism, so we should not expect negative Schwarzian derivative for that case. For maps with negative Schwarzian derivative each attracting or neutral periodic orbit has a critical point in its immediate basin of attraction. Thus, if we show that the Schwarzian derivative is negative, then we will know that all periodic orbits can be find by studying behavior of critical points of fa,bf_{a,b}. Therefore, we want to show that Sfa,b<0Sf_{a,b}<0 for sufficiently large aa for fa,bf_{a,b} determined by log-barrier regularizer.

In general, computation of Schwarzian derivative may be very complicated. However, there is a useful formula

(17) S(hf)=(f)2((Sh)f)+Sf.S(h\circ f)=(f^{\prime})^{2}\left((Sh)\circ f\right)+Sf.

The function fa,bf_{a,b} is given by (7). Consider

g(x):=(Ψfa,b)(x)=Ψ(x)+a(xb).g(x):=(\Psi\circ f_{a,b})(x)=\Psi(x)+a(x-b).

By (17) we have that

Sg=(fa,b)2((SΨ)fa,b)+Sfa,b.Sg=(f_{a,b}^{\prime})^{2}\left((S\Psi)\circ f_{a,b}\right)+Sf_{a,b}.

At the same time

Sg(x)=S(Ψ(x)+a(xb)).Sg(x)=S(\Psi(x)+a(x-b)).

Therefore,

(18) (fa,b(x))2((SΨ)fa,b(x))+Sfa,b(x)=S(Ψ(x)+a(xb)).(f_{a,b}^{\prime}(x))^{2}\left((S\Psi)\circ f_{a,b}(x)\right)+Sf_{a,b}(x)=S(\Psi(x)+a(x-b)).

Direct computations yield

SΨ(x)=6(x2+(1x)2)2>0S\Psi(x)=\frac{6}{\left(x^{2}+(1-x)^{2}\right)^{2}}>0

and

S(Ψ(x)+a(xb))=6[1a(x4+(1x)4)][x2+(1x)2ax2(1x)2]2.S(\Psi(x)+a(x-b))=\frac{6\left[1-a(x^{4}+(1-x)^{4})\right]}{\left[x^{2}+(1-x)^{2}-ax^{2}(1-x)^{2}\right]^{2}}.

Observe that x4+(1x)418x^{4}+(1-x)^{4}\geqslant\frac{1}{8} for all x[0,1]x\in[0,1]. Thus 1a(x4+(1x)4)1a81-a(x^{4}+(1-x)^{4})\leqslant 1-\frac{a}{8}, and

(19) S(Ψ(x)+a(xb))<0for a>8.S(\Psi(x)+a(x-b))<0\quad\text{for }a>8.

Therefore, Sfa,b<0Sf_{a,b}<0 for a>8a>8. Moreover,

maxb[0,1]Ψ(b)=Ψ(1/2)=8.\max_{b\in[0,1]}\Psi^{\prime}(b)=\Psi^{\prime}(1/2)=-8.

Thus,

Sfa,b(x)<0 for all a>Ψ(b)8.Sf_{a,b}(x)<0\text{ for all }a>-\Psi^{\prime}(b)\geq 8.