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

A sensitivity analysis of the PoA in non-atomic congestion games

Zijun Wu
Institute for Applied Optimization
School of Artificial Intelligence and Big Data
Hefei University
Jinxiu 99, 230601, Jingkai District, Hefei, Anhui, China
[email protected]
Rolf H. Möhring
Kombinatorische Optimierung und Graphenalgorithmen (COGA)
Fakultät II-Mathematik und Naturwissenschaften
Institut für Mathematik, Sekr. MA 5-1
Technische Universität Berlin
Straße des 17. Juni 136, Berlin, Germany, 10623
[email protected]
Abstract

The price of anarchy(PoA) is a standard measure to quantify the inefficiency of equilibria in non-atomic congestion games. Most publications have focused on worst-case bounds for the PoA, only few have analyzed the sensitivity of the PoA against changes of the demands or cost functions, although that is crucial for empirical computation of the PoA. We analyze the sensitivity of the PoA w.r.t. simultaneous changes of demands and cost functions. The key to this analysis is a metric for the distance between two games that defines a topological metric space consisting of all games with the same combinatorial structure. The PoA is then a locally pointwise Hölder continuous function of the demands and cost functions, and we analyze the Hölder exponent for different classes of cost functions. We also apply our approach to the convergence analysis of the PoA when the total demand tends to zero or infinity. Our results further develop the recent seminal work by Englert etal., Takalloo and Kwon, and Cominetti et al., who have considered the sensitivity of the PoA w.r.t. changes of the demands under special conditions.

1 Introduction

The Price of Anarchy (PoA) (Koutsoupias and Papadimitriou [1999] and Papadimitriou [2001]) is a classic notion in algorithmic game theory and has been intensively studied in the last two decades, see, e.g., the book by Nisan et al. [2007] for an overview. Much of this work has been devoted to worst-case upper bounds of the PoA in congestion games for cost functions of different types, starting with the pioneering paper of Roughgarden and Tardos [2002].

Much less attention has been paid to the evolution of the PoA as a “function” of the demands and the cost functions, although this is quite important for traffic networks. In fact, the real demands are usually hard to measure accurately since they may fluctuate within a certain range. Also the actual latency on a street in a traffic network is almost impossible to obtain, and usually modeled by an idealized flow-dependent cost function that is estimated empirically from real traffic data and may also include tolls. Thus modeling errors will inevitably occur. This raises the natural question if and how much such modeling errors will influence the PoA, in particular, if the PoA may differ largely under small modeling or measuring errors of the demands and cost functions. This is crucial for applications and asks for a sensitivity analysis of the PoA in a traffic network w.r.t. small changes of the demands and the cost functions of the network.

First results in this direction have been obtained by Englert et al. [2010], Takalloo and Kwon [2020], and Cominetti et al. [2020].

Englert et al. [2010] considered a traffic network with a single origin-destination (O/D) pair and polynomial cost functions of degree at most β.\beta. They view the PoA in the network as a function of the demand of that O/D pair when the cost functions stay unchanged. They showed that the increase of this PoA function is bounded by a factor of (1+ϵ)β(1+\epsilon)^{\beta} from above when the demand increases by a factor of 1+ϵ1+\epsilon.

Takalloo and Kwon [2020] have generalized this result to traffic networks with multiple O/D pairs and polynomial cost functions of degree at most β\beta. They obtained the same upper bound for the increase of the PoA when the demands of all O/D pairs increase synchronously by the same factor 1+ϵ1+\epsilon and the cost functions stay unchanged. Moreover, they showed that the increase of the PoA is bounded from below by a factor 1(1+ϵ)β\frac{1}{(1+\epsilon)^{\beta}} under the same conditions.

Similar to Englert et al. [2010], Cominetti et al. [2020] considered also a traffic network with a single O/D pair, and viewed the PoA as a function of the demand of that O/D pair when the cost functions are not varying. For cost functions with strictly positive derivatives, they showed that the resulting PoA function is differentiable at each demand level of the O/D pair where all the optimal paths carry a strictly positive flow. For affine linear cost functions, they showed further that the equilibrium cost is piece-wise linear and differentiable except at so-called \mathcal{E}-breakpoints, whose number is finite, though possibly exponentially large in the size of the network. They then showed that, in any interval between two consecutive \mathcal{E}-breakpoints, the PoA function is differentiable, and either monotone or unimodal with a unique minimum on the interior of that interval. So the maximum of the PoA function is attained at an \mathcal{E}-breakpoint. They also presented several examples showing how these properties fail for general cost functions.

We will come back to their results in more detail in Section 4.

These results undoubtedly indicate that the change of the PoA in a traffic network is within an acceptable and predictable range w.r.t. small changes of the demands when the cost functions are fixed and have certain good properties, and the changes of the demands fulfill certain regularity conditions. They are thus seminal first results for a sensitivity analysis of the PoA in traffic networks, but are still restricted to special cases. In particular, a sensitivity analysis of the PoA w.r.t. simultaneous changes of demands and cost functions is still missing.

1.1 Our contribution

This paper continues the studies of Englert et al. [2010], Takalloo and Kwon [2020] and Cominetti et al. [2020]. We consider a general traffic network that have multiple O/D pairs and general cost functions that may include tolls. We view the PoA as a real-valued map of the demands and the cost functions of that network, and then analyze the sensitivity of the resulting map when both the demands and the cost functions may change.

To that purpose, we need first a well defined measure for simultaneous changes of the demands and the cost functions in a traffic network with multiple O/D pairs.

We thus fix in our analysis an arbitrary directed network G=(V,A)G=(V,A) and an arbitrary finite set 𝒦\mathcal{K} of O/D pairs in that network, and consider only the cost functions τa()\tau_{a}(\cdot) of arcs aAa\in A and the demands dkd_{k} of O/D pairs k𝒦k\in\mathcal{K} as “variables”. Each pair Γ=(τ,d)\Gamma=(\tau,d) consisting of a cost function vector τ=(τa)aA\tau=(\tau_{a})_{a\in A} and a demand vector d=(dk)k𝒦d=(d_{k})_{k\in\mathcal{K}} then represents a “state” of the network and corresponds to an instance of the traffic game (Roughgarden and Tardos [2002]) defined on the network. We then view the collection of all these states as a game space, and the PoA in the network as a real-valued map ρ()\rho(\cdot) from this space to the unbounded interval [1,).[1,\infty). Then the value ρ(Γ)\rho(\Gamma) at a point Γ=(τ,d)\Gamma=(\tau,d) of the game space is the resulting PoA when the network is at state Γ,\Gamma, i.e., the resulting PoA when the network has the cost function vector τ\tau and the demand vector d.d.

By adapting the LL^{\infty}-norms of functions and vectors, we define a binary operator Dist(,),\text{Dist}(\cdot,\cdot), see (3.3), on the game space by putting

Dist(Γ,Γ)=max{dd,maxaA,x[0,min{T(d),T(d)}]|τa(x)σa(x)|,τ(T(d))σ(T(d))}\text{Dist}(\Gamma,\Gamma^{\prime})\!=\!\max\left\{|\!|d\!-\!d^{\prime}|\!|_{\infty},\ \max_{a\in A,x\in[0,\min\{T(d),T(d^{\prime})\}]}|\tau_{a}(x)\!-\!\sigma_{a}(x)|,\ |\!|\tau(T(d))\!-\!\sigma(T(d^{\prime}))|\!|_{\infty}\right\}

for two arbitrary points Γ=(τ,d)\Gamma=(\tau,d) and Γ=(σ,d)\Gamma^{\prime}=(\sigma,d^{\prime}) of the game space, where T(d)=k𝒦dkT(d)=\sum_{k\in\mathcal{K}}d_{k} and T(d)=k𝒦dkT(d^{\prime})=\sum_{k\in\mathcal{K}}d^{\prime}_{k} are the total demands of the demand vectors dd and dd^{\prime} respectively, τ(T(d))=(τa(T(d)))aA\tau(T(d))=(\tau_{a}(T(d)))_{a\in A} and σ(T(d))=(σa(T(d)))aA.\sigma(T(d^{\prime}))=(\sigma_{a}(T(d^{\prime})))_{a\in A}.

This binary operator Dist(,)\text{Dist}(\cdot,\cdot) is actually a metric on the game space w.r.t. an equivalence relation defined in (3.1), see Lemma 4. The game space then becomes a metric space, and a simultaneous change of the cost functions and the demands is then quantified by the “distance” ΓΓ:=Dist(Γ,Γ)|\!|\Gamma-\Gamma^{\prime}|\!|:=\text{Dist}(\Gamma,\Gamma^{\prime}) between two points Γ\Gamma and Γ\Gamma^{\prime} of the game space.

With this metric, our sensitivity analysis of the PoA in the network then transforms to an analysis of the pointwise Hölder continuity of the map ρ()\rho(\cdot) on the game space, aiming at finding for each point Γ=(τ,d)\Gamma=(\tau,d) a Hölder exponent γΓ>0\gamma_{\Gamma}>0 s.t. |ρ(Γ)ρ(Γ)|<ϰΓΓΓγΓ|\rho(\Gamma)-\rho(\Gamma^{\prime})|<\varkappa_{\Gamma}\cdot|\!|\Gamma-\Gamma^{\prime}|\!|^{\gamma_{\Gamma}} when ΓΓ<ϵΓ|\!|\Gamma-\Gamma^{\prime}|\!|<\epsilon_{\Gamma} for two positive constants ϰΓ,ϵΓ>0\varkappa_{\Gamma},\epsilon_{\Gamma}>0 depending only on Γ.\Gamma. Trivially, the larger the Hölder exponent γΓ\gamma_{\Gamma} at a point Γ=(τ,d)\Gamma=(\tau,d), the less sensitive is the PoA at the network state Γ\Gamma w.r.t. a small change ΓΓ|\!|\Gamma-\Gamma^{\prime}|\!| of the state Γ\Gamma.

As our first result, we show that both the equilibrium cost and the socially optimal cost are continuous maps (w.r.t. the metric) on the whole game space. This implies directly that the map ρ()\rho(\cdot) is continuous on the whole game space, see Theorem 6. Hence, the PoA changes only slightly when the changes of the cost functions and the demands are very small in terms of the metric.

We then show that ρ()\rho(\cdot) is not uniformly Hölder continuous on the whole game space, i.e., there are no constants γ,ϰ>0\gamma,\varkappa>0 such that

|ρ(Γ)ρ(Γ)|ϰΓΓγΓ,Γ𝒢(G,𝒦,𝒮),|\rho(\Gamma)-\rho(\Gamma^{\prime})|{\leq}\varkappa\cdot|\!|\Gamma-\Gamma^{\prime}|\!|^{\gamma}\quad\forall\Gamma,\Gamma^{\prime}\in\mathcal{G}(G,\mathcal{K},\mathcal{S}),

see Theorem 8a). Moreover, we show that no point Γ\Gamma of the game space has the whole game space as its Hölder neighborhood, see Theorem 8b). Here, a Hölder neighborhood of a game Γ\Gamma refers to an open subset UΓU_{\Gamma} of the game space with ΓUΓ,\Gamma\in U_{\Gamma}, for which there are constants γΓ,ϰΓ>0\gamma_{\Gamma},\varkappa_{\Gamma}>0 depending only on Γ\Gamma s.t. |ρ(Γ)ρ(Γ)|ϰΓΓΓγΓ|\rho(\Gamma)-\rho(\Gamma^{\prime})|{\leq}\varkappa_{\Gamma}\cdot|\!|\Gamma-\Gamma^{\prime}|\!|^{\gamma_{\Gamma}} for each ΓUΓ.\Gamma^{\prime}\in U_{\Gamma}.

Hence, ρ()\rho(\cdot) may only be pointwise and locally Hölder continuous on the game space. In particular, Theorem 8a) implies that ρ()\rho(\cdot) is not Lipschitz continuous on the whole game space.

Along with Theorem 8, our first Hölder continuity result shows that ρ()\rho(\cdot) is pointwise Hölder continuous with Hölder exponent 12\frac{1}{2} at each point Γ=(τ,d)\Gamma=(\tau,d) where the cost functions τa()\tau_{a}(\cdot) are Lipschitz continuous on the interval [0,T(d)],[0,T(d)], see Theorem 9. Hence, the change of the PoA is bounded from above by ϰΓΓΓ\varkappa_{\Gamma}\cdot\sqrt{|\!|\Gamma-\Gamma^{\prime}|\!|} for a Hölder constant ϰΓ>0\varkappa_{\Gamma}>0 when the network undergoes only a small change ΓΓ|\!|\Gamma-\Gamma^{\prime}|\!| at a state Γ=(τ,d)\Gamma=(\tau,d) with Lipschitz continuous cost functions τa()\tau_{a}(\cdot) on [0,T(d)].[0,T(d)]. Since the interval [0,T(d)][0,T(d)] is compact, this result applies obviously to all continuously differentiable cost functions. Because of Theorem 8b), however, it may not apply when the network undergoes a large change of the cost functions and/or the demands in terms of the metric.

To obtain this result, we show first that |ρ(Γ)ρ(Γ)|ϰ1,ΓΓΓ|\rho(\Gamma)-\rho(\Gamma^{\prime})|{\leq}\varkappa_{1,\Gamma}\cdot\sqrt{|\!|\Gamma-\Gamma^{\prime}|\!|} for a constant ϰ1,Γ>0\varkappa_{1,\Gamma}>0 when Γ=(τ,d)\Gamma=(\tau,d) and Γ=(σ,d)\Gamma^{\prime}=(\sigma,d) have the same demand vector d,d, the change ΓΓ|\!|\Gamma-\Gamma^{\prime}|\!| is small and the cost functions τa()\tau_{a}(\cdot) of Γ\Gamma are Lipschitz continuous on [0,T(d)],[0,T(d)], see Lemma 10. We then show that |ρ(Γ)ρ(Γ)|ϰ2,ΓΓΓ|\rho(\Gamma)-\rho(\Gamma^{\prime})|{\leq}\varkappa_{2,\Gamma}\cdot\sqrt{|\!|\Gamma-\Gamma^{\prime}|\!|} for a constant ϰ2,Γ>0\varkappa_{2,\Gamma}>0 when Γ=(τ,d)\Gamma=(\tau,d) and Γ=(τ,d)\Gamma^{\prime}=(\tau,d^{\prime}) have the same cost functions τa()\tau_{a}(\cdot), the change ΓΓ|\!|\Gamma-\Gamma^{\prime}|\!| is small and the cost functions τa()\tau_{a}(\cdot) are Lipschitz continuous on [0,T(d)],[0,T(d)], see Lemma 11.

Lemma 10 and Lemma 11 then imply Theorem 9. In particular, Lemma 11 generalizes the sensitivity results of Englert et al. [2010] and Takalloo and Kwon [2020] by considering more general cost functions and removing the requirement that the demands of different O/D pairs increase synchronously by the same factor. However, this comes at the cost of a weaker Hölder exponent 12\frac{1}{2} than that of Englert et al. [2010] and Takalloo and Kwon [2020], who obtain a Hölder exponent of 11.

The above Hölder continuity results build essentially upon Lemma 1c) in Section 2.3, which shows that the total cost of an ϵ\epsilon-approximate Wardrop equilibrium (Wardrop [1952]) deviates by at most O(ϵ)O(\sqrt{\epsilon}) from that of a precise Wardrop equilibrium. This is already a tight upper bound on the cost deviation for arbitrary Lipschitz continuous cost functions on [0,T(d)],[0,T(d)], see Example 2, and we are thus presently unable to improve the Hölder exponent 12\frac{1}{2} in Theorem 9, Lemma 10 and Lemma 11, see Remark 4.2. Nevertheless, a stronger Hölder exponent is still possible when the cost functions have special properties similar to those in the work of Englert et al. [2010], Cominetti et al. [2020], and Takalloo and Kwon [2020].

When the cost functions τa()\tau_{a}(\cdot) of a point Γ=(τ,d)\Gamma=(\tau,d) are constants or have strictly positive derivatives on the interval [0,T(d)],[0,T(d)], we obtain the stronger result that ρ()\rho(\cdot) is pointwise Hölder continuous with Hölder exponent 11 at the point Γ,\Gamma, see Theorem 12. Hence, the resulting change of the PoA is bounded from above by ϰΓΓΓ\varkappa_{\Gamma}\cdot|\!|\Gamma-\Gamma^{\prime}|\!| for a constant ϰΓ>0\varkappa_{\Gamma}>0 when the network undergoes a small change ΓΓ|\!|\Gamma-\Gamma^{\prime}|\!| at such a state Γ\Gamma. Again by Theorem 8, this may not hold when the network undergoes a large change of the cost functions and/or demands in terms of the metric.

Finally, we demonstrate that our Hölder continuity results also help to analyze the convergence rate of the PoA in traffic networks, which is an emerging research topic initiated recently by Colini-Baldeschi et al. [2017, 2020], see Section 1.2 or Section 5.1 for an overview of the related work on this new topic.

For T(d)0T(d)\to 0, Colini-Baldeschi et al. [2017, 2020] have obtained the first convergence result stating that the PoA converges to 11 at a rate of O(T(d))O(T(d)) when the cost functions are of the form τa(x)=nξa,nxn\tau_{a}(x)=\sum_{n\in\mathbb{N}}\xi_{a,n}\cdot x^{n} for each aAa\in A and each x[0,)x\in[0,\infty), and the demands of all O/D pairs follow a specific decreasing pattern, i.e., all of them decrease to 0 at the same rate of Θ(T(d)),\Theta(T(d)), see Section 1.2 or Section 5.1. With Theorem 12a), we show a stronger result that the PoA converges to 11 at a rate of O(T(d))O(T(d)) as T(d)0T(d)\to 0 regardless of the decreasing pattern of the demands when the cost functions are strictly positive and Lipschitz continuous within a small neighborhood around the origin 0, see Corollary 5.1.

When T(d),T(d)\to\infty, it has been shown by Colini-Baldeschi et al. [2017, 2020] and Wu et al. [2021] that the PoA converges to 11 as T(d)T(d)\to\infty for arbitrary regularly varying cost functions. For the special case of arbitrary polynomials as cost functions, Colini-Baldeschi et al. [2017, 2020] obtained the first convergence rate of O(1T(d))O(\frac{1}{T(d)}) for the PoA when the demands of all O/D pairs grow to \infty at the same rate of Θ(T(d))\Theta(T(d)). For the more specific case of BPR cost functions (Bureau of Public Roads [1964]) with degree β0\beta\geq 0, Wu et al. [2021] showed a stronger convergence rate of O(1T(d)β)O(\frac{1}{T(d)^{\beta}}) for the PoA when the total demand T(d)T(d) grows to .\infty. See Section 5 for an overview of related results on the convergence rates of the PoA. With Theorem 9, we show that the PoA converges to 11 at a rate of O(1/ln(T(d)+1))O(\sqrt{1/\ln(T(d)+1)}) as T(d)T(d)\to\infty for regularly varying (Bingham et al. [1987]) cost functions of the form ζaxβlnα(x+1),\zeta_{a}\cdot x^{\beta}\cdot\ln^{\alpha}(x+1), β>0,α0,\beta>0,\ \alpha\geq 0, see Corollary 5.2. This is the first explicit convergence rate of the PoA in traffic networks with regularly varying cost functions that are not polynomials for the case T(d)T(d)\to\infty.

Altogether, we have considerably enhanced the sensitivity results of Englert et al. [2010], Takalloo and Kwon [2020] and Cominetti et al. [2020] by considering general traffic networks and simultaneous changes of demands and cost functions. Our results establish the first sensitivity analysis of the PoA for simultaneous changes of demands and cost functions in traffic networks with multiple O/D pairs.

These sensitivity results give also new insights into congestion pricing with tolls. Tolls change the cost functions, and so—due to our sensitivity results—tolls need to be considerable in order to reduce the PoA significantly. This has, e.g., been observed by Harks et al. [2015]. They consider tolls on a limited number of streets and use steepest descent methods to reduce the PoA. Their empirical calculations stabilize quickly with decreasing changes of the tolls, as justified in hindsight by our results. Furthermore, our results help to analyze the convergence rates of the PoA when the total demand T(d)T(d) tends to zero or infinity.

Although we use the terminology of traffic networks in this paper, our analysis and results do not depend on this view and carry over naturally to arbitrary non-atomic congestion games.

1.2 Related work

Koutsoupias and Papadimitriou [1999] proposed to quantify the inefficiency of equilibria in arbitrary congestion games from a worst-case perspective. This then resulted in the concept of the price of anarchy (PoA) that is usually defined as the ratio of the worst case equilibrium cost over the socially optimal cost, see Papadimitriou [2001].

1.2.1 Early development

A wave of research has been started with the pioneering paper of Roughgarden and Tardos [2002] on the PoA in traffic networks with affine linear cost functions. Examples are Roughgarden [2001, 2003]; Roughgarden and Tardos [2004]; Roughgarden [2005, 2015]; Correa et al. [2004, 2005]. They investigated the worst-case upper bound of the PoA for different types of cost functions τa()\tau_{a}(\cdot), and analyzed the influence of the network topology on this bound. In particular, they showed that this bound is 43\tfrac{4}{3} when all τa()\tau_{a}(\cdot) are affine linear (Roughgarden and Tardos [2002]), and Θ(βlnβ)\Theta(\frac{\beta}{\ln\beta}) when all τa()\tau_{a}(\cdot) are polynomials with maximum degree β>0\beta>0 (Roughgarden and Tardos [2004] and Roughgarden [2015]). Moreover, they proved that this bound is independent of the network topology, see, e.g., Roughgarden [2003]. They also developed a (λ,μ)(\lambda,\mu)-smooth method by which one can obtain a tight and robust worst-case upper bound of the PoA for a large class of cost functions, see, e.g., Roughgarden [2003], Roughgarden and Tardos [2004] and Roughgarden [2015]. This method was reproved by Correa et al. [2005] from a geometric perspective. Moreover, Perakis [2007] considered worst-case upper bounds for non-separable cost functions. See Roughgarden and Tardos [2007] for a comprehensive overview of the early development of that research.

1.2.2 Convergence analysis of the PoA in traffic networks

Recent papers have empirically studied the PoA in traffic networks with BPR cost functions (Bureau of Public Roads [1964]) and real traffic data. Youn et al. [2008] observed that the empirical PoA in a traffic network depends crucially on the total demand. Starting from 1, it grows with some oscillations, and ultimately becomes 1 again as the total demand increases. A similar observation was made by O’Hare et al. [2016]. They even conjectured that the PoA converges to 11 in the power law 1+O(T(d)2β)1+O\big{(}T(d)^{-2\cdot\beta}\big{)} when the total demand T(d)T(d) becomes large. Monnot et al. [2017] showed that traffic choices of commuting students in Singapore are near-optimal and that the empirical PoA is much smaller than known worst-case upper bounds. Similar observations have been reported by Jahn et al. [2005].

These empirical observations have been recently confirmed by Colini-Baldeschi et al. [2016, 2017, 2020] and Wu et al. [2021]. Colini-Baldeschi et al. [2016, 2017, 2020] were the first to theoretically analyze the convergence of the PoA in traffic networks with growing or decreasing total demand.

As a first step, Colini-Baldeschi et al. [2016] showed the convergence of the PoA to 11 as the total demand T(d)T(d)\to\infty for traffic networks with a single O/D pair and regularly varying cost functions. This convergence result was then substantially extended by Colini-Baldeschi et al. [2017] to traffic networks with multiple O/D pairs for both the case T(d)0T(d)\to 0 and the case T(d),T(d)\to\infty, when the ratio of the demand of each O/D pair over the total demand T(d)T(d) remains positive as T(d)0T(d)\to 0 or \infty. Finally, Colini-Baldeschi et al. [2020] extended these results to the cases where the demands and the network together fulfill certain tightness and salience conditions that allow the ratios of demands to vary in a certain pattern as T(d)0T(d)\to 0 or .\infty. Moreover, they illustrated by an example that the PoA need not converge to 11 as T(d)T(d)\to\infty when the cost functions are not regularly varying. In particular, they obtained the first convergence rates of the PoA in traffic network with polynomial cost functions when the ratio of the demand of each O/D pair over the total demand T(d)T(d) stays positive as T(d)0T(d)\to 0 or .\infty. We will come back to these rates in Section 5.

Wu et al. [2021] extended the work of Colini-Baldeschi et al. [2016, 2017, 2020] for growing total demand by a new framework. They showed for traffic networks with arbitrary regularly varying functions that the PoA converges to 11 as the total demand tends to \infty regardless of the growth pattern of the demands. In particular, they proved a convergence rate of O(T(d)β)O(T(d)^{-\beta}) for BPR cost functions of degree β\beta and illustrated by examples that the conjecture proposed by O’Hare et al. [2016] need not hold.

1.2.3 Sensitivity analysis of the PoA in traffic networks

First results on the sensitivity of the PoA in traffic networks have been obtained recently by Englert et al. [2010], Takalloo and Kwon [2020], and Cominetti et al. [2020].

Englert et al. [2010] considered traffic networks with a single origin-destination (O/D) pair and polynomial cost functions of degree at most β.\beta. They viewed the PoA as a function of the demand of that O/D pair when the polynomial cost functions do not change. They showed that the increase of this PoA function is bounded by a factor of (1+ϵ)β(1+\epsilon)^{\beta} from above when the demand increases by a factor of 1+ϵ1+\epsilon.

Takalloo and Kwon [2020] generalized this result to traffic networks with multiple O/D pairs and polynomial cost functions of degree at most β\beta. They proved the same upper bound on the increase of the PoA when the demands of all O/D pairs increase synchronously by the same factor 1+ϵ1+\epsilon and the polynomial cost functions do not change. They also showed that the increase of the PoA is bounded by a factor 1(1+ϵ)β\frac{1}{(1+\epsilon)^{\beta}} from below under the same conditions.

Similar to Englert et al. [2010], Cominetti et al. [2020] considered also traffic networks with a single O/D pair, and viewed the PoA as a function of the demand of that O/D pair for fixed cost functions. For cost functions with strictly positive derivatives, they showed that the PoA function is differentiable at each demand level where all the optimal paths carry a strictly positive flow. For affine cost functions, they showed further that the equilibrium cost is piece-wise linear and differentiable except at so-called \mathcal{E}-breakpoints whose number is finite though possibly exponentially large in the size of the network. In the interval between any two consecutive \mathcal{E}-breakpoints, the PoA function is differentiable, and either monotone or unimodal with a unique minimum on the interior of that interval. So the maximum of the PoA function is attained at an \mathcal{E}-breakpoint. They also presented several examples showing how these properties fail for general cost functions.

1.2.4 Sensitivity analysis of equilibria in traffic networks

Related results on the sensitivity of Wardrop equilibria have been obtained by Hall [1978], Patriksson [2004], Josefsson and Patriksson [2007], Lu [2008], Klimm and Warode [2021] and others. Hall [1978] proved that the Wardrop equilibrium path cost of an O/D pair is continuous, and even montonically non-decreasing with the growth of its demand when both the demands of other O/D pairs and the cost functions are fixed. Consequently, the total cost of Wardrop equilibria is a continuous function of the demands. Theorem 6 generalizes this continuity to the whole topological game space.

Patriksson [2004], Josefsson and Patriksson [2007] and Lu [2008] considered the sensitivity of Wardrop equilibria w.r.t. changes of parameters of the demands and the cost functions when both the demands and the cost functions are parametric and contain parameters defined on finite dimensional Euclidean spaces. In this case, the non-atomic congestion game is characterized completely by a parameter vector μn\mu\in\mathbb{R}^{n} for a fixed integer n>0n>0, and the Wardrop equilibrium arc flows and cost are then functions of the parameter vector μn\mu\in\mathbb{R}^{n} that map μ\mu to “points” on the Euclidean space A\mathbb{R}^{A}. When the cost functions are differentiable, Patriksson [2004] characterized the existence of a directional derivative of the Wardrop equilibrium solution (arc flow and arc cost) in an arbitrary direction of μ\mu. Josefsson and Patriksson [2007] further improved Patriksson [2004], and showed that the Wardrop equilibrium arc cost is always directionally differentiable w.r.t. μ\mu, while the Wardrop equilibrium arc flow may not. Moreover, Lu [2008] derived methods to calculate the semiderivatives of the Wardrop equilibrium arc flow w.r.t. μ\mu under general conditions, and the derivatives of the Wardrop equilibrium arc flow w.r.t. μ\mu under more restrictive conditions. Here, the semiderivative of a function h:nAh:\mathbb{R}^{n}\to\mathbb{R}^{A} at a point μ0n\mu_{0}\in\mathbb{R}^{n} refers to a continuous and positively homogeneous function δμ0:nm\delta_{\mu_{0}}:\mathbb{R}^{n}\to\mathbb{R}^{m} s.t. h(μ)=h(μ0)+δμ0(μμ0)+o(μμ0)h(\mu)=h(\mu_{0})+\delta_{\mu_{0}}(\mu-\mu_{0})+o(\|\mu-\mu_{0}\|) for each μn\mu\in\mathbb{R}^{n} with sufficiently small μμ0.\|\mu-\mu_{0}\|. The vector δμ0(μμ0)\delta_{\mu_{0}}(\mu-\mu_{0}) is then the directional derivative of hh at μ0\mu_{0} w.r.t. direction μμ0.\mu-\mu_{0}. In particular, the derivative of hh at μ0\mu_{0} exists and equals δμ0()\delta_{\mu_{0}}(\cdot) when the semiderivative δμ0()\delta_{\mu_{0}}(\cdot) at μ0\mu_{0} is a linear function.

The recent seminal work by Klimm and Warode [2021], see also the conference version Klimm and Warode [2019], developed a Katzenelson’s homotopy method to compute all Wardrop equilibria for a non-atomic congestion game with piece-wise linear cost when the demand vector has the form d=λd(0)d=\lambda\cdot d^{(0)} for a fixed direction d(0)(0,)𝒦d^{(0)}\in(0,\infty)^{\mathcal{K}} and a variable parameter λ(0.).\lambda\in(0.\infty). They viewed the Wardrop equilibrium arc flow as a function of the parameter λ\lambda, and proved that this function is actually piece-wise linear in λ\lambda, and, in particular, may have exponentially many breakpoints when the cost functions are affine linear.

While these parametric sensitivity results of Wardrop equilibira are very interesting, they do not apply to our sensitivity analysis of the PoA, since neither the demands nor the cost functions are parametric. In fact, we consider the most general case that both the cost functions and the demands may vary arbitrarily, and so cannot be parameterized by a finite dimensional Euclidean space. Nevertheless, they are very inspiring, and pave a feasible way for future work on the differentiability of the PoA.

1.3 Outline of the paper

The paper is organized as follows. We develop our approach for general non-atomic congestion games but with the terminology of traffic networks. These are introduced in Section 2. Section 3 defines the metric and the topological space for games with the same combinatorial structure. Section 4 then presents our techniques and sensitivity analysis results. Section 5 demonstrates that our results also facilitate the analysis of the convergence rate of the PoA when the total demand tends to 0 or .\infty. We conclude with a short summary and discussion in Section 6.

2 Model and preliminaries

2.1 The model

We define an arbitrary non-atomic congestion game with the terminology of traffic games (see, e.g., Nisan et al. [2007]; Roughgarden and Tardos [2002]), since this is more intuitive. A non-atomic congestion game Γ\Gamma is then associated with a traffic network G=(V,A)G=(V,A), and consists of a tuple (𝒦,𝒮,τ,d)(\mathcal{K},\mathcal{S},\tau,d) with components defined in G1–G5 below.

  • G1

    𝒦\mathcal{K} is a finite non-empty set of groups or populations of users. Every group k𝒦k\in\mathcal{K} corresponds to a (transport) origin-destination (O/D) pair in the network G.G. We will write an O/D pair k𝒦k\in\mathcal{K} as (ok,tk)V2(o_{k},t_{k})\in V^{2} when the origin oko_{k} and the destination tkt_{k} are specified.

  • G2

    𝒮=k𝒦𝒮k,\mathcal{S}=\cup_{k\in\mathcal{K}}\mathcal{S}_{k}, where each 𝒮k2A\mathcal{S}_{k}\subseteq 2^{A}\setminus\emptyset denotes a non-empty collection of (ok,tk)(o_{k},t_{k})-paths that are (pure) strategies available only to users of O/D pair kk. Here, a (ok,tk)(o_{k},t_{k})-path is a non-empty subset of the arc set AA that is loop-free and leads from the origin oko_{k} to the destination tk.t_{k}. The sets 𝒮k\mathcal{S}_{k} are then mutually disjoint, i.e., 𝒮k𝒮k=\mathcal{S}_{k}\cap\mathcal{S}_{k^{\prime}}=\emptyset for any two distinct O/D pairs k,k𝒦.k,k^{\prime}\in\mathcal{K}.

  • G3

    τ=(τa)aA\tau=(\tau_{a})_{a\in A} is a cost function vector, in which each τa:[0,)[0,)\tau_{a}:[0,\infty)\to[0,\infty) is a continuous and non-decreasing latency or cost function of arc aAa\in A that depends on the flow value of arc aa and includes also all other extra cost for using arc aa such as tolls.

  • G4

    d=(dk)k𝒦d=(d_{k})_{k\in\mathcal{K}} is a demand vector whose component dk0d_{k}\geq 0 denotes the demand to be transported by (users of) O/D pair k𝒦k\in\mathcal{K} using paths in 𝒮k.\mathcal{S}_{k}. So Γ\Gamma has the total (transport) demand T(d):=k𝒦dk.T(d):=\sum_{k\in\mathcal{K}}d_{k}.

  • G5

    Users are non-cooperative. Each user of an arbitrary O/D pair k𝒦k\in\mathcal{K} is infinitesimal, i.e., controls an infinitesimal fraction of the demand dkd_{k}, and must satisfy that by choosing a path s𝒮ks\in\mathcal{S}_{k}. The demand dkd_{k} will then be arbitrarily split among paths in 𝒮k\mathcal{S}_{k} for each O/D pair k𝒦.k\in\mathcal{K}.

A (pure) strategy profile (or simply, profile) over all users is expressed as a multi-commodity (path) flow f=(fs)s𝒮=(fs)s𝒮k,k𝒦f=(f_{s})_{s\in\mathcal{S}}=(f_{s})_{s\in\mathcal{S}_{k},k\in\mathcal{K}} of the network GG with sSkfs=dk\sum_{s^{\prime}\in S_{k}}f_{s^{\prime}}=d_{k} for each O/D pair k𝒦.k\in\mathcal{K}. Herein, the flow value fs0f_{s}\geq 0 is just the demand transported along the path s𝒮.s\in\mathcal{S}. The flow value faf_{a} of an arc aAa\in A is then obtained as fa:=s𝒮:asfsf_{a}:=\sum_{s\in\mathcal{S}:a\in s}f_{s}. Hence, an arc aAa\in A has the cost τa(fa),\tau_{a}(f_{a}), a path s𝒮s\in\mathcal{S} has the cost τs(f):=aA:asτa(fa),\tau_{s}(f):=\sum_{a\in A:a\in s}\tau_{a}(f_{a}), and all users have the total cost C(Γ,f):=s𝒮fsτs(f)=aAfaτa(fa)C(\Gamma,f):=\sum_{s\in\mathcal{S}}f_{s}\cdot\tau_{s}(f)=\sum_{a\in A}f_{a}\cdot\tau_{a}(f_{a}).

We call the triple (G,𝒦,𝒮)(G,\mathcal{K},\mathcal{S}) the combinatorial structure of Γ\Gamma, and denote Γ\Gamma by simply the pair (τ,d)(\tau,d) when its combinatorial structure (G,𝒦,𝒮)(G,\mathcal{K},\mathcal{S}) is given.

Viewed as a general non-atomic congestion game, the arcs aAa\in A and paths s𝒮s\in\mathcal{S} correspond to resources and (pure) strategies of users, see, e.g., Dafermos and Sparrow [1969], Rosenthal [1973] and Correa et al. [2005]. Although we chose to use the terminology of traffic games, the analysis and results are independent of this view and carry over to general non-atomic congestion games.

2.2 Equilibria, optimality and the price of anarchy

A flow f=(fs)s𝒮f^{*}=(f^{*}_{s})_{s\in\mathcal{S}} of Γ\Gamma is called a social optimum (SO) if it has the minimum total cost, i.e., C(Γ,f)C(Γ,f)C(\Gamma,f^{*})\leq C(\Gamma,f) for an arbitrary flow ff of Γ.\Gamma.

A flow f~=(f~s)s𝒮\tilde{f}=(\tilde{f}_{s})_{s\in\mathcal{S}} of Γ\Gamma is called a Wardrop equilibrium (WE) if it fulfills Wardrop’s first principle (Wardrop [1952]), i.e.,

k𝒦s,s𝒮k(f~s>0τs(f~)τs(f~)).\forall k\in\mathcal{K}\ \forall s,s^{\prime}\in\mathcal{S}_{k}\Big{(}\tilde{f}_{s}>0\implies\tau_{s}(\tilde{f})\leq\tau_{s^{\prime}}(\tilde{f})\Big{)}. (2.1)

Clearly, every WE flow f~\tilde{f} of Γ\Gamma satisfies condition (2.2):

k𝒦s𝒮k:f~s>0τs(f~)=Lk(τ,d):=mins𝒮kτs(f~).\forall k\in\mathcal{K}\ \forall s\in\mathcal{S}_{k}:\quad\tilde{f}_{s}>0\implies\tau_{s}(\tilde{f})=L_{k}(\tau,d):=\min_{s^{\prime}\in\mathcal{S}_{k}}\tau_{s^{\prime}}(\tilde{f}). (2.2)

We call the constant Lk(τ,d)L_{k}(\tau,d) in condition (2.2) the user cost of O/D pair k𝒦.k\in\mathcal{K}. The total cost of WE flows f~\tilde{f} is then expressed equivalently by

C(Γ,f~)=s𝒮f~sτs(f~)=k𝒦Lk(τ,d)dk.C(\Gamma,\tilde{f})=\sum_{s\in\mathcal{S}}\tilde{f}_{s}\cdot\tau_{s}(\tilde{f})=\sum_{k\in\mathcal{K}}L_{k}(\tau,d)\cdot d_{k}. (2.3)

Trivially, a flow f~\tilde{f} of Γ\Gamma is a WE if and only if f~\tilde{f} satisfies the variational inequality

aAτa(f~a)(gaf~a)=s𝒮τs(f~)(gsf~s)>0\sum_{a\in A}\tau_{a}(\tilde{f}_{a})\cdot(g_{a}-\tilde{f}_{a})=\sum_{s\in\mathcal{S}}\tau_{s}(\tilde{f})\cdot(g_{s}-\tilde{f}_{s})>0 (2.4)

for an arbitrary flow g=(gs)s𝒮g=(g_{s})_{s\in\mathcal{S}} of Γ,\Gamma, see, e.g., Dafermos [1980].

Since the cost functions τa()\tau_{a}(\cdot) are non-negative, continuous and non-decreasing, and since the users are infinitesimal, Γ\Gamma has essentially unique WE flows, i.e., τa(f~a)=τa(f~a)\tau_{a}(\tilde{f}_{a})=\tau_{a}(\tilde{f}^{\prime}_{a}) for all aAa\in A for two arbitrary WE flows f~\tilde{f} and f~\tilde{f}^{\prime} of Γ,\Gamma, see, e.g., Schmeidler [1973], Smith [1979] and Roughgarden and Tardos [2002]. In particular, WE flows of Γ\Gamma coincide with (pure) Nash equilibria (NE) of Γ\Gamma, see, e.g., Roughgarden and Tardos [2002] for a definition of NE flows in non-atomic congestion games.

When all cost functions τa()\tau_{a}(\cdot) are differentiable, an SO flow ff^{*} is also a WE flow w.r.t. the marginal cost functions ca(x):=xτa(x)+τa(x),c_{a}(x):=x\cdot\tau^{\prime}_{a}(x)+\tau_{a}(x), and, vice versa, a WE flow is an SO flow w.r.t. the cost functions

1x0xτa(ξ)𝑑ξ,\frac{1}{x}\int_{0}^{x}\tau_{a}(\xi)d\xi,

see, e.g., Beckmann et al. [1956] or Roughgarden and Tardos [2002]. Hence, SO flows coincide with WE flows when all cost functions τa()\tau_{a}(\cdot) are monomials of the same degree β0.\beta\geq 0.

In general, SO flows and WE flows differ, see, e.g., Roughgarden and Tardos [2002]. The ratio of the worst-case total cost of a WE flow over the total cost of an SO flow is known as the Price of Anarchy (PoA, see Koutsoupias and Papadimitriou [1999] and Papadimitriou [2001]), and measures the inefficiency of WE flows. Formally, the PoA of Γ\Gamma is defined as

ρ(Γ):=C(Γ,f~)C(Γ,f)=s𝒮f~sτs(f~)s𝒮fsτs(f),\rho(\Gamma):=\frac{C(\Gamma,\tilde{f})}{C(\Gamma,f^{*})}=\frac{\sum_{s\in\mathcal{S}}\tilde{f}_{s}\cdot\tau_{s}(\tilde{f})}{\sum_{s\in\mathcal{S}}f^{*}_{s}\cdot\tau_{s}(f^{*})}, (2.5)

where f~\tilde{f} and ff^{*} are an arbitrary WE flow and an arbitrary SO flow of Γ,\Gamma, respectively. Definition (2.5) is unambiguous since WE flows are essentially unique.

2.3 Potential functions and ϵ\epsilon-approximate equilibria

A non-atomic congestion game Γ\Gamma is actually a potential game, see, e.g., Sandholm [2001]. The (Rosenthal) potential function of Γ\Gamma is given by

Φ(Γ,f)=aA0faτa(x)𝑑x,\Phi(\Gamma,f)=\sum_{a\in A}\int_{0}^{f_{a}}\tau_{a}(x)\ dx, (2.6)

and reaches its global minimum at its WE flows f~\tilde{f}, see, e.g., Roughgarden and Tardos [2002].

A flow ff is an ϵ\epsilon-approximate WE flow of Γ\Gamma for a small constant ϵ>0,\epsilon>0, if

aAτa(fa)(faga)=s𝒮τs(f)(fsgs)=k𝒦s𝒮kτs(f)(fsgs)ϵ\sum_{a\in A}\tau_{a}(f_{a})\cdot(f_{a}-g_{a})=\sum_{s\in\mathcal{S}}\tau_{s}(f)\cdot(f_{s}-g_{s})=\sum_{k\in\mathcal{K}}\sum_{s\in\mathcal{S}_{k}}\tau_{s}(f)\cdot(f_{s}-g_{s})\leq\epsilon (2.7)

for an arbitrary flow g=(gs)s𝒮g=(g_{s})_{s\in\mathcal{S}} of Γ\Gamma. The variational inequality (2.7) implies that

τs(f)τs(f)+ϵmins′′𝒮:fs′′>0fs′′=τs(f)+Θ(ϵ)\tau_{s}(f)\leq\tau_{s^{\prime}}(f)+\frac{\epsilon}{\min_{s^{\prime\prime}\in\mathcal{S}:\ f_{s^{\prime\prime}}>0}f_{s^{\prime\prime}}}=\tau_{s^{\prime}}(f)+\Theta(\epsilon) (2.8)

for two arbitrary paths s,s𝒮ks,s^{\prime}\in\mathcal{S}_{k} with fs>0,f_{s}>0, for every O/D pair k𝒦.k\in\mathcal{K}. Inequality (2.8) means that a unilateral change of paths in an ϵ\epsilon-approximate WE flow ff reduces the cost by at most Θ(ϵ),\Theta(\epsilon), and so ff indeed approximates a WE flow. In principle, we can alternatively define an ϵ\epsilon-approximate WE flow directly by inequality (2.8) (as was done, e.g., by Roughgarden and Tardos [2002]). But this does not considerably influence our analysis. We thus stick to the variational inequality definition (2.7), since it, together with the variational inequality (2.4), facilitates the cost comparison between an ϵ\epsilon-approximate WE flow and a precise WE flow, see, e.g., Lemma 1b).

Lemma 1 shows some useful properties of ϵ\epsilon-approximate WE flows. Here, a real-valued function h()h(\cdot) is Lipschitz continuous on an interval I[0,)I\subseteq[0,\infty) with Lipschitz constant MM if |h(x)h(y)|M|xy||h(x)-h(y)|\leq M\cdot|x-y| for all x,yI.x,y\in I.

Lemma 1.

Consider an arbitrary non-atomic congestion game Γ\Gamma with cost function vector τ\tau and demand vector d,d, and a constant ϵ>0.\epsilon>0. Let ff be an ϵ\epsilon-approximate WE flow, and let f~\tilde{f} be a WE flow. Then ff and f~\tilde{f} fulfill the following conditions.

  • a)

    For each O/D pair k𝒦,k\in\mathcal{K}, 0sSkfsτs(f)dkmins𝒮kτs(f)<ϵ.0\leq\sum_{s\in S_{k}}f_{s}\cdot\tau_{s}(f)-d_{k}\cdot\min_{s^{\prime}\in\mathcal{S}_{k}}\tau_{s^{\prime}}(f)<\epsilon. Moreover, k𝒦dkmins𝒮kτs(f)C(Γ,f)k𝒦dkmins𝒮kτs(f)+ϵ.\sum_{k\in\mathcal{K}}d_{k}\cdot\min_{s^{\prime}\in\mathcal{S}_{k}}\tau_{s^{\prime}}(f)\leq C(\Gamma,f)\leq\sum_{k\in\mathcal{K}}d_{k}\cdot\min_{s^{\prime}\in\mathcal{S}_{k}}\tau_{s^{\prime}}(f)+\epsilon.

  • b)

    0aAτa(f~a)(faf~a)Φ(Γ,f)Φ(Γ,f~)aAτa(fa)(faf~a)<ϵ,0\leq\sum_{a\in A}\tau_{a}(\tilde{f}_{a})\cdot(f_{a}-\tilde{f}_{a})\leq\Phi(\Gamma,f)-\Phi(\Gamma,\tilde{f})\leq\sum_{a\in A}\tau_{a}(f_{a})\cdot(f_{a}-\tilde{f}_{a})<\epsilon, and thus 0aA|τa(fa)τa(f~a)||faf~a|<ϵ.0\leq\sum_{a\in A}|\tau_{a}(f_{a})-\tau_{a}(\tilde{f}_{a})|\cdot|f_{a}-\tilde{f}_{a}|<\epsilon.

  • c)

    If every τa()\tau_{a}(\cdot) is Lipschitz continuous on [0,T(d)][0,T(d)] with Lipschitz constant M>0M>0, then |τa(fa)τa(f~a)|<Mϵ|\tau_{a}(f_{a})-\tau_{a}(\tilde{f}_{a})|<\sqrt{M\cdot\epsilon} for all arcs aA,a\in A, and |Lk(τ,d)mins𝒮kτs(f)||A|Mϵ|L_{k}(\tau,d)-\min_{s^{\prime}\in\mathcal{S}_{k}}\tau_{s^{\prime}}(f)|\leq|A|\cdot\sqrt{M\cdot\epsilon} for all O/D pairs k𝒦.k\in\mathcal{K}. Furthermore, |C(Γ,f)C(Γ,f~)||A|MϵT(d)+ϵ.|C(\Gamma,f)-C(\Gamma,\tilde{f})|\leq|A|\cdot\sqrt{M\cdot\epsilon}\cdot T(d)+\epsilon.

Lemma 1a) follows trivially from inequality (2.7). Lemma 1b) follows directly from (2.4), (2.6)–(2.7) and the fact that

τa(x)(yx)xyτa(z)𝑑zτa(y)(yx)aAx0y0.\tau_{a}(x)\cdot(y-x)\leq\int_{x}^{y}\tau_{a}(z)\ dz\leq\tau_{a}(y)\cdot(y-x)\quad\ \forall a\in A\ \forall x\in\mathbb{R}_{\geq 0}\ \forall y\in\mathbb{R}_{\geq 0}. (2.9)

Herein, 0:=[0,).\mathbb{R}_{\geq 0}:=[0,\infty). Inequality (2.9) holds because every cost function τa()\tau_{a}(\cdot) is non-decreasing, non-negative and continuous. Lemma 1c) is a direct consequence of Lemma 1a)–b). It yields an approximation bound when all cost functions are Lipschitz continuous on the compact interval [0,T(d)],[0,T(d)], which plays a pivotal role in the Hölder continuity analysis of the PoA in Section 4.2. Although this approximation bound is rather trivial, it is tight, see Example 2 below.

Example 2.

We illustrate the tightness of Lemma 1c) with Pigou’s game (Pigou [1920]). Pigou’s game Γ\Gamma has one unit of total demand and the simple traffic network shown in Figure 1. It thus has the unique WE flow f~=(f~u,f~)=(1,0),\tilde{f}=(\tilde{f}_{u},\tilde{f}_{\ell})=(1,0), where uu and \ell denote the upper and lower arcs (paths), respectively. Let ϵ>0\epsilon>0 be an arbitrary small constant, and put f=(fu,f)=(1ϵ,ϵ).f=(f_{u},f_{\ell})=(1-\sqrt{\epsilon},\sqrt{\epsilon}). Then ff is an ϵ\epsilon-approximate WE flow, since

τu(fu)(fugu)+τ(f)(fg)=(1ϵ)(1ϵgu)+ϵg=12ϵ+ϵgu+ϵgu+ϵg=ϵϵ(1gu)ϵ.\begin{split}\tau_{u}(f_{u})\cdot(f_{u}-g_{u})+&\tau_{\ell}(f_{\ell})\cdot(f_{\ell}-g_{\ell})=(1-\sqrt{\epsilon})\cdot(1-\sqrt{\epsilon}-g_{u})+\sqrt{\epsilon}-g_{\ell}\\ &=1-2\cdot\sqrt{\epsilon}+\epsilon-g_{u}+\sqrt{\epsilon}\cdot g_{u}+\sqrt{\epsilon}-g_{\ell}=\epsilon-\sqrt{\epsilon}\cdot(1-g_{u})\leq\epsilon.\end{split}

for an arbitrary flow g=(gu,g).g=(g_{u},g_{\ell}). Here, we use that gu+g=1g_{u}+g_{\ell}=1 for each g.g. Furthermore, |τu(f~u)τu(fu)|=|f~ufu|=ϵ|\tau_{u}(\tilde{f}_{u})-\tau_{u}(f_{u})|=|\tilde{f}_{u}-f_{u}|=\sqrt{\epsilon} and |C(Γ,f)C(Γ,f~)|=ϵϵΘ(ϵ),|C(\Gamma,f)-C(\Gamma,\tilde{f})|=\sqrt{\epsilon}-\epsilon\in\Theta(\sqrt{\epsilon}), which shows that Lemma 1c) is tight.

oottxx11
Figure 1: The Pigou’s game

Because of Lemma 1c), we may thus want to find for a given flow ff an ϵ>0\epsilon>0 such that ff is an ϵ\epsilon^{\prime}-approximate WE flow for all ϵ>ϵ,\epsilon^{\prime}>\epsilon, but not for all ϵ[0,ϵ).\epsilon^{\prime}\in[0,\epsilon). We call such a constant ϵ(0,)\epsilon\in(0,\infty) the approximation threshold of flow ff w.r.t. WE flows f~\tilde{f} of a non-atomic congestion game Γ=(τ,d)\Gamma=(\tau,d), and denote it by ϵ(τ,d,f)\epsilon(\tau,d,f) to indicate its dependence on τ,\tau, dd and ff. Note that ϵ(τ,d,f)=k𝒦sSk[τs(f)Lk(τ,d,f)]fs\epsilon(\tau,d,f)=\sum_{k\in\mathcal{K}}\sum_{s\in S_{k}}[\tau_{s}(f)-L_{k}(\tau,d,f)]\cdot f_{s} for each flow f,f, where Lk(τ,d,f):=mins𝒮kτs(f)L_{k}(\tau,d,f):=\min_{s\in\mathcal{S}_{k}}\ \tau_{s}(f) is the minimum path cost of O/D pair k𝒦k\in\mathcal{K} in flow f.f. This follows since

k𝒦Lk(τ,d,f)dk=ming is a flow of Γk𝒦sSkτs(f)gs.\sum_{k\in\mathcal{K}}L_{k}(\tau,d,f)\cdot d_{k}=\min_{g\text{ is a flow of }\Gamma}\sum_{k\in\mathcal{K}}\sum_{s\in S_{k}}\tau_{s}(f)\cdot g_{s}.

Lemma 1c) with this approximation threshold ϵ(τ,d,f)\epsilon(\tau,d,f) may then yield a tight upper bound of |C(Γ,f)C(Γ,f~)||C(\Gamma,f)-C(\Gamma,\tilde{f})| for arbitrary Lipschitz continuous cost functions on [0,T(d)][0,T(d)], see Example 2. We will obtain a tight upper bound for this approximation threshold in Section 4.2.

3 The topological space of all non-atomic congestion games with the same combinatorial structure

In the sequel, we fix an arbitrary combinatorial structure (G,𝒦,𝒮)(G,\mathcal{K},\mathcal{S}) and consider only non-atomic congestion games with this combinatorial structure. A non-atomic congestion game Γ\Gamma is then simply specified by the pair (τ,d).(\tau,d).

3.1 Assumptions

To avoid unnecessary discussions, we assume that the fixed combinatorial structure (G,𝒦,𝒮)(G,\mathcal{K},\mathcal{S}) satisfies Condition 1 below.

Condition 1

Every arc aAa\in A belongs to some path s𝒮=k𝒦𝒮k,s\in\mathcal{S}=\cup_{k^{\prime}\in\mathcal{K}}\mathcal{S}_{k^{\prime}}, and |𝒮k|2|\mathcal{S}_{k}|\geq 2 for each O/D pair k𝒦k\in\mathcal{K}.

Condition 1 can be required w.l.o.g., as arcs aAa\in A with asa\notin s for each s𝒮s\in\mathcal{S} play no role in a non-atomic congestion game Γ=(τ,d)\Gamma=(\tau,d), and an O/D pair k𝒦k\in\mathcal{K} with a singleton path set 𝒮k={s}\mathcal{S}_{k}=\{s\} can be removed by using τa(x+dk)\tau_{a}(x+d_{k}) instead of τa(x)\tau_{a}(x) for each arc aa belonging to the unique path ss in 𝒮k.\mathcal{S}_{k}.

The PoA ρ(Γ)\rho(\Gamma) is obviously 00\frac{0}{0} when the total demand T(d)=0.T(d)=0. In fact, even if T(d)>0,T(d)>0, the PoA may still be 00\frac{0}{0} when we allow τa(x)=0\tau_{a}(x)=0 for some aAa\in A and some x(0,)x\in(0,\infty). To avoid these undefined cases of the PoA, we assume that an arbitrary non-atomic congestion game Γ=(τ,d)\Gamma=(\tau,d) satisfies Condition 2 below.

Condition 2

T(d)>0T(d)>0 and τa(x)>0\tau_{a}(x)>0 for all aAa\in A and all x(0,T(d)].x\in(0,T(d)].

Lemma 3 shows that ρ(Γ)[1,)\rho(\Gamma)\in[1,\infty) is well defined for each non-atomic congestion game Γ=(τ,d)\Gamma=(\tau,d) fulfilling Condition 2. The proof of Lemma 3 is trivial, and thus omitted.

Lemma 3.

Consider an arbitrary non-atomic congestion game Γ=(τ,d)\Gamma=(\tau,d) fulfilling Condition 2. Let f,f~f,\tilde{f} and ff^{*} be an arbitrary flow, a WE flow and an SO flow of Γ,\Gamma, respectively. Then fa[0,T(d)]f_{a}\in[0,T(d)] for all arcs aA,a\in A, and

0<T(d)|𝒮|minaAτa(T(d)|𝒮|)C(Γ,f)C(Γ,f~)|A|T(d)maxaAτa(T(d)).0<\frac{T(d)}{|\mathcal{S}|}\cdot\min_{a\in A}\tau_{a}\Big{(}\frac{T(d)}{|\mathcal{S}|}\Big{)}\leq C(\Gamma,f^{*})\leq C(\Gamma,\tilde{f})\leq|A|\cdot T(d)\cdot\max_{a\in A}\tau_{a}\big{(}T(d)\big{)}.

So ρ(Γ)[1,|A||𝒮|maxaAτa(T(d))minaAτa(T(d)/|𝒮|)][1,)\rho(\Gamma)\in[1,\frac{|A|\cdot|\mathcal{S}|\cdot\max_{a\in A}\tau_{a}(T(d))}{\min_{a\in A}\tau_{a}(T(d)/|\mathcal{S}|)}]\subseteq[1,\infty) and is thus well defined.

As the definitions of cost functions τa(x)\tau_{a}(x) on the unbounded interval (T(d),)(T(d),\infty) play no role in a non-atomic congestion game Γ=(τ,d),\Gamma=(\tau,d), we define an equivalence relation between non-atomic congestion games as follows.

Equivalence

Two non-atomic congestion games Γ=(τ,d)\Gamma=(\tau,d) and Γ=(σ,d)\Gamma^{\prime}=(\sigma,d^{\prime}) are equivalent, denoted by ΓΓ,\Gamma\simeq\Gamma^{\prime}, when

dk=dkk𝒦andτa(x)=σa(x)aAx[0,T(d)]=[0,T(d)].d_{k}=d^{\prime}_{k}\ \forall k\in\mathcal{K}\quad\text{and}\quad\tau_{a}(x)=\sigma_{a}(x)\ {\forall a\in A\ \forall x\in[0,T(d)]=[0,T(d^{\prime})].} (3.1)

While the cost function values τa(x)\tau_{a}(x) and σa(x)\sigma_{a}(x) might differ largely on the unbounded interval (T(d),),(T(d),\infty), the corresponding non-atomic congestion games Γ=(τ,d)\Gamma=(\tau,d) and Γ=(σ,d)\Gamma^{\prime}=(\sigma,d) have the same game-theoretic properties when ΓΓ\Gamma\simeq\Gamma^{\prime}.

3.2 The game space, the metric and the topology

We now introduce a topology for the space of all non-atomic congestion games defined on the combinatorial structure (G,𝒦,S).(G,\mathcal{K},S). All topological notions not explicitly defined here are standard, and we recommend Kelley [1975] as a standard reference for them.

We denote by 𝒢(G,𝒦,𝒮)\mathcal{G}(G,\mathcal{K},\mathcal{S}) the set of all non-atomic congestion games Γ=(τ,d)\Gamma=(\tau,d) that have the fixed combinatorial structure (G,𝒦,𝒮)(G,\mathcal{K},\mathcal{S}) and satisfy Condition 2. We call 𝒢(G,𝒦,𝒮)\mathcal{G}(G,\mathcal{K},\mathcal{S}) the (G,𝒦,𝒮)(G,\mathcal{K},\mathcal{S})-game space (or, simply, the game space), and call each non-atomic congestion game Γ=(τ,d)𝒢(G,𝒦,𝒮)\Gamma=(\tau,d)\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) a game, or a point of the game space.

Clearly,

𝒢(G,𝒦,𝒮)𝒢¯(G,𝒦,𝒮):=𝒞+(0)A×0𝒦,\mathcal{G}(G,\mathcal{K},\mathcal{S})\subsetneq\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}):=\mathcal{C}_{+}^{\uparrow}(\mathbb{R}_{\geq 0})^{A}\times\mathbb{R}_{\geq 0}^{\mathcal{K}}, (3.2)

where 𝒞+(0)\mathcal{C}_{+}^{\uparrow}(\mathbb{R}_{\geq 0}) denotes the set of all non-decreasing, non-negative and continuous functions defined on 0=[0,),\mathbb{R}_{\geq 0}=[0,\infty), 𝒞+(0)A:={(τa)aA:τa𝒞+(0)aA}\mathcal{C}_{+}^{\uparrow}(\mathbb{R}_{\geq 0})^{A}:=\{(\tau_{a})_{a\in A}:\tau_{a}\in\mathcal{C}_{+}^{\uparrow}(\mathbb{R}_{\geq 0})\ \forall a\in A\} and 0𝒦:={d=(dk)k𝒦:dk0k𝒦}.\mathbb{R}_{\geq 0}^{\mathcal{K}}:=\{d=(d_{k})_{k\in\mathcal{K}}:d_{k}\in\mathbb{R}_{\geq 0}\ \forall k\in\mathcal{K}\}. We call the super set 𝒢¯(G,𝒦,𝒮)\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}) in (3.2) the generalized game space and each element Γ¯=(τ,d)𝒢¯(G,𝒦,𝒮)\bar{\Gamma}=(\tau,d)\in\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}) a generalized game. Trivially, 𝒢¯(G,𝒦,𝒮)\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}) is the collection of all non-atomic congestion games with the fixed combinatorial structure (G,𝒦,𝒮).(G,\mathcal{K},\mathcal{S}). Note that the PoA may be undefined for some generalized games Γ¯𝒢¯(G,𝒦,𝒮)𝒢(G,𝒦,𝒮).\bar{\Gamma}\in\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S})\setminus\mathcal{G}(G,\mathcal{K},\mathcal{S}).

We now define a “distance” on the generalized game space by the binary operator Dist:𝒢¯(G,𝒦,𝒮)×𝒢¯(G,𝒦,𝒮)[0.)\text{Dist}:\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S})\times\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S})\to[0.\infty) with

Dist(Γ¯,Γ¯):=max{dd,maxaAmaxx[0,min{T(d),T(d)}]|τa(x)σa(x)|,τ(T(d))σ(T(d))}\text{Dist}(\bar{\Gamma},\bar{\Gamma}^{\prime}):=\max\left\{|\!|d-d^{\prime}|\!|_{\infty},\max_{a\in A}\max_{x\in[0,\min\{T(d),T(d^{\prime})\}]}|\tau_{a}(x)-\sigma_{a}(x)|,|\!|\tau(T(d))-\sigma(T(d^{\prime}))|\!|_{\infty}\right\} (3.3)

for two arbitrary generalized games Γ¯=(τ,d)\bar{\Gamma}=(\tau,d) and Γ¯=(σ,d).\bar{\Gamma}^{\prime}=(\sigma,d^{\prime}). Here, y:=maxi=1n|yn||\!|y|\!|_{\infty}:=\max_{i=1}^{n}|y_{n}| is the LL^{\infty}-norm for an arbitrary vector y=(y1,,yn)y=(y_{1},\ldots,y_{n}) with an arbitrary length nn\in\mathbb{N}, and τ(T(d))\tau(T(d)) and σ(T(d))\sigma(T(d)) are the respective vectors (τa(T(d)))aA(\tau_{a}(T(d)))_{a\in A} and (σa(T(d)))aA.(\sigma_{a}(T(d)))_{a\in A}. To simplify notation, we denote by

τ|T(d)σ|T(d):=max{maxaAmaxx[0,min{T(d),T(d)}]|τa(x)σa(x)|,τ(T(d))σ(T(d))}|\!|\tau_{|T(d)}-\sigma_{|T(d^{\prime})}|\!|_{\infty}:=\max\left\{\max_{a\in A}\max_{x\in[0,\min\{T(d),T(d^{\prime})\}]}|\tau_{a}(x)-\sigma_{a}(x)|,|\!|\tau(T(d))-\sigma(T(d^{\prime}))|\!|_{\infty}\right\} (3.4)

the “distance” between τ\tau and σ\sigma w.r.t. the restricted domains [0,T(d)][0,T(d)] and [0,T(d)].[0,T(d^{\prime})]. Then Dist(Γ¯,Γ¯)\text{Dist}(\bar{\Gamma},\bar{\Gamma}^{\prime}) in (3.3) is expressed equivalently as

Dist(Γ¯,Γ¯)=max{dd,τ|T(d)σ|T(d)}.\text{Dist}(\bar{\Gamma},\bar{\Gamma}^{\prime})=\max\left\{|\!|d-d^{\prime}|\!|_{\infty},|\!|\tau_{|T(d)}-\sigma_{|T(d^{\prime})}|\!|_{\infty}\right\}. (3.5)

Note that Dist(,)\text{Dist}(\cdot,\cdot) is consistent with equivalence relation (3.1), i.e., Dist(Γ¯,Γ¯′′)=Dist(Γ¯,Γ¯′′)\text{Dist}(\bar{\Gamma},\bar{\Gamma}^{\prime\prime})=\text{Dist}(\bar{\Gamma}^{\prime},\bar{\Gamma}^{\prime\prime}) for three arbitrary generalized games Γ¯,Γ¯,Γ¯′′𝒢¯(G,𝒦,𝒮)\bar{\Gamma},\bar{\Gamma}^{\prime},\bar{\Gamma}^{\prime\prime}\in\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}) with Γ¯Γ¯\bar{\Gamma}\simeq\bar{\Gamma}^{\prime}.

Lemma 4 shows that Dist(,)\text{Dist}(\cdot,\cdot) is actually a metric on 𝒢¯(G,𝒦,𝒮)\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}) w.r.t. equivalence relation (3.1). We thus denote Dist(Γ¯,Γ¯)\text{Dist}(\bar{\Gamma},\bar{\Gamma}^{\prime}) symbolically by Γ¯Γ¯|\!|\bar{\Gamma}-\bar{\Gamma}^{\prime}|\!| for two arbitrary generalized games Γ¯,Γ¯𝒢¯(G,𝒦,𝒮),\bar{\Gamma},\bar{\Gamma}^{\prime}\in\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}), as this is a more intuitive way to denote the metric despite of the undefined operator Γ¯Γ¯\bar{\Gamma}-\bar{\Gamma}^{\prime}.

Lemma 4.

Consider now three arbitrary generalized games Γ¯1,Γ¯2,Γ¯3𝒢¯(G,𝒦,𝒮).\bar{\Gamma}_{1},\bar{\Gamma}_{2},\bar{\Gamma}_{3}\in\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}). Then:

  • a)

    Dist(Γ¯1,Γ¯2)=Dist(Γ¯2,Γ¯1)0.\text{Dist}(\bar{\Gamma}_{1},\bar{\Gamma}_{2})=\text{Dist}(\bar{\Gamma}_{2},\bar{\Gamma}_{1})\geq 0.

  • b)

    Dist(Γ¯1,Γ¯2)=0\text{Dist}(\bar{\Gamma}_{1},\bar{\Gamma}_{2})=0 if and only if Γ¯1Γ¯2\bar{\Gamma}_{1}\simeq\bar{\Gamma}_{2}.

  • c)

    Dist(Γ¯1,Γ¯2)Dist(Γ¯1,Γ¯3)+Dist(Γ¯3,Γ¯2).\text{Dist}(\bar{\Gamma}_{1},\bar{\Gamma}_{2})\leq\text{Dist}(\bar{\Gamma}_{1},\bar{\Gamma}_{3})+\text{Dist}(\bar{\Gamma}_{3},\bar{\Gamma}_{2}).

Proof.

Proof of Lemma 4 Lemma 4a)–4b) are trivial. We prove only Lemma 4c). Assume, w.l.o.g., that Γ¯1=(τ(1)=(τa(1))aA,d(1)),\bar{\Gamma}_{1}=(\tau^{(1)}=(\tau^{(1)}_{a})_{a\in A},d^{(1)}), Γ¯2=(τ(2)=(τa(2))aA,d(2))\bar{\Gamma}_{2}=(\tau^{(2)}=(\tau^{(2)}_{a})_{a\in A},d^{(2)}) and Γ¯3=(τ(3)=(τa(3))aA,d(3)).\bar{\Gamma}_{3}=(\tau^{(3)}=(\tau^{(3)}_{a})_{a\in A},d^{(3)}). Define

σ(i):=(σa(i))aA with σa(i)(x):={τa(i)(x)if xT(d(i)),τa(i)(T(d(i)))if x>T(d(i)),x0aAi=1,2,3.\sigma^{(i)}:=(\sigma_{a}^{(i)})_{a\in A}\text{ with }\sigma^{(i)}_{a}(x):=\begin{cases}\tau^{(i)}_{a}(x)&\text{if }x\leq T(d^{(i)}),\\ \tau_{a}^{(i)}(T(d^{(i)}))&\text{if }x>T(d^{(i)}),\end{cases}\quad\forall x\in\mathbb{R}_{\geq 0}\ \forall a\in A\ \forall i=1,2,3.

Let Γ¯1=(σ(1),d(1)),\bar{\Gamma}^{\prime}_{1}=(\sigma^{(1)},d^{(1)}), Γ¯2=(σ(2),d(2))\bar{\Gamma}^{\prime}_{2}=(\sigma^{(2)},d^{(2)}) and Γ¯3=(σ(3),d(3)).\bar{\Gamma}^{\prime}_{3}=(\sigma^{(3)},d^{(3)}). Then Γ¯iΓ¯i\bar{\Gamma}_{i}\simeq\bar{\Gamma}^{\prime}_{i} for i=1,2,3.i=1,2,3.

Let Tmax:=max{T(d(1)),T(d(2)),T(d(3))}.T_{\max}:=\max\{T(d^{(1)}),T(d^{(2)}),T(d^{(3)})\}. Then

σ|T(d(i))(i)σ|T(d(j))(j)=maxaAmaxx[0,Tmax]|σa(i)(x)σa(j)(x)|=σ|Tmax(i)σ|Tmax(j)i,j=1,2,3.|\!|\sigma_{|T(d^{(i)})}^{(i)}-\sigma_{|T(d^{(j)})}^{(j)}|\!|_{\infty}=\max_{a\in A}\max_{x\in[0,T_{max}]}|\sigma_{a}^{(i)}(x)-\sigma_{a}^{(j)}(x)|=|\!|\sigma_{|T_{max}}^{(i)}-\sigma_{|T_{\max}}^{(j)}|\!|_{\infty}\quad\forall i,j=1,2,3.

Hence,

Dist(Γ¯i,Γ¯j)=max{d(i)d(j),σ|Tmax(i)σ|Tmax(j)}i,j=1,2,3.\text{Dist}(\bar{\Gamma}^{\prime}_{i},\bar{\Gamma}^{\prime}_{j})=\max\left\{|\!|d^{(i)}-d^{(j)}|\!|_{\infty},|\!|\sigma_{|T_{max}}^{(i)}-\sigma_{|T_{\max}}^{(j)}|\!|_{\infty}\right\}\quad\forall i,j=1,2,3.

The triangle inequality Dist(Γ¯1,Γ¯2)Dist(Γ¯1,Γ¯3)+Dist(Γ¯3,Γ¯2)\text{Dist}(\bar{\Gamma}^{\prime}_{1},\bar{\Gamma}^{\prime}_{2})\leq\text{Dist}(\bar{\Gamma}^{\prime}_{1},\bar{\Gamma}^{\prime}_{3})+\text{Dist}(\bar{\Gamma}^{\prime}_{3},\bar{\Gamma}^{\prime}_{2}) follows from (3.6).

d(1)d(2)d(1)d(3)+d(3)d(2)σ|Tmax(1)σ|Tmax(2)σ|Tmax(1)σ|Tmax(3)+σ|Tmax(3)σ|Tmax(2)\begin{split}&|\!|d^{(1)}-d^{(2)}|\!|_{\infty}\leq|\!|d^{(1)}-d^{(3)}|\!|_{\infty}+|\!|d^{(3)}-d^{(2)}|\!|_{\infty}\\ &|\!|\sigma_{|T_{max}}^{(1)}-\sigma_{|T_{\max}}^{(2)}|\!|_{\infty}\leq|\!|\sigma_{|T_{max}}^{(1)}-\sigma_{|T_{\max}}^{(3)}|\!|_{\infty}+|\!|\sigma_{|T_{max}}^{(3)}-\sigma_{|T_{\max}}^{(2)}|\!|_{\infty}\end{split} (3.6)

Lemma 4c) then follows since the operator Dist(,)\text{Dist}(\cdot,\cdot) is consistent with equivalence relation (3.1).

\square

Remark 3.1.

Note that we cannot substitute the cost function distance (3.4) in the metric (3.3) (equivalently, (3.5)) by the LL^{\infty}-norm

τ|Tmaxσ|Tmax:=maxaA,x[0,Tmax]|τa(x)σa(x)|,Tmax:=max{T(d),T(d)},|\!|\tau_{|T_{max}}-\sigma_{|T_{max}}|\!|_{\infty}:=\max_{a\in A,\ x\in[0,T_{max}]}|\tau_{a}(x)-\sigma_{a}(x)|,\quad T_{max}:=\max\{T(d),T(d^{\prime})\}, (3.7)

although this is more intuitive and has been applied to the auxiliary cost functions σ(i)\sigma^{(i)} in the proof of Lemma 4c). The reason is that the resulting binary operator

D(Γ¯,Γ¯)=max{dd,τ|Tmaxσ|Tmax},Γ¯,Γ¯𝒢¯(G,𝒦,𝒮),D(\bar{\Gamma}^{\prime},\bar{\Gamma})=\max\{|\!|d-d^{\prime}|\!|_{\infty},|\!|\tau_{|T_{max}}-\sigma_{|T_{max}}|\!|_{\infty}\},\quad\forall\bar{\Gamma},\bar{\Gamma}^{\prime}\in\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}), (3.8)

is inconsistent with equivalence relation (3.1), as D(Γ¯,Γ¯′′)=D(Γ¯,Γ¯′′)D(\bar{\Gamma}^{\prime},\bar{\Gamma}^{\prime\prime})=D(\bar{\Gamma},\bar{\Gamma}^{\prime\prime}) need not hold for three arbitrary generalized games Γ¯,Γ¯,\bar{\Gamma},\ \bar{\Gamma}^{\prime}, and Γ¯′′\bar{\Gamma}^{\prime\prime} with Γ¯Γ¯.\bar{\Gamma}\simeq\bar{\Gamma}^{\prime}. Moreover, D(,)D(\cdot,\cdot) does not fulfill the triangle inequality in Lemma 4c), and is thus neither a metric nor a pseudo-metric and so does not induce a metric space. To see this, we consider three arbitrary generalized games Γ¯=(τ,d),Γ¯=(σ,d),Γ¯′′=(σ,d)𝒢¯(G,𝒦,𝒮)\bar{\Gamma}=(\tau,d),\ \bar{\Gamma}^{\prime}=(\sigma,d),\ \bar{\Gamma}^{\prime\prime}=(\sigma,d^{\prime})\in\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}) such that T(d)>T(d)>0,T(d^{\prime})>T(d)>0, τa(x)=σa(x)\tau_{a}(x)=\sigma_{a}(x) for all (a,x)[0,T(d)],(a,x)\in[0,T(d)], and τb(y)σb(y)\tau_{b}(y)\neq\sigma_{b}(y) for all (a,y)(T(d),T(d)].(a,y)\in(T(d),T(d^{\prime})]. Then D(Γ¯,Γ¯)=0D(\bar{\Gamma},\bar{\Gamma}^{\prime})=0 and D(Γ¯,Γ¯′′)=dd′′.D(\bar{\Gamma}^{\prime},\bar{\Gamma}^{\prime\prime})=|\!|d-d^{\prime\prime}|\!|_{\infty}. When τa()\tau_{a}(\cdot) and σa()\sigma_{a}(\cdot) differ more than dd|\!|d-d^{\prime}|\!|_{\infty} on the non-empty interval (T(d),T(d)],(T(d),T(d^{\prime})], then D(Γ¯,Γ¯′′)>dd=D(Γ¯,Γ¯)+D(Γ¯,Γ¯′′).D(\bar{\Gamma},\bar{\Gamma}^{\prime\prime})>|\!|d-d^{\prime}|\!|_{\infty}=D(\bar{\Gamma},\bar{\Gamma}^{\prime})+D(\bar{\Gamma}^{\prime},\bar{\Gamma}^{\prime\prime}). This follows since the binary operator (3.8) does not distinguish the cost functions of the two generalized games Γ¯=(σ,d)\bar{\Gamma}^{\prime}=(\sigma,d) and Γ¯′′=(σ,d)\bar{\Gamma}^{\prime\prime}=(\sigma,d^{\prime}) when (3.7) is used to quantify the cost function distance. Hence, the binary operator (3.8) does not fulfill the triangle inequality. Our definition (3.4) of the distance of cost functions takes also the domains of cost functions into account. So, Γ¯\bar{\Gamma}^{\prime} and Γ¯′′\bar{\Gamma}^{\prime\prime} have different cost functions under our definition, and Dist(Γ¯,Γ¯)+Dist(Γ¯,Γ¯′′)=Dist(Γ¯,Γ¯′′)=max{dd,σ(T(d))σ(T(d))}=Dist(Γ¯,Γ¯′′).\text{Dist}(\bar{\Gamma},\bar{\Gamma}^{\prime})+\text{Dist}(\bar{\Gamma}^{\prime},\bar{\Gamma}^{\prime\prime})=\text{Dist}(\bar{\Gamma}^{\prime},\bar{\Gamma}^{\prime\prime})=\max\{|\!|d-d^{\prime}|\!|_{\infty},|\!|\sigma(T(d))-\sigma(T(d^{\prime}))|\!|_{\infty}\}=\text{Dist}(\bar{\Gamma},\bar{\Gamma}^{\prime\prime}).

Equipped with the metric (3.3), 𝒢¯(G,𝒦,𝒮)\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}) becomes a metric space with the topology generated by open ϵ\epsilon-balls of the form (3.9),

Bϵ(Γ¯):={Γ¯𝒢¯(G,𝒦,𝒮):Γ¯Γ¯=Dist(Γ¯,Γ¯)<ϵ},Γ¯𝒢¯(G,𝒦,𝒮),ϵ>0.B_{\epsilon}(\bar{\Gamma}):=\{\bar{\Gamma}^{\prime}\in\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}):\ |\!|\bar{\Gamma}-\bar{\Gamma}^{\prime}|\!|=\text{Dist}(\bar{\Gamma},\bar{\Gamma}^{\prime})<\epsilon\},\quad\bar{\Gamma}\in\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}),\ \epsilon>0. (3.9)

3.3 The PoA is continuous

The metric (3.3) induces the definition of convergence of games and of the continuity of real-valued maps on 𝒢¯(G,𝒦,𝒮)\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}). A sequence (Γ¯n)n𝒢¯(G,𝒦,𝒮)(\bar{\Gamma}_{n})_{n\in\mathbb{N}}\in\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S})^{\mathbb{N}} converges to a limit Γ¯𝒢¯(G,𝒦,𝒮),\bar{\Gamma}\in\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}), denoted by limnΓ¯n=Γ¯,\lim_{n\to\infty}\bar{\Gamma}_{n}=\bar{\Gamma}, if for each ϵ>0,\epsilon>0, there is an N>0N>0 such that Γ¯nBϵ(Γ¯)\bar{\Gamma}_{n}\in B_{\epsilon}(\bar{\Gamma}) for all nN.n\geq N. Trivially, limnΓ¯n=Γ¯\lim_{n\to\infty}\bar{\Gamma}_{n}=\bar{\Gamma} if and only if limnΓ¯nΓ¯=0.\lim_{n\to\infty}|\!|\bar{\Gamma}_{n}-\bar{\Gamma}|\!|=0. Then a real-valued map φ¯:𝒢¯(G,𝒦,𝒮)\bar{\varphi}:\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S})\to\mathbb{R} is continuous if limnφ¯(Γ¯n)=φ¯(Γ¯)\lim_{n\to\infty}\bar{\varphi}(\bar{\Gamma}_{n})=\bar{\varphi}(\bar{\Gamma}) for each sequence (Γ¯n)n𝒢¯(G,𝒦,𝒮)(\bar{\Gamma}_{n})_{n\in\mathbb{N}}\in\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S})^{\mathbb{N}} and each Γ¯𝒢¯(G,𝒦,𝒮)\bar{\Gamma}\in\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}) with limnΓ¯n=Γ¯.\lim_{n\to\infty}\bar{\Gamma}_{n}=\bar{\Gamma}. In addition, a real-valued map φ:𝒢(G,𝒦,𝒮)\varphi:\mathcal{G}(G,\mathcal{K},\mathcal{S})\to\mathbb{R} is continuous if limnφ(Γn)=φ(Γ)\lim_{n\to\infty}\varphi(\Gamma_{n})=\varphi(\Gamma) for each sequence (Γn)n𝒢(G,𝒦,𝒮)(\Gamma_{n})_{n\in\mathbb{N}}\in\mathcal{G}(G,\mathcal{K},\mathcal{S})^{\mathbb{N}} and each Γ𝒢¯(G,𝒦,𝒮)\Gamma\in\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}) with limnΓn=Γ.\lim_{n\to\infty}\Gamma_{n}=\Gamma.

Note that every game Γ=(τ,d)𝒢(G,𝒦,𝒮)\Gamma=(\tau,d)\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) has the unique total cost C(Γ,f)C(\Gamma,f^{*}) for its SO flows ff^{*}. This defines an SO cost map 𝒞:𝒢(G,𝒦,𝒮)0\mathcal{C}^{*}:\mathcal{G}(G,\mathcal{K},\mathcal{S})\to\mathbb{R}_{\geq 0} with 𝒞(Γ):=C(Γ,f)\mathcal{C}^{*}(\Gamma):=C(\Gamma,f^{*}) for each Γ𝒢(G,𝒦,𝒮).\Gamma\in\mathcal{G}(G,\mathcal{K},\mathcal{S}). Similarly, we can define a WE cost map 𝒞~:𝒢(G,𝒦,𝒮)0\tilde{\mathcal{C}}:\mathcal{G}(G,\mathcal{K},\mathcal{S})\to\mathbb{R}_{\geq 0} by putting 𝒞~(Γ):=C(Γ,f~)\tilde{\mathcal{C}}(\Gamma):=C(\Gamma,\tilde{f}) for each Γ𝒢(G,𝒦,𝒮),\Gamma\in\mathcal{G}(G,\mathcal{K},\mathcal{S}), where f~\tilde{f} is an arbitrary WE flow of Γ.\Gamma. Lemma 5b)–c) imply that both 𝒞()\mathcal{C}^{*}(\cdot) and 𝒞~()\tilde{\mathcal{C}}(\cdot) are continuous maps on 𝒢(G,𝒦,𝒮)\mathcal{G}(G,\mathcal{K},\mathcal{S}).

Lemma 5.

Consider a convergent game sequence (Γn=(τ(n),d(n)))n𝒢(G,𝒦,𝒮)(\Gamma_{n}=(\tau^{(n)},d^{(n)}))_{n\in\mathbb{N}}\in\mathcal{G}(G,\mathcal{K},\mathcal{S})^{\mathbb{N}} with limnΓn=Γ\lim_{n\to\infty}\Gamma_{n}=\Gamma for a game Γ=(τ,d)𝒢(G,𝒦,𝒮).\Gamma=(\tau,d)\in\mathcal{G}(G,\mathcal{K},\mathcal{S}).

  • a)

    Let ff be an arbitrary flow of Γ.\Gamma. Then there is a sequence (fn)n(f_{n})_{n\in\mathbb{N}} of flows such that each fn,f_{n}, n,n\in\mathbb{N}, is a flow of Γn\Gamma_{n}, and f=limnf(n).f=\lim_{n\to\infty}f^{(n)}.

  • b)

    Let (fn(n))n(f^{*(n)}_{n})_{n\in\mathbb{N}} be a sequence of SO flows f(n)f^{*(n)} of Γn,\Gamma_{n}, and let (ni)i(n_{i})_{i\in\mathbb{N}} be an infinite subsequence with limif(ni)=f\lim_{i\to\infty}f^{*(n_{i})}=f^{*} for a vector f0𝒮.f^{*}\in\mathbb{R}_{\geq 0}^{\mathcal{S}}. Then ff^{*} is an SO flow of Γ.\Gamma. Moreover, (fn)n(f^{*}_{n})_{n\in\mathbb{N}} converges to an SO flow of Γ\Gamma when Γ\Gamma has a unique SO flow. In particular, limn𝒞(Γn)=𝒞(Γ).\lim_{n\to\infty}\mathcal{C}^{*}(\Gamma_{n})=\mathcal{C}^{*}(\Gamma).

  • c)

    Let (f~n)n(\tilde{f}_{n})_{n\in\mathbb{N}} be a sequence of WE flows f~(n)\tilde{f}^{(n)} of Γn,\Gamma_{n}, and let (ni)i(n_{i})_{i\in\mathbb{N}} be an infinite subsequence with limif~(ni)=f~\lim_{i\to\infty}\tilde{f}^{(n_{i})}=\tilde{f} for a vector f~0𝒮.\tilde{f}\in\mathbb{R}_{\geq 0}^{\mathcal{S}}. Then f~\tilde{f} is a WE flow of Γ.\Gamma. Moreover, (f~n)n(\tilde{f}_{n})_{n\in\mathbb{N}} converges to a WE flow of Γ\Gamma when Γ\Gamma has a unique WE flow. In particular, limn𝒞~(Γn)=𝒞~(Γ).\lim_{n\to\infty}\tilde{\mathcal{C}}(\Gamma_{n})=\tilde{\mathcal{C}}(\Gamma).

Proof.

Proof of Lemma 5

Proof.

Proof of Lemma 5a): Define δ:=mins𝒮:fs>0fs\delta:=\min_{s\in\mathcal{S}:f_{s}>0}f_{s} and sk:=argmins𝒮k:fs>0fss_{k}:=\arg\min_{s\in\mathcal{S}_{k}:f_{s}>0}f_{s} for each k𝒦.k\in\mathcal{K}. Then fsδ>0f_{s}\geq\delta>0 for all s𝒮s\in\mathcal{S} with fs>0,f_{s}>0, and fsfskδf_{s^{\prime}}\geq f_{s_{k}}\geq\delta for each s𝒮ks^{\prime}\in\mathcal{S}_{k} with fs>0f_{s^{\prime}}>0 for each O/D pair k𝒦.k\in\mathcal{K}. Since limnΓn=Γ,\lim_{n\to\infty}\Gamma_{n}=\Gamma, we have δ>ΓnΓd(n)d=maxk𝒦|dk(n)dk|\delta>|\!|\Gamma_{n}-\Gamma|\!|\geq|\!|d^{(n)}-d|\!|_{\infty}=\max_{k\in\mathcal{K}}|d_{k}^{(n)}-d_{k}| for each nNn\geq N for some integer N.N\in\mathbb{N}.

Define for each nNn\geq N a vector f(n)=(fs(n))s𝒮f^{(n)}=(f^{(n)}_{s})_{s\in\mathcal{S}} with

fs(n):={fsif s𝒮k{sk},fs+dk(n)dkif s=sk,s𝒮kk𝒦.f_{s}^{(n)}:=\begin{cases}f_{s}&\text{if }s\in\mathcal{S}_{k}\setminus\{s_{k}\},\\ f_{s}+d_{k}^{(n)}-d_{k}&\text{if }s=s_{k},\end{cases}\quad\forall s\in\mathcal{S}_{k}\ \forall k\in\mathcal{K}.

Then fs(n)0f_{s}^{(n)}\geq 0 for each s𝒮,s\in\mathcal{S}, since δ+dk(n)dk0\delta+d_{k}^{(n)}-d_{k}\geq 0 for each O/D pair k𝒦k\in\mathcal{K} when nN.n\geq N. Moreover, sSkfs(n)=sSkfs+dk(n)dk=dk(n)\sum_{s\in S_{k}}f_{s}^{(n)}=\sum_{s\in S_{k}}f_{s}+d_{k}^{(n)}-d_{k}=d_{k}^{(n)} for each O/D pair k𝒦.k\in\mathcal{K}. So f(n)f^{(n)} is a flow of Γn\Gamma_{n} for each nN.n\geq N.

limnf(n)=f\lim_{n\to\infty}f^{(n)}=f follows immediately from the definition of f(n)f^{(n)} and the fact that limnΓn=Γ.\lim_{n\to\infty}\Gamma_{n}=\Gamma.

Proof.

Proof of Lemma 5b): Trivially, ff^{*} is a flow of Γ.\Gamma. Let ff be an arbitrary flow of Γ.\Gamma. Lemma 5a) implies that f=limnf(n)f=\lim_{n\to\infty}f^{(n)} for a sequence (f(n))n(f^{(n)})_{n\in\mathbb{N}} of flows of games Γn.\Gamma_{n}. Then we obtain by the continuity of cost functions that C(Γ,f)=limiC(Γni,f(ni))limiC(Γni,f(ni))=C(Γ,f).C(\Gamma,f^{*})=\lim_{i\to\infty}C(\Gamma_{n_{i}},f^{*(n_{i})})\leq\lim_{i\to\infty}C(\Gamma_{n_{i}},f^{(n_{i})})=C(\Gamma,f). This shows that ff^{*} is an SO flow of Γ\Gamma due to the arbitrary choice of f.f. The remainder of Lemma 5b) then follows trivially.

Proof.

Proof of Lemma 5c): Similarly, f~\tilde{f} is a flow of Γ.\Gamma. Consider an arbitrary O/D pair k𝒦k\in\mathcal{K} and arbitrary two paths s,s𝒮ks,s^{\prime}\in\mathcal{S}_{k} with f~s>0.\tilde{f}_{s}>0. Then f~s(ni)>0\tilde{f}^{(n_{i})}_{s}>0 when ii is large enough. Since f~(ni)\tilde{f}^{(n_{i})} is a WE flow of Γni\Gamma_{n_{i}} for each i,i\in\mathbb{N}, we have τs(f~)=limiτs(ni)(f~(ni))limiτs(ni)(f~(ni))=τs(f~).\tau_{s}(\tilde{f})=\lim_{i\to\infty}\tau_{s}^{(n_{i})}(\tilde{f}^{(n_{i})})\leq\lim_{i\to\infty}\tau_{s^{\prime}}^{(n_{i})}(\tilde{f}^{(n_{i})})=\tau_{s^{\prime}}(\tilde{f}). This shows that f~\tilde{f} is a WE flow of Γ\Gamma due to the arbitrary choice of k,sk,s and s.s^{\prime}. The remainder of Lemma 5c) then follows trivially.

\square

Viewed as a real-valued map from 𝒢(G,𝒦,𝒮)\mathcal{G}(G,\mathcal{K},\mathcal{S}) to 0,\mathbb{R}_{\geq 0}, the PoA ρ()\rho(\cdot) is also continuous, since ρ()\rho(\cdot) is the quotient of two continuous maps on 𝒢(G,𝒦,𝒮)\mathcal{G}(G,\mathcal{K},\mathcal{S}), i.e., ρ(Γ)=𝒞~(Γ)𝒞(Γ)\rho(\Gamma)=\frac{\tilde{\mathcal{C}}(\Gamma)}{\mathcal{C}^{*}(\Gamma)} for each Γ𝒢(G,𝒦,𝒮).\Gamma\in\mathcal{G}(G,\mathcal{K},\mathcal{S}). Here, we recall that the PoA ρ()\rho(\cdot) is well defined on the whole game space 𝒢(G,𝒦,𝒮).\mathcal{G}(G,\mathcal{K},\mathcal{S}).

We summarize all these continuity results in Theorem 6.

Theorem 6.

The SO cost map 𝒞(),\mathcal{C}^{*}(\cdot), the WE cost map 𝒞~()\tilde{\mathcal{C}}(\cdot) and the PoA map ρ()\rho(\cdot) are continuous on the game space 𝒢(G,𝒦,𝒮)\mathcal{G}(G,\mathcal{K},\mathcal{S}).

Note that Hall [1978] has proved that the user cost Lk(τ,d)=mins𝒮k:f~s>0τs(f~)L_{k}(\tau,d)=\min_{s\in\mathcal{S}_{k}:\ \tilde{f}_{s}>0}\tau_{s}(\tilde{f}) is a continuous function of the demand vector dd when the cost function vector τ\tau is fixed. This then implies directly that 𝒞~()\tilde{\mathcal{C}}(\cdot) is continuous when τ\tau is fixed. Theorem 6 generalizes this continuity result to the game space 𝒢(G,𝒦,𝒮).\mathcal{G}(G,\mathcal{K},\mathcal{S}).

Non-atomic congestion games Γ¯\bar{\Gamma} in the gap 𝒢¯(G,𝒦,𝒮)𝒢(G,𝒦,𝒮)\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S})\setminus\mathcal{G}(G,\mathcal{K},\mathcal{S}) may have an undefined PoA of 00,\frac{0}{0}, and are thus not considered in our sequel analysis of the PoA map ρ()\rho(\cdot). One may of course wonder if we could include them in the analysis by constructing an extension ρ¯:𝒢¯(G,𝒦,𝒮)0\bar{\rho}:\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S})\to\mathbb{R}_{\geq 0} of the PoA map ρ()\rho(\cdot) to 𝒢¯(G,𝒦,𝒮).\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}). From a topological point of view, such an extension ρ¯()\bar{\rho}(\cdot) should not only satisfy the condition that ρ¯(Γ)=ρ(Γ)\bar{\rho}(\Gamma)=\rho(\Gamma) for each Γ𝒢(G,𝒦,𝒮),\Gamma\in\mathcal{G}(G,\mathcal{K},\mathcal{S}), but also preserve the continuity of the PoA map ρ().\rho(\cdot). When such a map ρ¯()\bar{\rho}(\cdot) exists, Condition 2 would then no longer be needed, which would considerably simplify the further analysis. Unfortunately, Example 7 illustrates that the PoA ρ()\rho(\cdot) cannot be continuously extended to the generalized game space 𝒢¯(G,𝒦,𝒮).\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}). Thus we need to exclude generalized games with an undefined PoA of 00\frac{0}{0} and have to accept the existence of this gap.

Example 7.

Consider the traffic network GG shown in Figure 2(a)–(b). This network has two vertices oo and tt with two parallel paths (arcs). Denote by 𝒦\mathcal{K} and 𝒮\mathcal{S} the respective sets of O/D pairs and paths of G.G. Let Γ¯=(τ,d)𝒢¯(G,𝒦,𝒮)\bar{\Gamma}=(\tau,d)\in\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}) be a generalized game with total demand T(d)=1T(d)=1 and the two cost functions τ1()\tau_{1}(\cdot) and τ2()\tau_{2}(\cdot) shown in Figure 2(a). For each n,n\in\mathbb{N}, let Γn=(τ(n),d(n))𝒢(G,𝒦,𝒮)\Gamma_{n}=(\tau^{(n)},d^{(n)})\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) be a game again with the same total demand T(d(n))=1T(d^{(n)})=1 but with the two cost functions τ1(n)()\tau_{1}^{(n)}(\cdot) and τ2(n)()\tau_{2}^{(n)}(\cdot) shown in Figure 2(b). The game sequence (Γn)n(\Gamma_{n})_{n\in\mathbb{N}} converges to Γ¯\bar{\Gamma} for every choice of β[0,)\beta\in[0,\infty).

ooτ1(x)0\tau_{1}(x)\equiv 0τ2(x)0\tau_{2}(x)\equiv 0tt
(a) Game Γ¯𝒢¯(G,𝒦,𝒮)\bar{\Gamma}\in\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}) with T(d)=1T(d)=1
ooτ1(n)(x)1n\tau_{1}^{(n)}(x)\equiv\frac{1}{n}τ2(n)(x)=xβn\tau_{2}^{(n)}(x)=\frac{x^{\beta}}{n}tt
(b) Game Γ(n)𝒢(G,𝒦,𝒮)\Gamma^{(n)}\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) with T(d(n))=1T(d^{(n)})=1
Figure 2: Non-extensibility of the PoA

We now illustrate with this convergent sequence that the PoA ρ()\rho(\cdot) can not be continuously extended to 𝒢¯(G,𝒦,𝒮).\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}). We do this by contradiction, and thus assume that there is a continuous extension ρ¯:𝒢¯(G,𝒦,𝒮)0\bar{\rho}:\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S})\to\mathbb{R}_{\geq 0} of the PoA ρ().\rho(\cdot). Then ρ¯(Γn)=ρ(Γn)\bar{\rho}(\Gamma_{n})=\rho(\Gamma_{n}) for each n,n\in\mathbb{N}, and ρ¯(Γ¯)=limnρ¯(Γn)=limnρ(Γn),\bar{\rho}(\bar{\Gamma})=\lim_{n\to\infty}\bar{\rho}(\Gamma_{n})=\lim_{n\to\infty}\rho(\Gamma_{n}), since ρ¯\bar{\rho} is continuous on 𝒢¯(G,𝒦,𝒮)\overline{\mathcal{G}}(G,\mathcal{K},\mathcal{S}). This means that the sequence (ρ(Γn))n=(ρ¯(Γn))n(\rho(\Gamma_{n}))_{n\in\mathbb{N}}=(\bar{\rho}(\Gamma_{n}))_{n\in\mathbb{N}} has a unique limit ρ¯(Γ¯)\bar{\rho}(\bar{\Gamma}) that is independent of β\beta. However, the limit of (ρ(Γn))n=(ρ¯(Γn))n(\rho(\Gamma_{n}))_{n\in\mathbb{N}}=(\bar{\rho}(\Gamma_{n}))_{n\in\mathbb{N}} depends crucially on the value β,\beta, and yields limnρ¯(Γn)=limnρ(Γn)=1\lim_{n\to\infty}\bar{\rho}(\Gamma_{n})=\lim_{n\to\infty}\rho(\Gamma_{n})=1 when β=0,\beta=0, and limnρ¯(Γn)=limnρ(Γn)=43\lim_{n\to\infty}\bar{\rho}(\Gamma_{n})=\lim_{n\to\infty}\rho(\Gamma_{n})=\frac{4}{3} when β=1.\beta=1. Hence, there is no continuous extension ρ¯()\bar{\rho}(\cdot) of the PoA ρ()\rho(\cdot).

3.4 Hölder continuous maps

Theorem 6 implies that |ρ(Γ)ρ(Γ)||\rho(\Gamma)-\rho(\Gamma^{\prime})| is small when the game Γ𝒢(G,𝒦,𝒮)\Gamma^{\prime}\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) deviates only slightly from the game Γ𝒢(G,𝒦,𝒮)\Gamma\in\mathcal{G}(G,\mathcal{K},\mathcal{S}). Section 4 below will further quantify the difference |ρ(Γ)ρ(Γ)||\rho(\Gamma)-\rho(\Gamma^{\prime})| of the PoA in terms of the metric ΓΓ|\!|\Gamma-\Gamma^{\prime}|\!|. To that end, we need the notion of Hölder continuity.

Definition 3.1.

Consider a real-valued map φ:𝒢(G,𝒦,𝒮).\varphi:\mathcal{G}(G,\mathcal{K},\mathcal{S})\to\mathbb{R}.

  • i)i)

    The map φ\varphi is pointwise Hölder continuous at a game Γ𝒢(G,𝒦,𝒮)\Gamma\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) with a Hölder exponent γΓ>0\gamma_{\Gamma}>0 (depending only on Γ\Gamma), if there are a Hölder constant ϰΓ>0\varkappa_{\Gamma}>0 (depending also only on Γ\Gamma) and a non-empty open set UΓ𝒢(G,𝒦,𝒮),U_{\Gamma}\subseteq\mathcal{G}(G,\mathcal{K},\mathcal{S}), s.t., ΓUΓ\Gamma\in U_{\Gamma} and |φ(Γ)φ(Γ)|ϰΓΓΓγΓ|\varphi(\Gamma)-\varphi(\Gamma^{\prime})|\leq\varkappa_{\Gamma}\cdot|\!|\Gamma-\Gamma^{\prime}|\!|^{\gamma_{\Gamma}} for each ΓUΓ.\Gamma^{\prime}\in U_{\Gamma}.

  • ii)ii)

    The map φ\varphi is pointwise Hölder continuous on the whole game space 𝒢(G,𝒦,𝒮)\mathcal{G}(G,\mathcal{K},\mathcal{S}) with a (uniform) Hölder exponent γ>0\gamma>0 when it is pointwise Hölder continuous at every Γ𝒢(G,𝒦,𝒮)\Gamma\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) with the Hölder exponent γ\gamma.

  • iii)iii)

    The map φ\varphi is uniformly Hölder continuous on the whole game space 𝒢(G,𝒦,𝒮)\mathcal{G}(G,\mathcal{K},\mathcal{S}) with a (uniform) Hölder exponent γ>0\gamma>0 and a (uniform) Hölder constant ϰ>0\varkappa>0 if |φ(Γ)φ(Γ)|ϰΓΓγ|\varphi(\Gamma)-\varphi(\Gamma^{\prime})|\leq\varkappa\cdot|\!|\Gamma-\Gamma^{\prime}|\!|^{\gamma} for all Γ,Γ𝒢(G,𝒦,𝒮)\Gamma^{\prime},\Gamma\in\mathcal{G}(G,\mathcal{K},\mathcal{S}).

  • iv)iv)

    The map φ\varphi is Lipschitz continuous on the whole game space 𝒢(G,𝒦,𝒮)\mathcal{G}(G,\mathcal{K},\mathcal{S}) with a Lipschitz constant ϰ>0\varkappa>0 when φ\varphi is uniformly Hölder continuous on 𝒢(G,𝒦,𝒮)\mathcal{G}(G,\mathcal{K},\mathcal{S}) with the Hölder exponent γ=1\gamma=1 and the Hölder constant ϰ\varkappa.

Uniform Hölder continuity implies pointwise Hölder continuity, but the converse need not hold, since the game space 𝒢(G,𝒦,𝒮)\mathcal{G}(G,\mathcal{K},\mathcal{S}) is not compact. Moreover, the smaller the Hölder exponent γΓ\gamma_{\Gamma} of the PoA map ρ()\rho(\cdot) at a game Γ𝒢(G,𝒦,𝒮)\Gamma\in\mathcal{G}(G,\mathcal{K},\mathcal{S}), the more sensitive is the PoA of the game Γ\Gamma w.r.t. small changes in its demands and cost functions. Hence, we can indeed quantify the sensitivity of the PoA by a Hölder continuity analysis of the map ρ()\rho(\cdot).

3.5 ρ\rho-invariant operators

The Hölder continuity analysis in Section 4 also involves the notion of ρ\rho-invariant operators. Formally, a continuous map Υ:𝒢(G,𝒦,𝒮)𝒢(G,𝒦,𝒮)\Upsilon:\mathcal{G}(G,\mathcal{K},\mathcal{S})\to\mathcal{G}(G,\mathcal{K},\mathcal{S}) is called a ρ\rho-invariant operator, if it is continuous and does not change the PoA. i.e., ρ(Γ)=ρ(Υ(Γ))\rho(\Gamma)=\rho\big{(}\Upsilon(\Gamma)\big{)} for each game Γ𝒢(G,𝒦,𝒮).\Gamma\in\mathcal{G}(G,\mathcal{K},\mathcal{S}). Examples are the cost and demand normalizations that have been used by Colini-Baldeschi et al. [2017, 2020] and by Wu et al. [2021].

A cost normalization is an operator Ψ:𝒢(G,𝒦,𝒮)𝒢(G,𝒦,𝒮)\Psi:\mathcal{G}(G,\mathcal{K},\mathcal{S})\to\mathcal{G}(G,\mathcal{K},\mathcal{S}) with Ψ(Γ)=Ψ(τ,d)=(τυ,d)𝒢(G,𝒦,𝒮)\Psi(\Gamma)=\Psi(\tau,d)=(\frac{\tau}{\upsilon},d)\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) for each Γ=(τ,d)𝒢(G,𝒦,𝒮),\Gamma=(\tau,d)\in\mathcal{G}(G,\mathcal{K},\mathcal{S}), where υ>0\upsilon>0 is a constant factor and τυ:=(τaυ)aA.\frac{\tau}{\upsilon}:=(\frac{\tau_{a}}{\upsilon})_{a\in A}. We employ the notation Ψυ\Psi_{\upsilon} to denote a cost normalization with factor υ.\upsilon. Cost normalizations Ψυ()\Psi_{\upsilon}(\cdot) are continuous since limnΨυ(Γn)=Ψυ(Γ)\lim_{n\to\infty}\Psi_{\upsilon}(\Gamma_{n})=\Psi_{\upsilon}(\Gamma) when limnΓnΓ=0.\lim_{n\to\infty}|\!|\Gamma_{n}-\Gamma|\!|=0. As the PoA map ρ()\rho(\cdot) is the quotient of the WE cost map 𝒞~()\tilde{\mathcal{C}}(\cdot) over the SO cost map 𝒞()\mathcal{C}^{*}(\cdot), ρ()\rho(\cdot) is then invariant w.r.t. arbitrary cost normalizations Ψυ()\Psi_{\upsilon}(\cdot).

However, a cost normalization does not leave the distance invariant. In fact, we have

Ψυ(Γ)Ψυ(Γ)=max{dd,τ|T(d)σ|T(d)υ}={ddif ddτ|T(d)σ|T(d)υ,τ|T(d)σ|T(d)υif dd<τ|T(d)σ|T(d)υ,\begin{split}|\!|\Psi_{\upsilon}(\Gamma)-\Psi_{\upsilon}(\Gamma^{\prime})|\!|&=\max\left\{|\!|d-d^{\prime}|\!|_{\infty},\frac{|\!|\tau_{|T(d)}-\sigma_{|T(d^{\prime})}|\!|_{\infty}}{\upsilon}\right\}\\ &=\begin{cases}|\!|d-d^{\prime}|\!|_{\infty}&\text{if }|\!|d-d^{\prime}|\!|_{\infty}\geq\frac{|\!|\tau_{|T(d)}-\sigma_{|T(d^{\prime})}|\!|_{\infty}}{\upsilon},\\ \frac{|\!|\tau_{|T(d)}-\sigma_{|T(d^{\prime})}|\!|_{\infty}}{\upsilon}&\text{if }|\!|d-d^{\prime}|\!|_{\infty}<\frac{|\!|\tau_{|T(d)}-\sigma_{|T(d^{\prime})}|\!|_{\infty}}{\upsilon},\end{cases}\end{split} (3.10)

for all Γ=(τ,d),Γ=(σ,d)𝒢(G,𝒦,𝒮)\Gamma=(\tau,d),\Gamma^{\prime}=(\sigma,d^{\prime})\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) and all υ>0,\upsilon>0, which may be different from ΓΓ.|\!|\Gamma-\Gamma^{\prime}|\!|.

By (3.10), Ψυ(Γ)Ψυ(Γ)=ΓΓ=dd|\!|\Psi_{\upsilon}(\Gamma)-\Psi_{\upsilon}(\Gamma^{\prime})|\!|=|\!|\Gamma-\Gamma^{\prime}|\!|=|\!|d-d^{\prime}|\!|_{\infty} when

ddmax{τ|T(d)σ|T(d),τ|T(d)σ|T(d)υ}.|\!|d-d^{\prime}|\!|_{\infty}\geq\max\left\{|\!|\tau_{|T(d)}-\sigma_{|T(d^{\prime})}|\!|_{\infty},\ \frac{|\!|\tau_{|T(d)}-\sigma_{|T(d^{\prime})}|\!|_{\infty}}{\upsilon}\right\}.

In particular, a cost normalization Ψυ\Psi_{\upsilon} will result in a scaling of the distance when d=d,d=d^{\prime}, i.e.,

Ψυ(Γ)Ψυ(Γ)=ΓΓυΓ=(τ,d),Γ=(σ,d)𝒢(G,𝒦,𝒮) with d=d.|\!|\Psi_{\upsilon}(\Gamma)-\Psi_{\upsilon}(\Gamma^{\prime})|\!|=\frac{|\!|\Gamma-\Gamma^{\prime}|\!|}{\upsilon}\quad\forall\Gamma=(\tau,d),\ \Gamma^{\prime}=(\sigma,d^{\prime})\in\mathcal{G}(G,\mathcal{K},\mathcal{S})\text{ with }d=d^{\prime}. (3.11)

We will see in Section 4 that equation (3.11) implies a rather unpleasant property of the map ρ().\rho(\cdot).

A demand normalization is an operator Λ:𝒢(G,𝒦,𝒮)𝒢(G,𝒦,𝒮)\Lambda:\mathcal{G}(G,\mathcal{K},\mathcal{S})\to\mathcal{G}(G,\mathcal{K},\mathcal{S}) with Λ(Γ)=Λ(τ,d)=(τυ,dυ)𝒢(G,𝒦,𝒮),\Lambda(\Gamma)=\Lambda(\tau,d)=(\tau\circ\upsilon,\frac{d}{\upsilon})\in\mathcal{G}(G,\mathcal{K},\mathcal{S}), where υ>0\upsilon>0 is again a constant factor, dυ:=(dkυ)k𝒦\frac{d}{\upsilon}:=(\frac{d_{k}}{\upsilon})_{k\in\mathcal{K}} and τυ:=(τaυ)aA\tau\circ\upsilon:=(\tau_{a}\circ\upsilon)_{a\in A} with τaυ(x):=τa(xυ)\tau_{a}\circ\upsilon(x):=\tau_{a}(x\cdot\upsilon) for all aAa\in A and all x[0,T(d)υ].x\in[0,\frac{T(d)}{\upsilon}]. We employ the notation Λυ\Lambda_{\upsilon} to denote a demand normalization with factor υ>0.\upsilon>0. Similar to cost normalizations, demand normalizations are continuous. The PoA ρ()\rho(\cdot) is invariant under demand normalizations, since f~\tilde{f} and ff^{*} are a WE flow and an SO flow of Γ\Gamma if and only if f~υ\frac{\tilde{f}}{\upsilon} and fυ\frac{f^{*}}{\upsilon} are a WE flow and an SO flow of Λυ(Γ),\Lambda_{\upsilon}(\Gamma), respectively, and since C(Γ,f)=υC(Λυ(Γ),fυ)C(\Gamma,f)=\upsilon\cdot C(\Lambda_{\upsilon}(\Gamma),\frac{f}{\upsilon}) for an arbitrary flow ff of Γ.\Gamma.

We will demonstrate in Section 5 that the cost and demand normalizations also help to analyze the convergence rate of the PoA when the total demand tends to 0 or .\infty.

4 Hölder continuity of the PoA

Initial Hölder continuity results have been obtained by Englert et al. [2010]; Takalloo and Kwon [2020] and Cominetti et al. [2020]. They considered the Hölder continuity of the PoA on subspaces of 𝒢(G,𝒦,𝒮),\mathcal{G}(G,\mathcal{K},\mathcal{S}), in which all games have the same cost functions.

Consider now an arbitrary cost function vector τ=(τa)aA\tau=(\tau_{a})_{a\in A} defined on [0,)[0,\infty) and satisfying Condition 2. We call the subspace 𝒢(G,𝒦,𝒮)|τ:={Γ=(σ,d)𝒢(G,𝒦,𝒮):σa(x)=τa(x)aAx[0,T(d)]}\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau}:=\{\Gamma^{\prime}=(\sigma,d)\in\mathcal{G}(G,\mathcal{K},\mathcal{S}):\sigma_{a}(x)=\tau_{a}(x)\ \forall a\in A\ \forall x\in[0,T(d)]\} of 𝒢(G,𝒦,𝒮)\mathcal{G}(G,\mathcal{K},\mathcal{S}) a cost slice w.r.t. the cost function vector τ.\tau. Trivially, two arbitrary games Γ1=(σ(1),d(1))\Gamma_{1}=(\sigma^{(1)},d^{(1)}) and Γ2=(σ(2),d(2))\Gamma_{2}=(\sigma^{(2)},d^{(2)}) from the same cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} satisfy that

Γ1Γ2=max{d(1)d(2),σ(1)(T(d(1)))σ(2)(T(d(2)))}.|\!|\Gamma_{1}-\Gamma_{2}|\!|=\max\left\{|\!|d^{(1)}-d^{(2)}|\!|_{\infty},\ |\!|\sigma^{(1)}(T(d^{(1)}))-\sigma^{(2)}(T(d^{(2)}))|\!|_{\infty}\right\}. (4.1)

Englert et al. [2010] considered the Hölder continuity of the PoA on a cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} with polynomial of degree at most β\beta cost functions τa()\tau_{a}(\cdot) on networks with only one O/D pair. They showed that

ρ(Γ)ρ(Γ)((1+ϵ)β1)ρ(Γ)\rho(\Gamma^{\prime})-\rho(\Gamma)\leq\big{(}(1+\epsilon)^{\beta}-1\big{)}\cdot\rho(\Gamma)

for two arbitrary games Γ=(τ,d)\Gamma=(\tau,d) and Γ=(τ,d)\Gamma^{\prime}=(\tau,d^{\prime}) of the cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} with d=(1+ϵ)dd^{\prime}=(1+\epsilon)\cdot d for an arbitrary constant ϵ>0\epsilon>0.

Takalloo and Kwon [2020] generalized the results of Englert et al. [2010] and considered the Hölder continuity of the PoA on a cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} with again polynomial cost functions τa()\tau_{a}(\cdot) of degree at most β,\beta, but for networks with multiple O/D pairs. They showed for this more general case that

O(ϵ)=(1(1+ϵ)β1)ρ(Γ)ρ(Γ)ρ(Γ)((1+ϵ)β1)ρ(Γ)=O(ϵ)-O(\epsilon)=\big{(}\frac{1}{(1+\epsilon)^{\beta}}-1\big{)}\cdot\rho(\Gamma)\leq\rho(\Gamma^{\prime})-\rho(\Gamma)\leq\big{(}(1+\epsilon)^{\beta}-1\big{)}\cdot\rho(\Gamma)=O(\epsilon) (4.2)

for two arbitrary games Γ=(τ,d)\Gamma=(\tau,d) and Γ=(τ,d)\Gamma^{\prime}=(\tau,d^{\prime}) of the cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} with d=(1+ϵ)dd^{\prime}=(1+\epsilon)\cdot d for an arbitrary constant ϵ>0\epsilon>0.

Their result implies that |ρ(Γ)ρ(Γ)|O(ρ(Γ)ϵ)|\rho(\Gamma)-\rho(\Gamma^{\prime})|\in O(\rho(\Gamma)\cdot\epsilon) when the cost functions τa()\tau_{a}(\cdot) are polynomials, and when Γ=(τ,d)\Gamma=(\tau,d) and Γ=(τ,(1+ϵ)d)\Gamma^{\prime}=(\tau,(1+\epsilon)\cdot d) for a constant ϵ>0,\epsilon>0, see (4.2). This together with (4.1) then yields that |ρ(Γ)ρ(Γ)|<ϰΓΓΓ|\rho(\Gamma)-\rho(\Gamma^{\prime})|<\varkappa_{\Gamma}\cdot|\!|\Gamma-\Gamma^{\prime}|\!| for a Hölder constant ϰΓ>0\varkappa_{\Gamma}>0 depending on Γ\Gamma when Γ=(τ,d)\Gamma=(\tau,d) and Γ=(τ,(1+ϵ)d)\Gamma^{\prime}=(\tau,(1+\epsilon)\cdot d) for a constant ϵ>0,\epsilon>0, so with a Hölder exponent of 11 though at the cost of the restrictive condition “d=(1+ϵ)dd^{\prime}=(1+\epsilon)\cdot d”. This result is quite inspiring and implies for each polynomial cost function vector τ\tau and each demand vector dd with T(d)>0T(d)>0 that the PoA map ρ()\rho(\cdot) is pointwise Hölder continuous with Hölder exponent 11 on the resulting one-dimensional affine subspace {Γ=(τ,d)𝒢(G,𝒦,𝒮)|τ:d=αd,α>0}\{\Gamma=(\tau,d^{\prime})\in\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau}:d^{\prime}=\alpha\cdot d,\ \alpha>0\} of the cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau}.

Cominetti et al. [2020] also considered the Hölder continuity on a cost slice 𝒢(G,𝒦,𝒮)|τ,\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau}, but on networks with one O/D pair and with cost functions τa()\tau_{a}(\cdot) that are affine linear or have strictly positive derivatives. Unlike Englert et al. [2010], Cominetti et al. [2020] focused on the differentiability of the resulting PoA map ρ()\rho(\cdot) on the cost slice 𝒢(G,𝒦,𝒮)|τ.\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau}. They showed for this case that the PoA map ρ()\rho(\cdot) is differentiable at each demand level except for a finite set of \mathcal{E}-breakpoints. This implies that the PoA map is pointwise Hölder continuous with Hölder exponent 11 on the cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} except for a finite set of points. This is the first relatively complete result on the Hölder continuity of the PoA as it needs no longer the condition “d=(1+ϵ)dd=(1+\epsilon)\cdot d^{\prime}”. But the restriction to one O/D pair and cost slice only is still strong.

We now generalize these results by analyzing the Hölder continuity of the PoA map on the whole game space 𝒢(G,𝒦,𝒮)\mathcal{G}(G,\mathcal{K},\mathcal{S}) and for an arbitrary finite set 𝒦\mathcal{K} of O/D pairs. Our Hölder continuity analysis is thus not restricted to a cost slice, but quantifies the changes of the PoA when both the cost functions and the demands change. As we consider the most general case, one cannot expect similar results on the differentiability of the PoA as in Cominetti et al. [2020].

4.1 The PoA is not uniformly Hölder continuous

We show first that the PoA map ρ()\rho(\cdot) is not uniformly Hölder continuous on 𝒢(G,𝒦,𝒮)\mathcal{G}(G,\mathcal{K},\mathcal{S}). This also means that the PoA map is not Lipschitz continuous on the whole game space 𝒢(G,𝒦,𝒮).\mathcal{G}(G,\mathcal{K},\mathcal{S}). We assume by contradiction that ρ()\rho(\cdot) is uniformly Hölder continuous on the whole game space 𝒢(G,𝒦,𝒮)\mathcal{G}(G,\mathcal{K},\mathcal{S}) with a uniform Hölder exponent γ>0\gamma>0 and a uniform Hölder constant ϰ>0,\varkappa>0, i.e., we assume for all Γ,Γ𝒢(G,𝒦,𝒮)\Gamma,\Gamma^{\prime}\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) that

|ρ(Γ)ρ(Γ)|ϰΓΓγ.|\rho(\Gamma)-\rho(\Gamma^{\prime})|\leq\varkappa\cdot|\!|\Gamma-\Gamma^{\prime}|\!|^{\gamma}. (4.3)

We now choose two arbitrary games Γ=(τ,d)\Gamma=(\tau,d) and Γ=(σ,d)\Gamma^{\prime}=(\sigma,d^{\prime}) with ρ(Γ)ρ(Γ)\rho(\Gamma)\neq\rho(\Gamma^{\prime}) and d=d.d=d^{\prime}. Note that there are indeed such two games in the demand slice 𝒢(G,𝒦,𝒮)|d:={(σ,d′′)𝒢(G,𝒦,𝒮):d′′=d},\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|d}:=\{(\sigma,d^{\prime\prime})\in\mathcal{G}(G,\mathcal{K},\mathcal{S}):d^{\prime\prime}=d\}, as every O/D pair has at least two paths, i.e., |𝒮k|2|\mathcal{S}_{k}|\geq 2 for each k𝒦,k\in\mathcal{K}, see Condition 1. Let υ>1\upsilon>1 be an arbitrary factor. By (4.3), (3.11), and a repeated application of the cost normalization Ψυ,\Psi_{\upsilon}, we obtain for each nn\in\mathbb{N} that

|ρ(Γ)ρ(Γ)|=|ρ(Ψυ(Γ))ρ(Ψυ(Γ))|==|ρ(ΨυΨυn(Γ))ρ(ΨυΨυn(Γ))|ϰΨυΨυ(Γ)ΨυΨυ(Γ)γ=ϰΓΓγυnγ,\begin{split}|\rho(\Gamma)-\rho(\Gamma^{\prime})|&=|\rho(\Psi_{\upsilon}(\Gamma))-\rho(\Psi_{\upsilon}(\Gamma^{\prime}))|=\cdots=|\rho(\underbrace{\Psi_{\upsilon}\circ\cdots\circ\Psi_{\upsilon}}_{n}(\Gamma))-\rho(\underbrace{\Psi_{\upsilon}\circ\cdots\circ\Psi_{\upsilon}}_{n}(\Gamma^{\prime}))|\\ &\leq\varkappa\cdot|\!|\Psi_{\upsilon}\circ\cdots\circ\Psi_{\upsilon}(\Gamma)-\Psi_{\upsilon}\circ\cdots\circ\Psi_{\upsilon}(\Gamma^{\prime})|\!|^{\gamma}=\varkappa\cdot\frac{|\!|\Gamma-\Gamma^{\prime}|\!|^{\gamma}}{{\upsilon^{n\cdot\gamma}}},\end{split} (4.4)

which implies |ρ(Γ)ρ(Γ)|=0|\rho(\Gamma)-\rho(\Gamma^{\prime})|=0 by letting nn\to\infty on both sides, and so contradicts with the fact that ρ(Γ)ρ(Γ).\rho(\Gamma)\neq\rho(\Gamma^{\prime}).

Hence, a uniform Hölder constant ϰ>0\varkappa>0 and a uniform Hölder exponent γ>0\gamma>0 cannot exist simultaneously. The Hölder continuity results of Englert et al. [2010] and Takalloo and Kwon [2020] seemingly indicate that there is a uniform Hölder exponent but only with a pointwise Hölder constant, see (4.2). We thus also focus on a uniform Hölder exponent γ\gamma with a pointwise Hölder constant ϰΓ\varkappa_{\Gamma} in our analysis.

Given a game Γ𝒢(G,𝒦,𝒮)\Gamma\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) and a Hölder exponent γ>0,\gamma>0, we call an open subset UΓ𝒢(G,𝒦,𝒮)U_{\Gamma}\subseteq\mathcal{G}(G,\mathcal{K},\mathcal{S}) a Hölder (continuity) neighborhood of order γ\gamma (in short, a γ\gamma-neighborhood) of Γ\Gamma if ΓUΓ,\Gamma\in U_{\Gamma}, and if there is a (pointwise) Hölder constant ϰΓ>0\varkappa_{\Gamma}>0 s.t. |ρ(Γ)ρ(Γ)|ϰΓΓΓγ|\rho(\Gamma^{\prime})-\rho(\Gamma)|\leq\varkappa_{\Gamma}\cdot|\!|\Gamma^{\prime}-\Gamma|\!|^{\gamma} for each ΓUΓ.\Gamma^{\prime}\in U_{\Gamma}. Here, we employ the convention that the empty set \emptyset is a γ\gamma-neighborhood of every game Γ𝒢(G,𝒦,𝒮)\Gamma\in\mathcal{G}(G,\mathcal{K},\mathcal{S}). Then ρ()\rho(\cdot) is Hölder continuous at a game Γ𝒢(G,𝒦,𝒮)\Gamma\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) if and only if Γ\Gamma has a non-empty γ\gamma-neighborhood UΓ𝒢(G,𝒦,𝒮)U_{\Gamma}\subseteq\mathcal{G}(G,\mathcal{K},\mathcal{S}) for some Hölder exponent γ>0.\gamma>0.

We now show that every γ\gamma-neighborhood UΓU_{\Gamma} is a proper subset of 𝒢(G,𝒦,𝒮),\mathcal{G}(G,\mathcal{K},\mathcal{S}), i.e., UΓ𝒢(G,𝒦,𝒮).U_{\Gamma}\neq\mathcal{G}(G,\mathcal{K},\mathcal{S}). This means that the Hölder continuity of the PoA map ρ()\rho(\cdot) can hold only locally at a game Γ𝒢(G,𝒦,𝒮)\Gamma\in\mathcal{G}(G,\mathcal{K},\mathcal{S}), even when the Hölder constant ϰΓ\varkappa_{\Gamma} is pointwise, i.e., may depend on the game Γ.\Gamma.

We again show this by contradiction, and thus assume that there is a game Γ=(τ,d)𝒢(G,𝒦,𝒮)\Gamma=(\tau,d)\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) whose γ\gamma-neighborhood is the entire space 𝒢(G,𝒦,𝒮)\mathcal{G}(G,\mathcal{K},\mathcal{S}) for some Hölder exponent γ>0.\gamma>0. Then there is a Hölder constant ϰΓ>0\varkappa_{\Gamma}>0 such that

|ρ(Γ)ρ(Γ)|ϰΓΓΓγΓ𝒢(G,𝒦,𝒮).|\rho(\Gamma)-\rho(\Gamma^{\prime})|\leq\varkappa_{\Gamma}\cdot|\!|\Gamma-\Gamma^{\prime}|\!|^{\gamma}\quad\forall\Gamma^{\prime}\in\mathcal{G}(G,\mathcal{K},\mathcal{S}). (4.5)

To obtain a contradiction, we consider now an arbitrary game Γ=(σ,d)\Gamma^{\prime}=(\sigma,d) from the same demand slice 𝒢(G,𝒦,𝒮)|d\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|d} of Γ.\Gamma. Let υ>0\upsilon>0 be an arbitrary factor. Then the cost normalization Ψυ()\Psi_{\upsilon}(\cdot) and (4.5) yield

|ρ(Γ)ρ(Γ)|=|ρ(Γ)ρ(Ψυ(Γ))|ϰΓΓΨυ(Γ)γ.|\rho(\Gamma)-\rho(\Gamma^{\prime})|=|\rho(\Gamma)-\rho(\Psi_{\upsilon}(\Gamma^{\prime}))|\leq\varkappa_{\Gamma}\cdot|\!|\Gamma-\Psi_{\upsilon}(\Gamma^{\prime})|\!|^{\gamma}. (4.6)

Note that

ΓΨυ(Γ)=τ|T(d)σ|T(d)υ=maxaA,x[0,T(d)]|τa(x)σa(x)υ|maxaA,x[0,T(d)]τa(x) as υ.|\!|\Gamma-\Psi_{\upsilon}(\Gamma^{\prime})|\!|=|\!|\tau_{|T(d)}-\frac{\sigma_{|T(d)}}{\upsilon}|\!|_{\infty}=\max_{a\in A,x\in[0,T(d)]}\left|\tau_{a}(x)-\frac{\sigma_{a}(x)}{\upsilon}\right|\to\max_{a\in A,x\in[0,T(d)]}\tau_{a}(x)\quad\text{ as }\upsilon\to\infty.

Inequality (4.6) then implies that

|ρ(Γ)ρ(Γ)|ϰΓmaxaA,x[0,T(d)]τa(x)γ=:ϰΓ||τ|T(d)||γ.|\rho(\Gamma)-\rho(\Gamma^{\prime})|\leq\varkappa_{\Gamma}\cdot\max_{a\in A,x\in[0,T(d)]}\tau_{a}(x)^{\gamma}=:\varkappa_{\Gamma}\cdot|\!|\tau_{|T(d)}|\!|_{\infty}^{\gamma}.

Since Γ\Gamma^{\prime} is an arbitrary game of the demand slice 𝒢(G,𝒦,𝒮)|d={Γ=(σ,d)𝒢(G,𝒦,𝒮):d=d},\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|d}=\{\Gamma^{\prime}=(\sigma,d^{\prime})\in\mathcal{G}(G,\mathcal{K},\mathcal{S}):d^{\prime}=d\}, the PoA map ρ()\rho(\cdot) has a uniform finite upper bound ρ(Γ)+ϰΓτ|T(d)γ\rho(\Gamma)+\varkappa_{\Gamma}\cdot|\!|\tau_{|T(d)}|\!|_{\infty}^{\gamma} on the demand slice 𝒢(G,𝒦,𝒮)|d.\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|d}.

However, the demand slice 𝒢(G,𝒦,𝒮)|d\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|d} contains games with polynomial cost functions of degree at most β\beta for arbitrary β>0\beta>0. The PoA map ρ()\rho(\cdot) is then unbounded on the demand slice 𝒢(G,𝒦,𝒮)|d,\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|d}, since these games have a tight upper bound Θ(β/lnβ)\Theta(\beta/\ln\beta) tending to \infty as β,\beta\to\infty, see Roughgarden [2015]. Here, we notice that |𝒮k|2|\mathcal{S}_{k}|\geq 2 for each O/D pair k𝒦,k\in\mathcal{K}, see Condition 1, and thus there is a game Γβ=(σ,d)𝒢(G,𝒦,𝒮)|d\Gamma^{\prime}_{\beta}=(\sigma,d)\in\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|d} for each β>0,\beta>0, who performs similarly to Pigou’s game (Pigou [1920]) with cost functions xβx^{\beta} and 1,1, and whose PoA reaches the upper bound Θ(β/lnβ).\Theta(\beta/\ln\beta).

We summarize this in Theorem 8.

Theorem 8.
  • a)

    There are no constants γ>0\gamma>0 and ϰ>0\varkappa>0 s.t. the PoA map ρ()\rho(\cdot) is uniformly Hölder continuous with Hölder exponent γ\gamma and Hölder constant ϰ.\varkappa.

  • b)

    For every open subset U𝒢(G,𝒦,𝒮)U\subseteq\mathcal{G}(G,\mathcal{K},\mathcal{S}) and every arbitrary Hölder exponent γ>0,\gamma>0, if UU is a γ\gamma-neighborhood of a game Γ𝒢(G,𝒦,𝒮),\Gamma\in\mathcal{G}(G,\mathcal{K},\mathcal{S}), then U𝒢(G,𝒦,𝒮).U\neq\mathcal{G}(G,\mathcal{K},\mathcal{S}).

Remark 4.1.

We actually have proved that ρ()\rho(\cdot) is neither uniformly Hölder continuous on a demand slice, nor has a γ\gamma-neighborhood including a demand slice 𝒢(G,𝒦,𝒮)|d\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|d} as a subspace. However, the PoA map ρ()\rho(\cdot) may be Lipschitz continuous on a cost slice 𝒢(G,𝒦,𝒮)|τ.\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau}. To see this, we consider a vector τ=(τa)aA\tau=(\tau_{a})_{a\in A} of cost functions with τa(x)ca\tau_{a}(x)\equiv c_{a} for some constant ca>0c_{a}>0 and all (a,x)A×[0,).(a,x)\in A\times[0,\infty). Then ρ(Γ)1\rho(\Gamma)\equiv 1 for all Γ=(τ,d)𝒢(G,𝒦,𝒮)|τ,\Gamma=(\tau,d)\in\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau}, and so ρ()\rho(\cdot) is Lipschitz continuous on 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} when we restrict the map ρ()\rho(\cdot) onto 𝒢(G,𝒦,𝒮)|τ.\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau}. In fact, there is even a 11-neighborhood including the cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} as a subspace. To show this, pick an arbitrary Γ=(τ,d)𝒢(G,𝒦,𝒮)|τ\Gamma=(\tau,d)\in\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} and an arbitrary constant κ>0.\kappa>0. Then

𝒢(G,𝒦,𝒮)|τ{Γ}VΓ(κ):={Γ𝒢(G,𝒦,𝒮):|ρ(Γ)ρ(Γ)|κΓΓ<0},\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau}\setminus\{\Gamma\}\subseteq V_{\Gamma}(\kappa):=\{\Gamma^{\prime}\in\mathcal{G}(G,\mathcal{K},\mathcal{S}):\ |\rho(\Gamma^{\prime})-\rho(\Gamma)|-\kappa\cdot|\!|\Gamma-\Gamma^{\prime}|\!|<0\},

and VΓ(κ)V_{\Gamma}(\kappa) is open since ρ()\rho(\cdot) is continuous. Theorem 12a) in Section 4.3 shows that there are a Hölder constant ϰΓ>0\varkappa_{\Gamma}>0 and a nonempty 11-neighborhood UΓU_{\Gamma} with |ρ(Γ)ρ(Γ)|ϰΓΓΓ|\rho(\Gamma^{\prime})-\rho(\Gamma)|\leq\varkappa_{\Gamma}\cdot|\!|\Gamma-\Gamma^{\prime}|\!| for each ΓUΓ.\Gamma^{\prime}\in U_{\Gamma}. Thus

|ρ(Γ)ρ(Γ)|max{κ,ϰΓ}ΓΓΓUΓVΓ(κ).|\rho(\Gamma^{\prime})-\rho(\Gamma)|\leq\max\{\kappa,\varkappa_{\Gamma}\}\cdot|\!|\Gamma^{\prime}-\Gamma|\!|\quad\forall\Gamma^{\prime}\in U_{\Gamma}\cup V_{\Gamma}(\kappa).

Clearly, UΓVΓ(κ)U_{\Gamma}\cup V_{\Gamma}(\kappa) is a 11-neighborhood of Γ\Gamma that includes the whole cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} as a subspace. Hence, the Hölder continuity of the PoA may differ largely on the two types of slices. We thus need separate discussions for them when we analyze the pointwise Hölder continuity of the PoA.

4.2 Pointwise Hölder continuity of the PoA

Because of Theorem 8, we now consider Hölder continuity of the PoA map ρ()\rho(\cdot) pointwise and locally at each game Γ𝒢(G,𝒦,𝒮)\Gamma\in\mathcal{G}(G,\mathcal{K},\mathcal{S}).

Theorem 9 shows that the PoA is Hölder continuous with Hölder exponent γ=12\gamma=\frac{1}{2} at every game Γ=(τ,d)𝒢(G,𝒦,𝒮)\Gamma=(\tau,d)\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) whose cost functions τa()\tau_{a}(\cdot) are Lipschitz continuous on the compact interval [0,T(d)].[0,T(d)]. Note that games satisfying these assumptions are dense in 𝒢(G,𝒦,𝒮),\mathcal{G}(G,\mathcal{K},\mathcal{S}), i.e., every game Γ𝒢(G,𝒦,𝒮)\Gamma^{\prime}\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) is the limit of a convergent sequence of such games. Thus every non-empty open subset UU of 𝒢(G,𝒦,𝒮)\mathcal{G}(G,\mathcal{K},\mathcal{S}) contains a non-empty 12\frac{1}{2}-neighborhood, although Theorem 8b) implies that this neighborhood might not be large.

Theorem 9.

Consider an arbitrary game Γ=(τ,d)𝒢(G,𝒦,𝒮)\Gamma=(\tau,d)\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) whose cost functions τa()\tau_{a}(\cdot) are Lipschitz continuous on the compact interval [0,T(d)].[0,T(d)]. Then the PoA map ρ()\rho(\cdot) is Hölder continuous at game Γ\Gamma with Hölder exponent γ=12\gamma=\frac{1}{2} within a γ\gamma-neighborhood BϵΓ(Γ)B_{\epsilon_{\Gamma}}(\Gamma) for a small constant ϵΓ>0\epsilon_{\Gamma}>0 depending only on the game Γ.\Gamma.

Theorem 9 presents the first pointwise Hölder continuity result of the PoA map ρ()\rho(\cdot) on the whole game space 𝒢(G,𝒦,𝒮)\mathcal{G}(G,\mathcal{K},\mathcal{S}). It applies to the most general case that the cost functions and the demands change simultaneously.

To prove Theorem 9, we will now analyze the Hölder continuity of the PoA on cost and demand slices separately, since it may differ on cost and demand slices, see Remark 4.1. Theorem 9 then follows by combining the results appropriately.

4.2.1 Hölder continuity of the PoA on demand slices

Lemma 10 presents the first results about Hölder continuity of the PoA on a demand slice 𝒢(G,𝒦,𝒮)|w\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|w} for an arbitrary demand vector w=(wk)k𝒦w=(w_{k})_{k\in\mathcal{K}} with T(w)=k𝒦wk>0.T(w)=\sum_{k\in\mathcal{K}}w_{k}>0. Lemma 10a) shows that the SO map 𝒞()\mathcal{C}^{*}(\cdot) is Lipschitz continuous with Lipschitz constant |A|T(w)|A|\cdot T(w) on the demand slice 𝒢(G,𝒦,𝒮)|w.\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|w}. Lemma 10b) shows a similar continuity result for the potential function values of WE flows of games in 𝒢(G,𝒦,𝒮)|w.\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|w}. Lemma 10c) shows that the WE flows of two arbitrary games Γ1\Gamma_{1} and Γ2\Gamma_{2} in 𝒢(G,𝒦,𝒮)|w\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|w} are mutually |A|T(w)Γ1Γ2|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|-approximate WE flows of each other. Finally, with Lemma 1c) and Lemma 10c), Lemma 10d) shows that, when restricted to the demand slice 𝒢(G,𝒦,𝒮)|w,\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|w}, both the WE cost map 𝒞~()\tilde{\mathcal{C}}(\cdot) and the PoA map ρ()\rho(\cdot) are pointwise Hölder continuous with Hölder exponent 12\frac{1}{2} at each game Γ1𝒢(G,𝒦,𝒮)|w\Gamma_{1}\in\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|w} whose cost functions are Lipschitz continuous on the compact interval [0,T(w)].[0,T(w)].

Lemma 10.

Consider an arbitrary demand vector w=(wk)k𝒦w=(w_{k})_{k\in\mathcal{K}} with T(w)>0,T(w)>0, and two arbitrary games Γ1=(π(1),w)\Gamma_{1}=(\pi^{(1)},w) and Γ2=(π(2),w)\Gamma_{2}=(\pi^{(2)},w) of the demand slice 𝒢(G,𝒦,𝒮)|w.\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|w}. Let f~\tilde{f} and g~\tilde{g} be WE flows of Γ1\Gamma_{1} and Γ2,\Gamma_{2}, respectively. Then, the following statements hold.

  • a)

    |𝒞(Γ1)𝒞(Γ2)||A|T(w)Γ1Γ2.|\mathcal{C}^{*}(\Gamma_{1})-\mathcal{C}^{*}(\Gamma_{2})|\leq|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|.

  • b)

    |Φ(Γ1,f)Φ(Γ2,f)||A|T(w)Γ1Γ2|\Phi(\Gamma_{1},f)-\Phi(\Gamma_{2},f)|\leq|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!| for every flow f.f. Moreover, |Φ(Γ1,f~)Φ(Γ2,g~)||A|T(w)Γ1Γ2,|\Phi(\Gamma_{1},\tilde{f})-\Phi(\Gamma_{2},\tilde{g})|\leq|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|, 0Φ(Γ1,g~)Φ(Γ1,f~)2|A|T(w)Γ1Γ2,0\leq\Phi(\Gamma_{1},\tilde{g})-\Phi(\Gamma_{1},\tilde{f})\leq 2\cdot|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|, and 0Φ(Γ2,f~)Φ(Γ2,g~)2|A|T(w)Γ1Γ2.0\leq\Phi(\Gamma_{2},\tilde{f})-\Phi(\Gamma_{2},\tilde{g})\leq 2\cdot|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|.

  • c)

    f~\tilde{f} is an |A|T(w)Γ1Γ2|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|-approximate WE flow of the game Γ2,\Gamma_{2}, and g~\tilde{g} is an |A|T(w)Γ1Γ2|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|-approximate WE flow of the game Γ1.\Gamma_{1}.

  • d)

    If πa(1)()\pi_{a}^{(1)}(\cdot) is Lipschitz continuous on [0,T(w)][0,T(w)] with Lipschitz constant M>0M>0 for each aA,a\in A, then |𝒞~(Γ1)𝒞~(Γ2)|(M|A|T(w)+2)|A|T(w)max{Γ1Γ2,Γ1Γ2},|\tilde{\mathcal{C}}(\Gamma_{1})-\tilde{\mathcal{C}}(\Gamma_{2})|\leq(\sqrt{M\cdot|A|\cdot T(w)}+2)\cdot|A|\cdot T(w)\cdot\max\{\sqrt{|\!|\Gamma_{1}-\Gamma_{2}|\!|},\ |\!|\Gamma_{1}-\Gamma_{2}|\!|\}, and so |ρ(Γ1)ρ(Γ2)|νΓ1max{Γ1Γ2,Γ1Γ2}|\rho(\Gamma_{1})-\rho(\Gamma_{2})|\leq\nu_{\Gamma_{1}}\cdot\max\{\sqrt{|\!|\Gamma_{1}-\Gamma_{2}|\!|},\ |\!|\Gamma_{1}-\Gamma_{2}|\!|\} with

    νΓ1:=2ρ(Γ1)+M|A|T(w)+2𝒞(Γ1)|A|T(w)\nu_{\Gamma_{1}}:=2\cdot\frac{\rho(\Gamma_{1})+\sqrt{M\cdot|A|\cdot T(w)}+2}{\mathcal{C}^{*}(\Gamma_{1})}\cdot|A|\cdot T(w)

    when Γ1Γ2𝒞(Γ1)2|A|T(w).|\!|\Gamma_{1}-\Gamma_{2}|\!|\leq\frac{\mathcal{C}^{*}(\Gamma_{1})}{2\cdot|A|\cdot T(w)}.

Proof.

Proof of Lemma 10 Let ff^{*} and gg^{*} be an SO flow of Γ1\Gamma_{1} and an SO flow of Γ2,\Gamma_{2}, respectively. Note that Γ1\Gamma_{1} and Γ2\Gamma_{2} have the same set of flows, since both belong to the demand slice 𝒢(G,𝒦,𝒮)|w.\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|w}. So g~\tilde{g} and gg^{*} are also flows of Γ1,\Gamma_{1}, and f~\tilde{f} and ff^{*} are also flows of Γ2.\Gamma_{2}.

Proof.

Proof of Lemma 10a): Note that

𝒞(Γ1)=C(Γ1,f)C(Γ1,g)=aAπa(1)(ga)gaaAπa(2)(ga)ga+aAgaΓ1Γ2=𝒞(Γ2)+aAgaΓ1Γ2𝒞(Γ2)+|A|T(w)Γ1Γ2.\begin{split}\mathcal{C}^{*}(\Gamma_{1})&=C(\Gamma_{1},f^{*})\leq C(\Gamma_{1},g^{*})=\sum_{a\in A}\pi^{(1)}_{a}(g_{a}^{*})\cdot g_{a}^{*}\leq\sum_{a\in A}\pi^{(2)}_{a}(g_{a}^{*})\cdot g_{a}^{*}+\sum_{a\in A}g_{a}^{*}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|\\ &=\mathcal{C}^{*}(\Gamma_{2})+\sum_{a\in A}g_{a}^{*}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|\leq\mathcal{C}^{*}(\Gamma_{2})+|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|.\end{split}

Here, we have used that gaT(w)g_{a}^{*}\leq T(w) for each aA,a\in A, and that

|πa(1)(x)πa(2)(x)|Γ1Γ2=maxbA,y[0,T(w)]|πb(1)(y)πb(2)(y)|aAx[0,T(w)].|\pi_{a}^{(1)}(x)-\pi^{(2)}_{a}(x)|\leq|\!|\Gamma_{1}-\Gamma_{2}|\!|=\max_{b\in A,y\in[0,T(w)]}|\pi_{b}^{(1)}(y)-\pi^{(2)}_{b}(y)|\quad\forall a\in A\ \forall x\in[0,T(w)].

Similarly, we have 𝒞(Γ2)𝒞(Γ1)+|A|T(w)Γ1Γ2.\mathcal{C}^{*}(\Gamma_{2})\leq\mathcal{C}^{*}(\Gamma_{1})+|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|. Then Lemma 10a) follows.

Proof.

Proof of Lemma 10b): Consider an arbitrary flow ff for the demand vector w.w. Definition (2.6) of the potential function Φ(,)\Phi(\cdot,\cdot) implies that

Φ(Γ1,f)=aA0faπa(1)(x)𝑑xaA0faπa(2)(x)𝑑x+aAfaΓ1Γ2Φ(Γ2,f)+|A|T(w)Γ1Γ2,\Phi(\Gamma_{1},f)=\sum_{a\in A}\int_{0}^{f_{a}}\pi_{a}^{(1)}(x)dx\leq\sum_{a\in A}\int_{0}^{f_{a}}\pi_{a}^{(2)}(x)dx+\sum_{a\in A}f_{a}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|\leq\Phi(\Gamma_{2},f)+|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|,

and, similarly that Φ(Γ2,f)Φ(Γ1,f)+|A|T(w)Γ1Γ2.\Phi(\Gamma_{2},f)\leq\Phi(\Gamma_{1},f)+|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|. Thus we have |Φ(Γ1,f)Φ(Γ2,f)||A|T(w)Γ1Γ2|\Phi(\Gamma_{1},f)-\Phi(\Gamma_{2},f)|\leq|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!| for an arbitrary flow f.f.

Hence

Φ(Γ1,f~)Φ(Γ1,g~)Φ(Γ2,g~)+|A|T(w)Γ1Γ2,\Phi(\Gamma_{1},\tilde{f})\leq\Phi(\Gamma_{1},\tilde{g})\leq\Phi(\Gamma_{2},\tilde{g})+|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|,

and, similarly, Φ(Γ2,g~)Φ(Γ1,f~)+|A|T(w)Γ1Γ2.\Phi(\Gamma_{2},\tilde{g})\leq\Phi(\Gamma_{1},\tilde{f})+|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|. Here, we have used the fact that WE flows of a game minimize the potential function of that game. So |Φ(Γ1,f~)Φ(Γ2,g~)||A|T(w)Γ1Γ2.|\Phi(\Gamma_{1},\tilde{f})-\Phi(\Gamma_{2},\tilde{g})|\leq|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|.

Hence

0Φ(Γ1,g~)Φ(Γ1,f~)=Φ(Γ1,g~)Φ(Γ2,g~)+Φ(Γ2,g~)Φ(Γ1,f~)2|A|T(w)Γ1Γ2,0Φ(Γ2,f~)Φ(Γ2,g~)Φ(Γ2,f~)Φ(Γ1,f~)+Φ(Γ1,f~)Φ(Γ2,g~)2|A|T(w)Γ1Γ2.\begin{split}&0\leq\Phi(\Gamma_{1},\tilde{g})-\Phi(\Gamma_{1},\tilde{f})=\Phi(\Gamma_{1},\tilde{g})-\Phi(\Gamma_{2},\tilde{g})+\Phi(\Gamma_{2},\tilde{g})-\Phi(\Gamma_{1},\tilde{f})\leq 2\cdot|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|,\\ &0\leq\Phi(\Gamma_{2},\tilde{f})-\Phi(\Gamma_{2},\tilde{g})\leq\Phi(\Gamma_{2},\tilde{f})-\Phi(\Gamma_{1},\tilde{f})+\Phi(\Gamma_{1},\tilde{f})-\Phi(\Gamma_{2},\tilde{g})\leq 2\cdot|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|.\end{split}

This completes the proof of Lemma 10b).

Proof.

Proof of Lemma 10c): Consider again an arbitrary flow ff for the demand vector w.w. Since f~\tilde{f} is a WE flow of Γ1=(π(1),w),\Gamma_{1}=(\pi^{(1)},w), we have

aAπa(2)(f~a)(f~afa)=aAπa(1)(f~a)(f~afa)+aA(πa(2)(f~a)πa(1)(f~a))(f~afa)aA(πa(2)(f~a)πa(1)(f~a))(f~afa)|A|T(w)Γ1Γ2.\begin{split}\sum_{a\in A}\pi_{a}^{(2)}(\tilde{f}_{a})\cdot(\tilde{f}_{a}-f_{a})&=\sum_{a\in A}\pi_{a}^{(1)}(\tilde{f}_{a})\cdot(\tilde{f}_{a}-f_{a})+\sum_{a\in A}(\pi_{a}^{(2)}(\tilde{f}_{a})-\pi_{a}^{(1)}(\tilde{f}_{a}))\cdot(\tilde{f}_{a}-f_{a})\\ &\leq\sum_{a\in A}(\pi_{a}^{(2)}(\tilde{f}_{a})-\pi_{a}^{(1)}(\tilde{f}_{a}))\cdot(\tilde{f}_{a}-f_{a})\leq|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|.\end{split} (4.7)

Similarly,

aAπa(1)(g~a)(g~afa)aA(πa(1)(g~a)πa(2)(g~a))(g~afa)|A|T(w)Γ1Γ2.\sum_{a\in A}\pi_{a}^{(1)}(\tilde{g}_{a})\cdot(\tilde{g}_{a}-f_{a})\leq\sum_{a\in A}(\pi_{a}^{(1)}(\tilde{g}_{a})-\pi_{a}^{(2)}(\tilde{g}_{a}))\cdot(\tilde{g}_{a}-f_{a})\leq|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|. (4.8)

Then Lemma 10c) follows from (4.7)–(4.8) and the definition of ϵ\epsilon-approximate WE flows.

Proof.

Proof of Lemma 10d): Lemma 1c) and Lemma 10c) together imply that

|𝒞~(Γ1)C(Γ1,g~)|=|C(Γ1,f~)C(Γ1,g~)||A|T(w)M|A|T(w)Γ1Γ2+|A|T(w)Γ1Γ2.\begin{split}|\tilde{\mathcal{C}}(\Gamma_{1})-C(\Gamma_{1},\tilde{g})|&=|C(\Gamma_{1},\tilde{f})-C(\Gamma_{1},\tilde{g})|\\ &\leq|A|\cdot T(w)\cdot\sqrt{M\cdot|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|}+|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|.\end{split} (4.9)

Trivially,

|𝒞~(Γ2)C(Γ1,g~)|=|C(Γ2,g~)C(Γ1,g~)|aA|πa(2)(g~a)πa(1)(g~a)|g~a|A|T(w)Γ1Γ2.\begin{split}|\tilde{\mathcal{C}}(\Gamma_{2})-C(\Gamma_{1},\tilde{g})|&=|C(\Gamma_{2},\tilde{g})-C(\Gamma_{1},\tilde{g})|\leq\sum_{a\in A}|\pi_{a}^{(2)}(\tilde{g}_{a})-\pi_{a}^{(1)}(\tilde{g}_{a})|\cdot\tilde{g}_{a}\\ &\leq|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|.\end{split} (4.10)

Hence

|𝒞~(Γ1)𝒞~(Γ2)||A|T(w)M|A|T(w)Γ1Γ2+2|A|T(w)Γ1Γ2{(M|A|T(w)+2)|A|T(w)Γ1Γ2if Γ1Γ21,(M|A|T(w)+2)|A|T(w)Γ1Γ2if Γ1Γ2>1.\begin{split}|\tilde{\mathcal{C}}(\Gamma_{1})-\tilde{\mathcal{C}}(\Gamma_{2})|&\leq|A|\cdot T(w)\cdot\sqrt{M\cdot|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|}+2\cdot|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|\\ &\leq\begin{cases}(\sqrt{M\cdot|A|\cdot T(w)}+2)\cdot|A|\cdot T(w)\cdot\sqrt{|\!|\Gamma_{1}-\Gamma_{2}|\!|}&\text{if }|\!|\Gamma_{1}-\Gamma_{2}|\!|\leq 1,\\ (\sqrt{M\cdot|A|\cdot T(w)}+2)\cdot|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|&\text{if }|\!|\Gamma_{1}-\Gamma_{2}|\!|>1.\end{cases}\end{split} (4.11)

This together with Lemma 10a) implies that

ρ(Γ1)ρ(Γ2)𝒞~(Γ1)𝒞(Γ1)𝒞~(Γ1)(M|A|T(w)+2)|A|T(w)max{Γ1Γ2,Γ1Γ2}𝒞(Γ1)+|A|T(w)Γ1Γ2=ρ(Γ1)|A|T(w)Γ1Γ2+(M|A|T(w)+2)|A|T(w)max{Γ1Γ2,Γ1Γ2}𝒞(Γ1)+|A|T(w)Γ1Γ2ρ(Γ1)+M|A|T(w)+2𝒞(Γ1)+|A|T(w)Γ1Γ2|A|T(w)max{Γ1Γ2,Γ1Γ2},\begin{split}&\rho(\Gamma_{1})-\rho(\Gamma_{2})\leq\frac{\tilde{\mathcal{C}}(\Gamma_{1})}{\mathcal{C}^{*}(\Gamma_{1})}-\frac{\tilde{\mathcal{C}}(\Gamma_{1})-(\sqrt{M\cdot|A|\cdot T(w)}+2)\cdot|A|\cdot T(w)\cdot\max\{\sqrt{\Gamma_{1}-\Gamma_{2}},\ |\!|\Gamma_{1}-\Gamma_{2}|\!|\}}{\mathcal{C}^{*}(\Gamma_{1})+|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|}\\ &=\frac{\rho(\Gamma_{1})\cdot|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|+(\sqrt{M\cdot|A|\cdot T(w)}+2)\cdot|A|\cdot T(w)\cdot\max\{\sqrt{\Gamma_{1}-\Gamma_{2}},\ |\!|\Gamma_{1}-\Gamma_{2}|\!|\}}{\mathcal{C}^{*}(\Gamma_{1})+|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|}\\ &\leq\frac{\rho(\Gamma_{1})+\sqrt{M\cdot|A|\cdot T(w)}+2}{\mathcal{C}^{*}(\Gamma_{1})+|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|}\cdot|A|\cdot T(w)\cdot\max\{\sqrt{\Gamma_{1}-\Gamma_{2}},\ |\!|\Gamma_{1}-\Gamma_{2}|\!|\},\end{split}

and that

ρ(Γ2)ρ(Γ1)\displaystyle\rho(\Gamma_{2})-\rho(\Gamma_{1}) 𝒞~(Γ1)+(M|A|T(w)+2)|A|T(w)max{Γ1Γ2,Γ1Γ2}𝒞(Γ1)|A|T(w)Γ1Γ2𝒞~(Γ1)𝒞(Γ1)\displaystyle\leq\frac{\tilde{\mathcal{C}}(\Gamma_{1})+(\sqrt{M\cdot|A|\cdot T(w)}+2)\cdot|A|\cdot T(w)\cdot\max\{\sqrt{|\!|\Gamma_{1}-\Gamma_{2}|\!|},\ |\!|\Gamma_{1}-\Gamma_{2}|\!|\}}{\mathcal{C}^{*}(\Gamma_{1})-|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|}-\frac{\tilde{\mathcal{C}}(\Gamma_{1})}{\mathcal{C}^{*}(\Gamma_{1})}
ρ(Γ1)+M|A|T(w)+2𝒞(Γ1)|A|T(w)||Γ1Γ2|||A|T(w)max{Γ1Γ2,||Γ1Γ2||}\displaystyle\leq\frac{\rho(\Gamma_{1})+\sqrt{M\cdot|A|\cdot T(w)}+2}{\mathcal{C}^{*}(\Gamma_{1})-|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|}\cdot|A|\cdot T(w)\cdot\max\{\sqrt{\Gamma_{1}-\Gamma_{2}},\ |\!|\Gamma_{1}-\Gamma_{2}|\!|\} (4.12)

when ||Γ1Γ2||𝒞(Γ1)|A|T(w).|\!|\Gamma_{1}-\Gamma_{2}|\!|\leq\frac{\mathcal{C}^{*}(\Gamma_{1})}{|A|\cdot T(w)}. Therefore, |ρ(Γ1)ρ(Γ2)|νΓ1max{||Γ1Γ2||,||Γ1Γ2||}|\rho(\Gamma_{1})-\rho(\Gamma_{2})|\leq\nu_{\Gamma_{1}}\cdot\max\{\sqrt{|\!|\Gamma_{1}-\Gamma_{2}|\!|},\ |\!|\Gamma_{1}-\Gamma_{2}|\!|\} with

νΓ1:=2ρ(Γ1)+M|A|T(w)+2𝒞(Γ1)|A|T(w)\nu_{\Gamma_{1}}:=2\cdot\frac{\rho(\Gamma_{1})+\sqrt{M\cdot|A|\cdot T(w)}+2}{\mathcal{C}^{*}(\Gamma_{1})}\cdot|A|\cdot T(w)

when ||Γ1Γ2||𝒞(Γ1)2|A|T(w).|\!|\Gamma_{1}-\Gamma_{2}|\!|\leq\frac{\mathcal{C}^{*}(\Gamma_{1})}{2\cdot|A|\cdot T(w)}. This completes the proof of Lemma 10d).

\square

Remark 4.2.

In the proof of Lemma 10d), we have used Lemma 1c) to bound |C(Γ1,f~)C(Γ1,g~)|,|C(\Gamma_{1},\tilde{f})-C(\Gamma_{1},\tilde{g})|, see (4.9), since Lemma 10c) has shown that g~\tilde{g} is an |A|T(w)||Γ1Γ2|||A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|-approximate WE flow of Γ1.\Gamma_{1}. Note that |A|T(w)||Γ1Γ2||Θ(||Γ1Γ2||)|A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|\in\Theta(|\!|\Gamma_{1}-\Gamma_{2}|\!|) is already a tight upper bound on the approximation threshold ϵ(π(1),w,g~)\epsilon(\pi^{(1)},w,\tilde{g}) of the flow g~\tilde{g} for Lipschitz continuous cost functions on [0,T(w)][0,T(w)]. We illustrate this with the two games Γ1\Gamma_{1} and Γ2\Gamma_{2} shown in Figure 3(a)–(b). Clearly, ||Γ1Γ2||=ϵ|\!|\Gamma_{1}-\Gamma_{2}|\!|=\epsilon since Γ1\Gamma_{1} and Γ2\Gamma_{2} have the same demand vector w=(1).w=(1). When viewed as a flow of Γ1,\Gamma_{1}, the WE flow g~=(0.5,0.5)\tilde{g}=(0.5,0.5) of Γ2\Gamma_{2} has the approximation threshold ϵ(π(1),w,g~)=g~[π(1)(g~)πu(1)(g~u)]=0.5ϵΘ(||Γ1Γ2||),\epsilon(\pi^{(1)},w,\tilde{g})=\tilde{g}_{\ell}\cdot[\pi_{\ell}^{(1)}(\tilde{g}_{\ell})-\pi_{u}^{(1)}(\tilde{g}_{u})]=0.5\cdot\epsilon\in\Theta(|\!|\Gamma_{1}-\Gamma_{2}|\!|), which shows that the upper bound |A|T(w)||Γ1Γ2|||A|\cdot T(w)\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!| is tight (w.r.t. the magnitude and the exponent of ||Γ1Γ2|||\!|\Gamma_{1}-\Gamma_{2}|\!|). This means that the exponent 12\frac{1}{2} in the right-hand of the inequality (4.9) cannot be improved when the cost functions are Lipschitz continuous on the interval [0,T(w)],[0,T(w)], and when we use Lemma 1c) to bound |C(Γ1,f~)C(Γ1,g~)||C(\Gamma_{1},\tilde{f})-C(\Gamma_{1},\tilde{g})|, since Example 2 has also shown the tightness of Lemma 1c). Hence, to improve the Hölder exponent, we need a finer analysis, which we will develop for cost functions with special properties in Section 4.3.

oottπ(1)u(x)=x\pi^{(1)}_{u}(x)=xπ(1)(x)=x+ϵ\pi^{(1)}_{\ell}(x)=x+\epsilon
(a) Γ1\Gamma_{1} with f~=(1+ϵ2,1ϵ2)\tilde{f}=(\frac{1+\epsilon}{2},\frac{1-\epsilon}{2})
oottπ(2)u(x)=x\pi^{(2)}_{u}(x)=xπ(2)(x)=x\pi^{(2)}_{\ell}(x)=x
(b) Γ2\Gamma_{2} with g~=(0.5,0.5)\tilde{g}=(0.5,0.5)
Figure 3: Games with total demand of 11

4.2.2 Hölder continuity of the PoA on cost slices

Lemma 11 shows results similar to Lemma 10 for two arbitrary games Γ1=(π,w)\Gamma_{1}=(\pi,w) and Γ2=(π,w)\Gamma_{2}=(\pi,w^{\prime}) from the same cost slice 𝒢(G,𝒦,𝒮)|π\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\pi} for a cost function vector π\pi defined on [0,)[0,\infty) and satisfying Condition 2. Using Lemma 10d), Lemma 11a) shows that |𝒞~(Γ1)𝒞~(Γ2)|ν~Γ1||Γ1Γ2|||\tilde{\mathcal{C}}(\Gamma_{1})-\tilde{\mathcal{C}}(\Gamma_{2})|\leq\tilde{\nu}_{\Gamma_{1}}\cdot\sqrt{|\!|\Gamma_{1}-\Gamma_{2}|\!|} for a constant ν~Γ1>0\tilde{\nu}_{\Gamma_{1}}>0 depending only on Γ1,\Gamma_{1}, when the cost functions πa()\pi_{a}(\cdot) are Lipschitz continuous on the compact interval [0,T(w)],[0,T(w)], T(w)<T(w)T(w^{\prime})<T(w) and ||Γ1Γ2|||\!|\Gamma_{1}-\Gamma_{2}|\!| is small. Using Lemma 10a), Lemma 11b) shows with similar assumptions that |𝒞(Γ1)𝒞(Γ2)|νΓ1||Γ1Γ2|||\mathcal{C}^{*}(\Gamma_{1})-\mathcal{C}^{*}(\Gamma_{2})|\leq\nu^{*}_{\Gamma_{1}}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!| for a constant νΓ1>0\nu^{*}_{\Gamma_{1}}>0 depending only on Γ1.\Gamma_{1}. Lemma 11c) then presents an upper bound for |ρ(Γ1)ρ(Γ2)||\rho(\Gamma_{1})-\rho(\Gamma_{2})|. Note that the Lipschitz constant MΓ1M_{\Gamma_{1}} is required to be at least 11 in Lemma 11. However, this is not an additional restriction, as every Lipschitz continuous function always has a Lipschitz constant of at least 1.1.

Lemma 11.

Consider an arbitrary game Γ1=(π,w)𝒢(G,𝒦,𝒮)\Gamma_{1}=(\pi,w)\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) such that the cost functions πa()\pi_{a}(\cdot) are Lipschitz continuous on the compact interval [0,T(w)][0,T(w)] with a Lipschitz constant MΓ11.M_{\Gamma_{1}}\geq 1. Then the following statements hold.

  • a)

    For each game Γ2=(π,w)𝒢(G,𝒦,𝒮)\Gamma_{2}=(\pi,w^{\prime})\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) with T(w)T(w),T(w^{\prime})\leq T(w),

    |𝒞~(Γ2)𝒞~(Γ1)|2((MΓ1|A|T(w)+2)|A|T(w)+|A|πmax(T(w))|𝒦|)MΓ1max{||Γ1Γ2||,MΓ1||Γ1Γ2||}=:ν~Γ1max{||Γ1Γ2||,MΓ1||Γ1Γ2||}\begin{split}|\tilde{\mathcal{C}}(\Gamma_{2})-\tilde{\mathcal{C}}(\Gamma_{1})|&\leq 2\cdot\Big{(}(\sqrt{M_{\Gamma_{1}}\cdot|A|\cdot T(w)}+2)\cdot|A|\cdot T(w)\\ &\hskip 14.22636pt+|A|\cdot\pi_{max}(T(w))\cdot|\mathcal{K}|\Big{)}\cdot\sqrt{M_{\Gamma_{1}}}\cdot\max\{\sqrt{|\!|\Gamma_{1}-\Gamma_{2}|\!|},\ \sqrt{M_{\Gamma_{1}}}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|\}\\ &=:\tilde{\nu}_{\Gamma_{1}}\cdot\max\{\sqrt{|\!|\Gamma_{1}-\Gamma_{2}|\!|},\ \sqrt{M_{\Gamma_{1}}}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|\}\end{split}

    when ||Γ1Γ2||<T(w)|𝒦|,|\!|\Gamma_{1}\!-\!\Gamma_{2}|\!|\!<\!\frac{T(w)}{|\mathcal{K}|}, where ν~Γ1:=2((MΓ1|A|T(w)+2)|A|T(w)+|A||𝒦|πmax(T(w)))MΓ1\tilde{\nu}_{\Gamma_{1}}:=2\cdot\big{(}\!\big{(}\sqrt{M_{\Gamma_{1}}\!\cdot\!|A|\!\cdot\!T(w)}\!+\!2\big{)}\cdot|A|\cdot T(w)\!+\!|A|\cdot|\mathcal{K}|\cdot\pi_{max}(T(w))\!\big{)}\cdot\sqrt{M_{\Gamma_{1}}} and πmax(T(w)):=maxaAπa(T(w)).\pi_{max}(T(w))\!:=\!\max_{a\in A}\pi_{a}(T(w)).

  • b)

    For each game Γ2=(π,w)𝒢(G,𝒦,𝒮)\Gamma_{2}=(\pi,w^{\prime})\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) with T(w)T(w),T(w^{\prime})\leq T(w),

    |𝒞(Γ1)𝒞(Γ2)|2(|A||𝒦|πmax(T(w))+|A|T(w)MΓ1)||Γ1Γ2||=:νΓ1||Γ1Γ2|||\mathcal{C}^{*}(\Gamma_{1})-\mathcal{C}^{*}(\Gamma_{2})|\leq 2\cdot\big{(}|A|\cdot|\mathcal{K}|\cdot\pi_{max}(T(w))+|A|\cdot T(w)\cdot M_{\Gamma_{1}}\big{)}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|=:\nu^{*}_{\Gamma_{1}}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|

    when ||Γ1Γ2||<T(w)|𝒦|,|\!|\Gamma_{1}-\Gamma_{2}|\!|<\frac{T(w)}{|\mathcal{K}|}, where νΓ1:=2(|A||𝒦|πmax(T(w))+|A|T(w)MΓ1).\nu^{*}_{\Gamma_{1}}:=2\cdot\big{(}|A|\cdot|\mathcal{K}|\cdot\pi_{max}(T(w))+|A|\cdot T(w)\cdot M_{\Gamma_{1}}\big{)}.

  • c)

    For each game Γ2=(π,w)𝒢(G,𝒦,𝒮)\Gamma_{2}=(\pi,w^{\prime})\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) with T(w)T(w),T(w^{\prime})\leq T(w), it holds that |ρ(Γ1)ρ(Γ2)|<νΓ1max{||Γ1Γ2||,MΓ1||Γ1Γ2||}|\rho(\Gamma_{1})-\rho(\Gamma_{2})|<\nu^{\prime}_{\Gamma_{1}}\cdot\max\{\sqrt{|\!|\Gamma_{1}-\Gamma_{2}|\!|},\sqrt{M_{\Gamma_{1}}}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|\} when νΓ1:=2ρ(Γ1)νΓ1+2ν~Γ2𝒞(Γ1)\nu^{\prime}_{\Gamma_{1}}:=\frac{2\cdot\rho(\Gamma_{1})\cdot\nu^{*}_{\Gamma_{1}}+2\cdot\tilde{\nu}_{\Gamma_{2}}}{\mathcal{C}^{*}(\Gamma_{1})} and ||Γ1Γ2||min{T(w)|𝒦|,𝒞(Γ1)2νΓ1}.|\!|\Gamma_{1}-\Gamma_{2}|\!|\leq\min\{\frac{T(w)}{|\mathcal{K}|},\ \frac{\mathcal{C}^{*}(\Gamma_{1})}{2\cdot\nu^{*}_{\Gamma_{1}}}\}.

Proof.

Proof of Lemma 11 Define an auxiliary demand vector w^=(w^k)k𝒦\hat{w}=(\hat{w}_{k})_{k\in\mathcal{K}} with w^k=min{wk,wk}0\hat{w}_{k}=\min\{w_{k},w^{\prime}_{k}\}\geq 0 for each k𝒦.k\in\mathcal{K}. Denote by Γ^\hat{\Gamma} the pair Γ^=(π,w^).\hat{\Gamma}=(\pi,\hat{w}). Since T(w)>0,T(w)>0, T(w^)T(w)|𝒦|||Γ1Γ2||>0T(\hat{w})\geq T(w)-|\mathcal{K}|\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|>0 and Γ^\hat{\Gamma} is a game, when the condition ||Γ1Γ2||<T(w)|𝒦||\!|\Gamma_{1}-\Gamma_{2}|\!|<\frac{T(w)}{|\mathcal{K}|} is fulfilled. Here, we use that

||Γ1Γ2||=max{||ww||,||π(T(w))π(T(w))||}||ww||=maxk𝒦|wkwk|,|\!|\Gamma_{1}-\Gamma_{2}|\!|=\max\{|\!|w-w^{\prime}|\!|_{\infty},|\!|\pi(T(w))-\pi(T(w^{\prime}))|\!|_{\infty}\}\geq|\!|w-w^{\prime}|\!|_{\infty}=\max_{k\in\mathcal{K}}|w_{k}-w^{\prime}_{k}|,

and so 0T(w)T(w^)|𝒦|||Γ1Γ2||,0\leq T(w)-T(\hat{w})\leq|\mathcal{K}|\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|, 0T(w)T(w^)|𝒦|||Γ1Γ2||,0\leq T(w^{\prime})-T(\hat{w})\leq|\mathcal{K}|\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|, 0wkw^k||Γ1Γ2||0\leq w_{k}-\hat{w}_{k}\leq|\!|\Gamma_{1}-\Gamma_{2}|\!| and 0wkw^k||Γ1Γ2||0\leq w^{\prime}_{k}-\hat{w}_{k}\leq|\!|\Gamma_{1}-\Gamma_{2}|\!| for each k𝒦.k\in\mathcal{K}.

We assume now that Γ^\hat{\Gamma} is a game, i.e., ||Γ1Γ2||<T(w)|𝒦|.|\!|\Gamma_{1}-\Gamma_{2}|\!|<\frac{T(w)}{|\mathcal{K}|}.

Since the cost functions πa()\pi_{a}(\cdot) are Lipschitz continuous on the compact interval [0,T(w)][0,T(w)] with Lipschitz constant MΓ1,M_{\Gamma_{1}}, they are also Lipschitz continuous on the compact interval [0,T(w^)][0,T(w)][0,T(\hat{w})]\subseteq[0,T(w)] with the same Lipschitz constant MΓ1.M_{\Gamma_{1}}. Then Lemma 10a) and d) yield, respectively, that

|𝒞(Γ^)𝒞(Γ3)||A|T(w^)||Γ3Γ^||Γ3=(π^,w^)𝒢(G,𝒦,𝒮)|w^,|\mathcal{C}^{*}(\hat{\Gamma})-\mathcal{C}^{*}(\Gamma_{3})|\leq|A|\cdot T(\hat{w})\cdot|\!|\Gamma_{3}-\hat{\Gamma}|\!|\quad\forall\Gamma_{3}=(\hat{\pi},\hat{w})\in\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\hat{w}}, (4.13)

and that

|𝒞~(Γ^)𝒞~(Γ3)|(MΓ1|A|T(w^)+2)|A|T(w^)max{||Γ3Γ^||,||Γ3Γ^||}|\tilde{\mathcal{C}}(\hat{\Gamma})-\tilde{\mathcal{C}}(\Gamma_{3})|\leq(\sqrt{M_{\Gamma_{1}}\cdot|A|\cdot T(\hat{w})}+2)\cdot|A|\cdot T(\hat{w})\cdot\max\{\sqrt{|\!|\Gamma_{3}-\hat{\Gamma}|\!|},\ |\!|\Gamma_{3}-\hat{\Gamma}|\!|\} (4.14)

for every game Γ3=(π^,w^)𝒢(G,𝒦,𝒮)|w^.\Gamma_{3}=(\hat{\pi},\hat{w})\in\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\hat{w}}.

We now prove Lemma 11 with inequalities (4.13) and (4.14).

Proof.

Proof of Lemma 11a):

Consider an arbitrary WE flow f~\tilde{f} of the game Γ1=(π,w).\Gamma_{1}=(\pi,w). Since w^kwk\hat{w}_{k}\leq w_{k} for each k𝒦,k\in\mathcal{K}, there is a vector μ=(μs)s𝒮k\mu=(\mu_{s})_{s\in\mathcal{S}_{k}} such that 0μsf~s0\leq\mu_{s}\leq\tilde{f}_{s} for all s𝒮,s\in\mathcal{S}, and sSk(f~sμs)=w^k\sum_{s^{\prime}\in S_{k}}(\tilde{f}_{s^{\prime}}-\mu_{s^{\prime}})=\hat{w}_{k} for each k𝒦.k\in\mathcal{K}. Then f~μ=(f~sμs)s𝒮k\tilde{f}-\mu=(\tilde{f}_{s}-\mu_{s})_{s\in\mathcal{S}_{k}} is a WE flow of the game Γ1:=(π|μ,w^)𝒢(G,𝒦,𝒮)|w^,\Gamma^{\prime}_{1}:=(\pi_{|\mu},\hat{w})\in\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\hat{w}}, where π|μ:=(πa|μ)aA\pi_{|\mu}:=(\pi_{a|\mu})_{a\in A} with πa|μ(x):=πa(x+μa)\pi_{a|\mu}(x):=\pi_{a}(x+\mu_{a}) and μa:=s𝒮:asμs\mu_{a}:=\sum_{s\in\mathcal{S}:a\in s}\mu_{s} for each (a,x)A×[0,T(w^)].(a,x)\in A\times[0,T(\hat{w})]. This follows since the cost πs|μ(f~μ):=aA:asπa|μ(f~aμa)=aA:asπa(f~a)=πs(f~)\pi_{s|\mu}(\tilde{f}-\mu):=\sum_{a\in A:a\in s}\pi_{a|\mu}(\tilde{f}_{a}-\mu_{a})=\sum_{a\in A:a\in s}\pi_{a}(\tilde{f}_{a})=\pi_{s}(\tilde{f}) remains unchanged for each path s𝒮.s\in\mathcal{S}.

Inequality (4.14) then yields that

|𝒞~(Γ^)𝒞~(Γ1)|(MΓ1|A|T(w^)+2)|A|T(w^)max{||Γ1Γ^||,||Γ1Γ^||}.|\tilde{\mathcal{C}}(\hat{\Gamma})-\tilde{\mathcal{C}}(\Gamma^{\prime}_{1})|\leq(\sqrt{M_{\Gamma_{1}}\cdot|A|\cdot T(\hat{w})}+2)\cdot|A|\cdot T(\hat{w})\cdot\max\{\sqrt{|\!|\Gamma^{\prime}_{1}-\hat{\Gamma}|\!|},\ |\!|\Gamma^{\prime}_{1}-\hat{\Gamma}|\!|\}. (4.15)

Note that

|𝒞~(Γ1)𝒞~(Γ1)|=|C(Γ1,f~μ)C(Γ1,f~)|=k𝒦Lk(Γ1)(wkw^k)Lmax(Γ1)|𝒦|||Γ1Γ2||Lmax(Γ1)|𝒦|max{||Γ1Γ2||,||Γ1Γ2||}|A|πmax(T(w))|𝒦|max{||Γ1Γ2||,||Γ1Γ2||},\begin{split}|\tilde{\mathcal{C}}(\Gamma^{\prime}_{1})-\tilde{\mathcal{C}}(\Gamma_{1})|&=|C(\Gamma^{\prime}_{1},\tilde{f}-\mu)-C(\Gamma_{1},\tilde{f})|=\sum_{k\in\mathcal{K}}L_{k}(\Gamma_{1})\cdot(w_{k}-\hat{w}_{k})\\ &\leq L_{\max}(\Gamma_{1})\cdot|\mathcal{K}|\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|\leq L_{\max}(\Gamma_{1})\cdot|\mathcal{K}|\cdot\max\{\sqrt{|\!|\Gamma_{1}-\Gamma_{2}|\!|},\ |\!|\Gamma_{1}-\Gamma_{2}|\!|\}\\ &\leq|A|\cdot\pi_{max}(T(w))\cdot|\mathcal{K}|\cdot\max\{\sqrt{|\!|\Gamma_{1}-\Gamma_{2}|\!|},\ |\!|\Gamma_{1}-\Gamma_{2}|\!|\},\end{split} (4.16)

where we recall that Lk(Γ1)L_{k}(\Gamma_{1}) is the user cost of O/D pair k𝒦k\in\mathcal{K} of Γ1\Gamma_{1} in the WE flow f~,\tilde{f}, and Lmax(Γ1):=maxk𝒦Lk(Γ1)|A|maxaAπa(T(w))=|A|πmax(T(w)).L_{\max}(\Gamma_{1}):=\max_{k\in\mathcal{K}}\ L_{k}(\Gamma_{1})\leq|A|\cdot\max_{a\in A}\pi_{a}(T(w))=|A|\cdot\pi_{max}(T(w)). Note that

||Γ1Γ^||=maxaA,x[0,T(w^)]|πa(x)πa(x+μa)|MΓ1maxaAμaMΓ1||ww||MΓ1||Γ1Γ2||.|\!|\Gamma^{\prime}_{1}-\hat{\Gamma}|\!|=\max_{a\in A,x\in[0,T(\hat{w})]}|\pi_{a}(x)-\pi_{a}(x+\mu_{a})|\leq M_{\Gamma_{1}}\cdot\max_{a\in A}\ \mu_{a}\leq M_{\Gamma_{1}}\cdot|\!|w-w^{\prime}|\!|_{\infty}\leq M_{\Gamma_{1}}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|.

This together with (4.15) and (4.16) yields that

|𝒞~(Γ^)𝒞~(Γ1)|((MΓ1|A|T(w)+2)|A|T(w)+|A|πmax(T(w))|𝒦|)MΓ1max{||Γ1Γ2||,MΓ1||Γ1Γ2||},\begin{split}|\tilde{\mathcal{C}}(\hat{\Gamma})-\tilde{\mathcal{C}}(\Gamma_{1})|&\leq\Big{(}(\sqrt{M_{\Gamma_{1}}\cdot|A|\cdot T(w)}+2)\cdot|A|\cdot T(w)\\ &\hskip 14.22636pt+|A|\cdot\pi_{max}(T(w))\cdot|\mathcal{K}|\Big{)}\cdot\sqrt{M_{\Gamma_{1}}}\cdot\max\{\sqrt{|\!|\Gamma_{1}-\Gamma_{2}|\!|},\ \sqrt{M_{\Gamma_{1}}}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|\},\end{split} (4.17)

since T(w^)T(w).T(\hat{w})\leq T(w). Here, we observe that MΓ11.M_{\Gamma_{1}}\geq 1.

Similarly, we have

|𝒞~(Γ^)𝒞~(Γ2)|((MΓ1|A|T(w)+2)|A|T(w)+|A|πmax(T(w))|𝒦|)MΓ1max{||Γ2Γ1||,MΓ1||Γ2Γ1||},\begin{split}|\tilde{\mathcal{C}}(\hat{\Gamma})-\tilde{\mathcal{C}}(\Gamma_{2})|&\leq\Big{(}(\sqrt{M_{\Gamma_{1}}\cdot|A|\cdot T(w)}+2)\cdot|A|\cdot T(w)\\ &\hskip 14.22636pt+|A|\cdot\pi_{max}(T(w))\cdot|\mathcal{K}|\Big{)}\cdot\sqrt{M_{\Gamma_{1}}}\cdot\max\{\sqrt{|\!|\Gamma_{2}-\Gamma_{1}|\!|},\ \sqrt{M_{\Gamma_{1}}}\cdot|\!|\Gamma_{2}-\Gamma_{1}|\!|\},\end{split} (4.18)

since T(w)T(w).T(w^{\prime})\leq T(w). Inequalities (4.17)–(4.18) then yield

|𝒞~(Γ2)𝒞~(Γ1)|2((MΓ1|A|T(w)+2)|A|T(w)+|A|πmax(T(w))|𝒦|)MΓ1max{||Γ1Γ2||,MΓ1||Γ1Γ2||}.\begin{split}|\tilde{\mathcal{C}}(\Gamma_{2})-\tilde{\mathcal{C}}(\Gamma_{1})|&\leq 2\cdot\Big{(}(\sqrt{M_{\Gamma_{1}}\cdot|A|\cdot T(w)}+2)\cdot|A|\cdot T(w)\\ &\hskip 14.22636pt+|A|\cdot\pi_{max}(T(w))\cdot|\mathcal{K}|\Big{)}\cdot\sqrt{M_{\Gamma_{1}}}\cdot\max\{\sqrt{|\!|\Gamma_{1}-\Gamma_{2}|\!|},\ \sqrt{M_{\Gamma_{1}}}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|\}.\end{split}
Proof.

Proof of Lemma 11b):

Consider an arbitrary SO flow ff^{*} of Γ1.\Gamma_{1}. Similar to the proof of Lemma 11a), there is a vector λ=(λs)s𝒮\lambda=(\lambda_{s})_{s\in\mathcal{S}} such that 0λsfs0\leq\lambda_{s}\leq f_{s}^{*} for each s𝒮,s\in\mathcal{S}, and s𝒮k(fsλs)=w^k\sum_{s^{\prime}\in\mathcal{S}_{k}}(f_{s^{\prime}}^{*}-\lambda_{s^{\prime}})=\hat{w}_{k} for each k𝒦.k\in\mathcal{K}. Redefine Γ1:=(π|λ,w^)\Gamma^{\prime}_{1}:=(\pi_{|\lambda},\hat{w}) with π|λ:=(πa|λ)aA,\pi_{|\lambda}:=(\pi_{a|\lambda})_{a\in A}, πa|λ(x):=πa(x+λa)\pi_{a|\lambda}(x):=\pi_{a}(x+\lambda_{a}) and λa:=s𝒮:asλs\lambda_{a}:=\sum_{s\in\mathcal{S}:a\in s}\lambda_{s} for each (a,x)A×[0,T(w^)].(a,x)\in A\times[0,T(\hat{w})].

Then Γ1\Gamma^{\prime}_{1} is a game when ||Γ1Γ2||<T(w)|𝒦|,|\!|\Gamma_{1}-\Gamma_{2}|\!|<\frac{T(w)}{|\mathcal{K}|}, and fλ=(fsλs)s𝒮f^{*}-\lambda=(f^{*}_{s}-\lambda_{s})_{s\in\mathcal{S}} is a flow of Γ1,\Gamma^{\prime}_{1}, but need not be an SO flow of Γ1\Gamma^{\prime}_{1}.

Let ff be a flow of Γ1\Gamma_{1} such that fλf-\lambda is an SO flow of Γ1.\Gamma^{\prime}_{1}. Such a flow ff exists since f+λf^{*^{\prime}}+\lambda is a flow of Γ1\Gamma_{1} when ff^{*^{\prime}} is an SO flow of Γ1.\Gamma^{\prime}_{1}. Then

𝒞(Γ1)C(Γ1,f)=aAfaπa(fa)=aAλaπa(fa)+aA(faλa)πa(fa)|A||𝒦|πmax(T(w))||Γ1Γ2||+aA(faλa)πa(fa)=|A||𝒦|πmax(T(w))||Γ1Γ2||+aA(faλa)πa|λ(faλa)=|A||𝒦|πmax(T(w))||Γ1Γ2||+𝒞(Γ1)|A||𝒦|πmax(T(w))||Γ1Γ2||+C(Γ1,fλa)|A||𝒦|πmax(T(w))||Γ1Γ2||+C(Γ1,f)=|A||𝒦|πmax(T(w))||Γ1Γ2||+𝒞(Γ1).\begin{split}\mathcal{C}^{*}(\Gamma_{1})\leq C(\Gamma_{1},f)&=\sum_{a\in A}f_{a}\cdot\pi_{a}(f_{a})=\sum_{a\in A}\lambda_{a}\cdot\pi_{a}(f_{a})+\sum_{a\in A}(f_{a}-\lambda_{a})\cdot\pi_{a}(f_{a})\\ &\leq|A|\cdot|\mathcal{K}|\cdot\pi_{max}(T(w))\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|+\sum_{a\in A}(f_{a}-\lambda_{a})\cdot\pi_{a}(f_{a})\\ &=|A|\cdot|\mathcal{K}|\cdot\pi_{max}(T(w))\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|+\sum_{a\in A}(f_{a}-\lambda_{a})\cdot\pi_{a|\lambda}(f_{a}-\lambda_{a})\\ &=|A|\cdot|\mathcal{K}|\cdot\pi_{max}(T(w))\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|+\mathcal{C}^{*}(\Gamma^{\prime}_{1})\\ &\leq|A|\cdot|\mathcal{K}|\cdot\pi_{max}(T(w))\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|+C(\Gamma^{\prime}_{1},f^{*}-\lambda_{a})\\ &\leq|A|\cdot|\mathcal{K}|\cdot\pi_{max}(T(w))\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|+C(\Gamma_{1},f^{*})\\ &=|A|\cdot|\mathcal{K}|\cdot\pi_{max}(T(w))\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|+\mathcal{C}^{*}(\Gamma_{1}).\end{split}

Here, we have used λa=k𝒦s𝒮k:asλsk𝒦s𝒮kλs=k𝒦(wkw^k)|𝒦|||ww^|||𝒦|||ww|||𝒦|||Γ1Γ2||.\lambda_{a}=\sum_{k\in\mathcal{K}}\sum_{s\in\mathcal{S}_{k}:a\in s}\lambda_{s}\leq\sum_{k\in\mathcal{K}}\sum_{s\in\mathcal{S}_{k}}\lambda_{s}=\sum_{k\in\mathcal{K}}(w_{k}-\hat{w}_{k})\leq|\mathcal{K}|\cdot|\!|w-\hat{w}|\!|_{\infty}\leq|\mathcal{K}|\cdot|\!|w-w^{\prime}|\!|_{\infty}\leq|\mathcal{K}|\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|. Then we obtain

0𝒞(Γ1)𝒞(Γ1)|A||𝒦|πmax(T(w))||Γ1Γ2||.0\leq\mathcal{C}^{*}(\Gamma_{1})-\mathcal{C}^{*}(\Gamma^{\prime}_{1})\leq|A|\cdot|\mathcal{K}|\cdot\pi_{max}(T(w))\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|. (4.19)

Inequality (4.13) yields

|𝒞(Γ1)𝒞(Γ^)|<|A|T(w)MΓ1||Γ1Γ2||,|\mathcal{C}^{*}(\Gamma^{\prime}_{1})-\mathcal{C}^{*}(\hat{\Gamma})|<|A|\cdot T(w)\cdot M_{\Gamma_{1}}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|, (4.20)

where we observe that T(w^)T(w)T(\hat{w})\leq T(w) and ||Γ1Γ^||MΓ1||Γ1Γ2||.|\!|\Gamma^{\prime}_{1}-\hat{\Gamma}|\!|\leq M_{\Gamma_{1}}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|. Inequalities (4.19) and (4.20) together imply

|𝒞(Γ^𝒞(Γ1)|(|A||𝒦|πmax(T(w))+|A|T(w)MΓ1)||Γ1Γ2||.|\mathcal{C}^{*}(\hat{\Gamma}-\mathcal{C}^{*}(\Gamma_{1})|\leq\big{(}|A|\cdot|\mathcal{K}|\cdot\pi_{max}(T(w))+|A|\cdot T(w)\cdot M_{\Gamma_{1}}\big{)}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|. (4.21)

Similar to the proof of Lemma 11a), we have

|𝒞(Γ^)𝒞(Γ2)|(|A||𝒦|πmax(T(w))+|A|T(w)MΓ1)||Γ1Γ2||,|\mathcal{C}^{*}(\hat{\Gamma})-\mathcal{C}^{*}(\Gamma_{2})|\leq\big{(}|A|\cdot|\mathcal{K}|\cdot\pi_{max}(T(w))+|A|\cdot T(w)\cdot M_{\Gamma_{1}}\big{)}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|, (4.22)

since T(w)T(w).T(w^{\prime})\leq T(w). Lemma 11b) then follows immediately from (4.21) and (4.22).

Proof.

Proof of Lemma 11c): Lemma 11a)–b) imply that

|ρ(Γ1)ρ(Γ2)|𝒞~(Γ1)+ν~Γ1max{||Γ1Γ2||,MΓ1||Γ1Γ2||}𝒞(Γ1)νΓ1||Γ1Γ2||𝒞~(Γ1)𝒞(Γ1)=ρ(Γ1)νΓ1||Γ1Γ2||+ν~Γ1max{||Γ1Γ2||,MΓ1||Γ1Γ2||}𝒞(Γ1)νΓ1||Γ1Γ2||2ρ(Γ1)νΓ1+2ν~Γ1𝒞(Γ1)max{||Γ1Γ2||,MΓ1||Γ1Γ2||}\begin{split}|\rho(\Gamma_{1})-\rho(\Gamma_{2})|&\leq\frac{\tilde{\mathcal{C}}(\Gamma_{1})+\tilde{\nu}_{\Gamma_{1}}\cdot\max\{\sqrt{|\!|\Gamma_{1}-\Gamma_{2}|\!|},\sqrt{M_{\Gamma_{1}}}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|\}}{\mathcal{C}^{*}(\Gamma_{1})-\nu^{*}_{\Gamma_{1}}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|}-\frac{\tilde{\mathcal{C}}(\Gamma_{1})}{\mathcal{C}^{*}(\Gamma_{1})}\\ &=\frac{\rho(\Gamma_{1})\cdot\nu^{*}_{\Gamma_{1}}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|+\tilde{\nu}_{\Gamma_{1}}\cdot\max\{\sqrt{|\!|\Gamma_{1}-\Gamma_{2}|\!|},\sqrt{M_{\Gamma_{1}}}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|\}}{\mathcal{C}^{*}(\Gamma_{1})-\nu^{*}_{\Gamma_{1}}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|}\\ &\leq\frac{2\cdot\rho(\Gamma_{1})\cdot\nu^{*}_{\Gamma_{1}}+2\cdot\tilde{\nu}_{\Gamma_{1}}}{\mathcal{C}^{*}(\Gamma_{1})}\cdot\max\{\sqrt{|\!|\Gamma_{1}-\Gamma_{2}|\!|},\sqrt{M_{\Gamma_{1}}}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|\}\end{split}

when νΓ1||Γ1Γ2||𝒞(Γ1)2\nu^{*}_{\Gamma_{1}}\cdot|\!|\Gamma_{1}-\Gamma_{2}|\!|\leq\frac{\mathcal{C}^{*}(\Gamma_{1})}{2} and ||Γ1Γ2||<T(w)|𝒦|.|\!|\Gamma_{1}-\Gamma_{2}|\!|<\frac{T(w)}{|\mathcal{K}|}.

This completes the proof of Lemma 11.

\square

While Lemma 11 shows a weaker Hölder exponent 12\frac{1}{2} than the exponent 11 of Englert et al. [2010] and Takalloo and Kwon [2020], it applies to more general cases. On the one hand, two games Γ1=(π,w)\Gamma_{1}=(\pi,w) and Γ2=(π,w)\Gamma_{2}=(\pi,w^{\prime}) from the same cost slice 𝒢(G,𝒦,𝒮)|π\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\pi} need only satisfy the weaker condition T(w)T(w)T(w)\geq T(w^{\prime}) instead of the stronger condition “w=(1+ϵ)ww=(1+\epsilon)\cdot w^{\prime}”. On the other hand, Lemma 11 applies to a cost slice 𝒢(G,𝒦,𝒮)|π\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\pi} with arbitrary continuously differentiable cost functions, as it assumes only Lipschitz continuity of the cost functions on the compact interval [0,T(w)],[0,T(w)], but not on the whole unbounded interval [0,).[0,\infty).

The proof of Lemma 11 essentially builds on Lemma 10. Since we cannot improve the Hölder exponent 12\frac{1}{2} of Lemma 10 for demand slices (see Remark 4.2), we are presently also unable to do that for cost slices in Lemma 11, although the PoA might have stronger Hölder continuity properties on cost slices than demand slices (see Remark 4.1). While Englert et al. [2010], Takalloo and Kwon [2020] and Cominetti et al. [2020] have already proposed independent techniques to analyze the Hölder continuity of the PoA on cost slices for some inspiring cases, we are still eager for a general technique to independently analyze the Hölder continuity of the PoA on cost slices with arbitrary Lipschitz continuous cost functions on [0,T(d)].[0,T(d)]. Nonetheless, we will see in Section 4.3 that the current technical framework used in the proof of Lemma 11 yields a stronger Hölder exponent of 11 for cost slices when the cost functions have good properties similar to those in Englert et al. [2010], Englert et al. [2010] and Cominetti et al. [2020].

4.2.3 Proof of Theorem 9

We now combine Lemma 10 and Lemma 11 to prove Theorem 9. Consider an arbitrary game Γ=(τ,d)𝒢(G,𝒦,𝒮)\Gamma=(\tau,d)\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) whose cost functions τa(x)\tau_{a}(x) are Lipschitz continuous on [0,T(d)][0,T(d)] with a Lipschitz constant MΓ>0.M_{\Gamma}>0.

For an arbitrary game Γ=(σ,d)𝒢(G,𝒦,𝒮),\Gamma^{\prime}=(\sigma,d^{\prime})\in\mathcal{G}(G,\mathcal{K},\mathcal{S}), we define an auxiliary game Γ=(τ^,d)\Gamma^{\prime\prime}=(\hat{\tau},d^{\prime}) with

τ^a(x)={τa(x)if xTmin:=min{T(d),T(d)}τa(Tmin)if x>Tmin.aA.\hat{\tau}_{a}(x)=\begin{cases}\tau_{a}(x)&\text{if }x\leq T_{min}:=\min\{T(d),T(d^{\prime})\}\\ \tau_{a}(T_{\min})&\text{if }x>T_{min}.\end{cases}\quad\forall a\in A. (4.23)

Trivially, each auxiliary cost function τ^a(x)\hat{\tau}_{a}(x) is Lipschitz continuous on [0,T(d)][0,T(d^{\prime})] with the same Lipschitz constant MΓ,M_{\Gamma}, since τa()\tau_{a}(\cdot) is non-decreasing. Moreover, we obtain by (3.3) and (3.5) that

||ΓΓ||=max{||dd||,||τ(Tmin)τ(T(d))||}||ΓΓ||=max{maxaA,x[0,Tmin]|τa(x)σa(x)|,||τ(Tmin)σ(T(d))||}.\begin{split}&|\!|\Gamma-\Gamma^{\prime\prime}|\!|=\max\left\{|\!|d-d^{\prime}|\!|_{\infty},|\!|\tau(T_{min})-\tau(T(d))|\!|_{\infty}\right\}\\ &|\!|\Gamma^{\prime\prime}-\Gamma^{\prime}|\!|=\max\left\{\max_{a\in A,x\in[0,T_{min}]}|\tau_{a}(x)-\sigma_{a}(x)|,|\!|\tau(T_{min})-\sigma(T(d^{\prime}))|\!|_{\infty}\right\}.\end{split} (4.24)

Again by (3.3) and (3.5), we have ||τ(Tmin)τ(T(d))||MΓ|T(d)T(d)|MΓ|𝒦|||dd||MΓ|𝒦|||ΓΓ||,|\!|\tau(T_{min})-\tau(T(d))|\!|_{\infty}\leq M_{\Gamma}\cdot|T(d^{\prime})-T(d)|\leq M_{\Gamma}\cdot|\mathcal{K}|\cdot|\!|d-d^{\prime}|\!|_{\infty}\leq M_{\Gamma}\cdot|\mathcal{K}|\cdot|\!|\Gamma-\Gamma^{\prime}|\!|, and

||τ(Tmin)σ(T(d))||||τ(Tmin)τ(T(d))||+||τ(T(d))σ(T(d))||(MΓ|𝒦|+1)||ΓΓ||.|\!|\tau(T_{min})-\sigma(T(d^{\prime}))|\!|_{\infty}\leq|\!|\tau(T_{min})-\tau(T(d))|\!|_{\infty}+|\!|\tau(T(d))-\sigma(T(d^{\prime}))|\!|_{\infty}\leq(M_{\Gamma}\cdot|\mathcal{K}|+1)\cdot|\!|\Gamma-\Gamma^{\prime}|\!|.

This together with (4.24) yields

||ΓΓ||max{MΓ|𝒦|,1}||ΓΓ|| and ||ΓΓ||(MΓ|𝒦|+1)||ΓΓ||.|\!|\Gamma-\Gamma^{\prime\prime}|\!|\leq\max\{M_{\Gamma}\cdot|\mathcal{K}|,1\}\cdot|\!|\Gamma-\Gamma^{\prime}|\!|\text{ and }|\!|\Gamma^{\prime}-\Gamma^{\prime\prime}|\!|\leq(M_{\Gamma}\cdot|\mathcal{K}|+1)\cdot|\!|\Gamma-\Gamma^{\prime}|\!|. (4.25)

So both ||ΓΓ|||\!|\Gamma-\Gamma^{\prime\prime}|\!| and ||ΓΓ|||\!|\Gamma^{\prime}-\Gamma^{\prime\prime}|\!| converge to 0 as ||ΓΓ|||\!|\Gamma-\Gamma^{\prime}|\!| tends to 0.0.

Since Γ=(σ,d)\Gamma^{\prime}=(\sigma,d^{\prime}) and Γ=(τ^,d)\Gamma^{\prime\prime}=(\hat{\tau},d^{\prime}) belong the same demand slice 𝒢(G,𝒦,𝒮)|d,\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|d^{\prime}}, and since the auxiliary cost functions τ^\hat{\tau} are Lipschitz continuous on [0,T(d)][0,T(d^{\prime})] with Lipschitz constant MΓ,M_{\Gamma}, Lemma 10d) implies that

|ρ(Γ)ρ(Γ)|νΓmax{||ΓΓ||,||ΓΓ||}|\rho(\Gamma^{\prime\prime})-\rho(\Gamma^{\prime})|\leq\nu_{\Gamma^{\prime\prime}}\cdot\max\{\sqrt{|\!|\Gamma^{\prime\prime}-\Gamma^{\prime}|\!|},|\!|\Gamma^{\prime\prime}-\Gamma^{\prime}|\!|\} (4.26)

with a constant νΓ>0\nu_{\Gamma^{\prime\prime}}>0 depending only on Γ,\Gamma^{\prime\prime}, when ||ΓΓ|||\!|\Gamma^{\prime}-\Gamma^{\prime\prime}|\!| is small enough. Moreover, Lemma 10d) and Theorem 6 together imply that the constant νΓ\nu_{\Gamma^{\prime\prime}} converges to a constant νΓ>0\nu^{\prime}_{\Gamma}>0 depending only on Γ\Gamma as ΓΓ\Gamma^{\prime}\to\Gamma (which implies ΓΓ\Gamma^{\prime\prime}\to\Gamma by (4.25)), since the constant νΓ\nu_{\Gamma^{\prime\prime}} depends only on MΓ,M_{\Gamma}, T(d),T(d^{\prime}), 𝒞(Γ),\mathcal{C}^{*}(\Gamma^{\prime\prime}), and ρ(Γ),\rho(\Gamma^{\prime\prime}), and since the Lipschitz constant MΓM_{\Gamma} does not change in the limit. This together with (4.25) implies that there is a small constant ϵ1,Γ>0\epsilon_{1,\Gamma}>0 such that

|ρ(Γ)ρ(Γ)|<2νΓ(MΓ|𝒦|+1)||ΓΓ|||\rho(\Gamma^{\prime})-\rho(\Gamma^{\prime\prime})|<2\cdot\nu^{\prime}_{\Gamma}\cdot\sqrt{(M_{\Gamma}\cdot|\mathcal{K}|+1)\cdot|\!|\Gamma-\Gamma^{\prime}|\!|} (4.27)

when ||ΓΓ||<ϵ1,Γ.|\!|\Gamma-\Gamma^{\prime}|\!|<\epsilon_{1,\Gamma}.

When T(d)T(d),T(d^{\prime})\geq T(d), we obtain by the definition (4.23) of the auxiliary cost functions τ^a\hat{\tau}_{a} that Γ=(τ,d)=(τ^,d)\Gamma=(\tau,d)=(\hat{\tau},d) w.r.t. equivalence relation (3.1). Then Γ\Gamma and Γ\Gamma^{\prime\prime} are from the same cost slice 𝒢(G,𝒦,𝒮)|τ^\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\hat{\tau}} with T(d)T(d).T(d)\leq T(d^{\prime}). Lemma 11c) then yields

|ρ(Γ)ρ(Γ)|νΓmax{||ΓΓ||,||ΓΓ||}|\rho(\Gamma)-\rho(\Gamma^{\prime\prime})|\leq\nu^{\prime}_{\Gamma^{\prime\prime}}\cdot\max\{\sqrt{|\!|\Gamma-\Gamma^{\prime\prime}|\!|},|\!|\Gamma-\Gamma^{\prime\prime}|\!|\} (4.28)

for a constant νΓ>0\nu^{\prime}_{\Gamma^{\prime\prime}}>0 depending on Γ,\Gamma^{\prime\prime}, when ||ΓΓ|||\!|\Gamma-\Gamma^{\prime\prime}|\!| is small enough. Similar to the previous arguments, we obtain by Theorem 6 and Lemma 11c) that this constant νΓ\nu^{\prime}_{\Gamma^{\prime\prime}} converges also to a constant νΓ\nu^{\prime\prime}_{\Gamma} as ΓΓ,\Gamma^{\prime}\to\Gamma, and so

|ρ(Γ)ρ(Γ)|2νΓmax{MΓ|𝒦|,1}||ΓΓ|||\rho(\Gamma)-\rho(\Gamma^{\prime\prime})|\leq 2\cdot\nu^{\prime\prime}_{\Gamma}\cdot\sqrt{\max\{M_{\Gamma}\cdot|\mathcal{K}|,1\}\cdot|\!|\Gamma-\Gamma^{\prime}|\!|} (4.29)

when ||ΓΓ||<ϵ2,Γ|\!|\Gamma-\Gamma^{\prime}|\!|<\epsilon_{2,\Gamma} for a small constant ϵ2,Γ>0.\epsilon_{2,\Gamma}>0.

When T(d)<T(d),T(d^{\prime})<T(d), then Γ=(τ^,d)=(τ,d)\Gamma^{\prime\prime}=(\hat{\tau},d^{\prime})=(\tau,d^{\prime}) by equivalence relation (3.1). Then Γ\Gamma and Γ\Gamma^{\prime\prime} are from the same cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} with T(d)<T(d).T(d^{\prime})<T(d). Lemma 11c) then yields

|ρ(Γ)ρ(Γ)|νΓmax{||ΓΓ||,||ΓΓ||}νΓmax{MΓ|𝒦|,1}||ΓΓ|||\rho(\Gamma)-\rho(\Gamma^{\prime\prime})|\leq\nu^{\prime\prime\prime}_{\Gamma}\cdot\max\{\sqrt{|\!|\Gamma-\Gamma^{\prime\prime}|\!|},|\!|\Gamma-\Gamma^{\prime\prime}|\!|\}\leq\nu^{\prime\prime\prime}_{\Gamma}\cdot\sqrt{\max\{M_{\Gamma}\cdot|\mathcal{K}|,1\}\cdot|\!|\Gamma-\Gamma^{\prime}|\!|} (4.30)

for a constant νΓ>0\nu^{\prime\prime\prime}_{\Gamma}>0 depending only on Γ,\Gamma, when ||ΓΓ||<ϵ3,Γ|\!|\Gamma-\Gamma^{\prime}|\!|<\epsilon_{3,\Gamma} for a small constant ϵ3,Γ>0\epsilon_{3,\Gamma}>0.

Inequalities (4.27), (4.29) and (4.30) together then imply Theorem 9, since |ρ(Γ)ρ(Γ)||ρ(Γ)ρ(Γ)|+|ρ(Γ)ρ(Γ)||\rho(\Gamma)-\rho(\Gamma^{\prime})|\leq|\rho(\Gamma)-\rho(\Gamma^{\prime\prime})|+|\rho(\Gamma^{\prime\prime})-\rho(\Gamma^{\prime})| and the arbitrary choice of Γ.\Gamma^{\prime}.

This completes the proof of Theorem 9.

\square

4.3 Hölder continuity of the PoA for special cost functions

Although Theorem 8 shows that the PoA map ρ()\rho(\cdot) is not Lipschitz continuous on the whole game space 𝒢(G,𝒦,𝒮),\mathcal{G}(G,\mathcal{K},\mathcal{S}), the differentiability results of Cominetti et al. [2020] seemingly suggest that the map ρ()\rho(\cdot) might be pointwise Lipschitz continuous (i.e., pointwise Hölder continuous with Hölder exponent 11) at each game Γ=(τ,d)\Gamma=(\tau,d) whose cost functions τa()\tau_{a}(\cdot) have strictly positive derivatives on the compact interval [0,T(d)].[0,T(d)]. Theorem 12 confirms this.

Theorem 12.

Consider an arbitrary game Γ=(τ,d)𝒢(G,𝒦,𝒮).\Gamma=(\tau,d)\in\mathcal{G}(G,\mathcal{K},\mathcal{S}).

  • a)

    If τa(x)\tau_{a}(x) is constant for all aAa\in A and x[0,T(d)],x\in[0,T(d)], then there are a small ϵΓ>0\epsilon_{\Gamma}>0 and a Hölder constant ϰΓ>0\varkappa_{\Gamma}>0 depending only on Γ\Gamma such that |ρ(Γ)ρ(Γ)|<ϰΓ||ΓΓ|||\rho(\Gamma)-\rho(\Gamma^{\prime})|<\varkappa_{\Gamma}\cdot|\!|\Gamma-\Gamma^{\prime}|\!| for every game Γ𝒢(G,𝒦,𝒮)\Gamma^{\prime}\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) with ||ΓΓ||<ϵΓ.|\!|\Gamma-\Gamma^{\prime}|\!|<\epsilon_{\Gamma}.

  • b)

    If τa()\tau_{a}(\cdot) is continuously differentiable on [0,T(d)][0,T(d)] and τa(x)mΓ\tau^{\prime}_{a}(x)\geq m_{\Gamma} for all aAa\in A and x[0,T(d)]x\in[0,T(d)] for a constant mΓ>0m_{\Gamma}>0 depending only on Γ,\Gamma, then there are a small ϵΓ>0\epsilon_{\Gamma}>0 and a Hölder constant ϰΓ>0\varkappa_{\Gamma}>0 depending only on Γ\Gamma such that |ρ(Γ)ρ(Γ)|<ϰΓ||ΓΓ|||\rho(\Gamma)-\rho(\Gamma^{\prime})|<\varkappa_{\Gamma}\cdot|\!|\Gamma-\Gamma^{\prime}|\!| for every game Γ𝒢(G,𝒦,𝒮)\Gamma^{\prime}\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) with ||ΓΓ||<ϵΓ.|\!|\Gamma-\Gamma^{\prime}|\!|<\epsilon_{\Gamma}.

Proof.

Proof of Theorem 12

Proof.

Proof of Theorem 12a): Assume that τa(x)τa(0)\tau_{a}(x)\equiv\tau_{a}(0) for all aAa\in A and x[0,T(d)].x\in[0,T(d)]. Trivially, the cost functions τa()\tau_{a}(\cdot) are Lipschitz continuous on [0,T(d)][0,T(d)] with Lipschitz constant MΓ=1.M_{\Gamma}=1.

Let Γ=(σ,d)𝒢(G,𝒦,𝒮)\Gamma^{\prime}=(\sigma,d^{\prime})\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) be an arbitrary game. Define an auxiliary cost function vector τ^=(τ^a)aA\hat{\tau}=(\hat{\tau}_{a})_{a\in A} with τ^a(x):=τa(0)\hat{\tau}_{a}(x):=\tau_{a}(0) for all aAa\in A and x[0,).x\in[0,\infty). Let Γ=(τ^,d)\Gamma^{\prime\prime}=(\hat{\tau},d^{\prime}) be the corresponding auxiliary game. Then ρ(Γ)=ρ(Γ)=1,\rho(\Gamma)=\rho(\Gamma^{\prime\prime})=1, inequality (4.25) holds and ΓΓ\Gamma^{\prime\prime}\to\Gamma as ||ΓΓ||0.|\!|\Gamma-\Gamma^{\prime}|\!|\to 0.

Lemma 10a) and (4.25) yield that

|𝒞(Γ)𝒞(Γ)|<|A|T(d)(|𝒦|+1)||ΓΓ||2|A|T(d)(|𝒦|+1)||ΓΓ|||\mathcal{C}^{*}(\Gamma^{\prime\prime})-\mathcal{C}^{*}(\Gamma^{\prime})|<|A|\cdot T(d^{\prime})\cdot(|\mathcal{K}|+1)\cdot|\!|\Gamma-\Gamma^{\prime}|\!|\leq 2\cdot|A|\cdot T(d)\cdot(|\mathcal{K}|+1)\cdot|\!|\Gamma-\Gamma^{\prime}|\!| (4.31)

when T(d)2T(d).T(d^{\prime})\leq 2\cdot T(d).

Let f~\tilde{f} and g~\tilde{g} be a WE flow of Γ=(τ^,d)\Gamma^{\prime\prime}=(\hat{\tau},d^{\prime}) and a WE flow of Γ=(σ,d),\Gamma^{\prime}=(\sigma,d^{\prime}), respectively. We then have

𝒞~(Γ)𝒞~(Γ)=C(Γ,g~)C(Γ,f~)=aAσa(g~a)g~aaAτa(0)f~aaA(σa(g~a)τa(0))f~a|A|T(d)||ΓΓ||<2|A|T(d)(|𝒦|+1)||ΓΓ||\begin{split}\tilde{\mathcal{C}}(\Gamma^{\prime})&-\tilde{\mathcal{C}}(\Gamma^{\prime\prime})=C(\Gamma^{\prime},\tilde{g})-C(\Gamma^{\prime\prime},\tilde{f})=\sum_{a\in A}\sigma_{a}(\tilde{g}_{a})\cdot\tilde{g}_{a}-\sum_{a\in A}\tau_{a}(0)\cdot\tilde{f}_{a}\\ &\leq\sum_{a\in A}\big{(}\sigma_{a}(\tilde{g}_{a})-\tau_{a}(0)\big{)}\cdot\tilde{f}_{a}\leq|A|\cdot T(d^{\prime})\cdot|\!|\Gamma^{\prime}-\Gamma|\!|<2\cdot|A|\cdot T(d)\cdot(|\mathcal{K}|+1)\cdot|\!|\Gamma-\Gamma^{\prime}|\!|\end{split} (4.32)

when T(d)2T(d).T(d^{\prime})\leq 2\cdot T(d). Inequalities (4.31) and (4.32) together imply that

0ρ(Γ)ρ(Γ)=ρ(Γ)ρ(Γ)=𝒞~(Γ)𝒞(Γ)𝒞~(Γ)𝒞(Γ)𝒞~(Γ)+2|A|T(d)(|𝒦|+1)||ΓΓ||𝒞(Γ)2|A|T(d)(|𝒦|+1)||ΓΓ||𝒞~(Γ)𝒞(Γ)<8|A|T(d)(|𝒦|+1)𝒞(Γ)||ΓΓ||\begin{split}0\leq\rho(\Gamma^{\prime})-\rho(\Gamma)&=\rho(\Gamma^{\prime})-\rho(\Gamma^{\prime\prime})=\frac{\tilde{\mathcal{C}}(\Gamma^{\prime})}{\mathcal{C}^{*}(\Gamma^{\prime})}-\frac{\tilde{\mathcal{C}}(\Gamma^{\prime\prime})}{\mathcal{C}^{*}(\Gamma^{\prime\prime})}\\ &\leq\frac{\tilde{\mathcal{C}}(\Gamma^{\prime\prime})+2\cdot|A|\cdot T(d)\cdot(|\mathcal{K}|+1)\cdot|\!|\Gamma-\Gamma^{\prime}|\!|}{\mathcal{C}^{*}(\Gamma^{\prime\prime})-2\cdot|A|\cdot T(d)\cdot(|\mathcal{K}|+1)\cdot|\!|\Gamma-\Gamma^{\prime}|\!|}-\frac{\tilde{\mathcal{C}}(\Gamma^{\prime\prime})}{\mathcal{C}^{*}(\Gamma^{\prime\prime})}\\ &<\frac{8\cdot|A|\cdot T(d)\cdot(|\mathcal{K}|+1)}{\mathcal{C}^{*}(\Gamma)}\cdot|\!|\Gamma-\Gamma^{\prime}|\!|\end{split} (4.33)

when ||ΓΓ|||\!|\Gamma-\Gamma^{\prime}|\!| is small. Here, we notice that T(d)<2T(d)T(d^{\prime})<2\cdot T(d) holds when ||ΓΓ|||\!|\Gamma-\Gamma^{\prime}|\!| is small. This completes the proof of Theorem 12a).

Proof.

Proof of Theorem 12b): Assume now that τa()\tau_{a}(\cdot) is continuously differentiable on [0,T(d)][0,T(d)] and τa(x)mΓ\tau^{\prime}_{a}(x)\geq m_{\Gamma} for all aAa\in A and x[0,T(d)]x\in[0,T(d)] for a constant mΓ>0m_{\Gamma}>0 depending only on Γ.\Gamma. Trivially, τa()\tau_{a}(\cdot) is Lipschitz continuous on [0,T(d)][0,T(d)] with Lipschitz constant MΓ:=maxbAmaxx[0,1]τb(x)mΓ>0M_{\Gamma}:=\max_{b\in A}\max_{x\in[0,1]}\ \tau^{\prime}_{b}(x)\geq m_{\Gamma}>0 for each arc aA.a\in A.

Let Γ=(σ,d)𝒢(G,𝒦,𝒮)\Gamma^{\prime}=(\sigma,d^{\prime})\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) be an arbitrary game. Define an auxiliary cost function vector τ^\hat{\tau} with

τ^a(x):=τa(x)aAx[0,T(d)]\hat{\tau}_{a}(x):=\tau_{a}(x)\quad\forall a\in A\ \forall x\in[0,T(d^{\prime})] (4.34)

when T(d)T(d),T(d)\geq T(d^{\prime}), and

τ^a(x):={τa(x)if xT(d)τa(T(d))+τa(T(d))(xT(d))if x(T(d),T(d)]aAx[0,T(d)]\hat{\tau}_{a}(x):=\begin{cases}\tau_{a}(x)&\text{if }x\leq T(d)\\ \tau_{a}(T(d))+\tau^{\prime}_{a}(T(d))\cdot(x-T(d))&\text{if }x\in(T(d),T(d^{\prime})]\end{cases}\quad\forall a\in A\ \forall x\in[0,T(d^{\prime})] (4.35)

when T(d)>T(d).T(d^{\prime})>T(d). Denote by Γ\Gamma^{\prime\prime} the game (τ^,d).(\hat{\tau},d^{\prime}). Then the cost functions τ^a\hat{\tau}_{a} are continuously differentiable on [0,T(d)],[0,T(d^{\prime})], and 0<mΓminaA,x[0,T(d)]τ^a(x)maxaA,x[0,T(d)]τ^a(x)MΓ.0<m_{\Gamma}\leq\min_{a\in A,x\in[0,T(d^{\prime})]}\hat{\tau}^{\prime}_{a}(x)\leq\max_{a\in A,x\in[0,T(d^{\prime})]}\hat{\tau}^{\prime}_{a}(x)\leq M_{\Gamma}. Similar to (4.25), we have

||ΓΓ||=||σ|T(d)τ^|T(d)||||ΓΓ||+MΓ|T(d)T(d)|(1+MΓ|𝒦|)||ΓΓ||||ΓΓ||=max{||dd||,||τ|T(d)τ^|T(d)||}||ΓΓ||+MΓ|T(d)T(d)|(1+MΓ|𝒦|)||ΓΓ||.\begin{split}|\!|\Gamma^{\prime\prime}-\Gamma^{\prime}|\!|&=|\!|\sigma_{|T(d^{\prime})}-\hat{\tau}_{|T(d^{\prime})}|\!|_{\infty}\leq|\!|\Gamma-\Gamma^{\prime}|\!|+M_{\Gamma}\cdot|T(d^{\prime})-T(d)|\leq(1+M_{\Gamma}\cdot|\mathcal{K}|)\cdot|\!|\Gamma-\Gamma^{\prime}|\!|\\ |\!|\Gamma-\Gamma^{\prime\prime}|\!|&=\max\{|\!|d-d^{\prime}|\!|_{\infty},|\!|\tau_{|T(d)}-\hat{\tau}_{|T(d^{\prime})}|\!|_{\infty}\}\leq|\!|\Gamma-\Gamma^{\prime}|\!|+M_{\Gamma}\cdot|T(d^{\prime})-T(d)|\\ &\hskip 56.9055pt\leq(1+M_{\Gamma}\cdot|\mathcal{K}|)\cdot|\!|\Gamma-\Gamma^{\prime}|\!|.\end{split} (4.36)

Lemma 10a) and inequality (4.36) then yield

|𝒞(Γ)𝒞(Γ)|<|A|T(d)(MΓ|𝒦|+1)||ΓΓ||2|A|T(d)(MΓ|𝒦|+1)||ΓΓ||,|\mathcal{C}^{*}(\Gamma^{\prime\prime})-\mathcal{C}^{*}(\Gamma^{\prime})|<|A|\cdot T(d^{\prime})\cdot(M_{\Gamma}\cdot|\mathcal{K}|+1)\cdot|\!|\Gamma-\Gamma^{\prime}|\!|\leq 2\cdot|A|\cdot T(d)\cdot(M_{\Gamma}\cdot|\mathcal{K}|+1)\cdot|\!|\Gamma-\Gamma^{\prime}|\!|, (4.37)

when ||ΓΓ|||\!|\Gamma-\Gamma^{\prime}|\!| is small, since the cost functions τ^a\hat{\tau}_{a} are Lipschitz continuous on [0,T(d)][0,T(d^{\prime})] with Lipschitz constant MΓ,M_{\Gamma}, and since both Γ=(σ,d)\Gamma^{\prime}=(\sigma,d^{\prime}) and Γ=(τ^,d)\Gamma^{\prime\prime}=(\hat{\tau},d^{\prime}) are from the same demand slice 𝒢(G,𝒦,𝒮)|d.\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|d^{\prime}}.

Let f~\tilde{f} and g~\tilde{g} be a WE flow of Γ\Gamma^{\prime\prime} and a WE flow of Γ,\Gamma^{\prime}, respectively. Lemma 10c) and Lemma 1b) together yield that

0aAτ^a(f~a)(g~af~a)aAτ^a(g~a)(g~af~a)|A|T(d)||ΓΓ||,0aAσa(g~a)(f~ag~a)aAσa(f~a)(f~ag~a)|A|T(d)||ΓΓ||,\begin{split}&0\leq\sum_{a\in A}\hat{\tau}_{a}(\tilde{f}_{a})\cdot(\tilde{g}_{a}-\tilde{f}_{a})\leq\sum_{a\in A}\hat{\tau}_{a}(\tilde{g}_{a})\cdot(\tilde{g}_{a}-\tilde{f}_{a})\leq|A|\cdot T(d^{\prime})\cdot|\!|\Gamma^{\prime}-\Gamma^{\prime\prime}|\!|,\\ &0\leq\sum_{a\in A}\sigma_{a}(\tilde{g}_{a})\cdot(\tilde{f}_{a}-\tilde{g}_{a})\leq\sum_{a\in A}\sigma_{a}(\tilde{f}_{a})\cdot(\tilde{f}_{a}-\tilde{g}_{a})\leq|A|\cdot T(d^{\prime})\cdot|\!|\Gamma^{\prime}-\Gamma^{\prime\prime}|\!|,\end{split} (4.38)

which in turn implies that

0aA(τ^a(f~a)σa(g~a))(g~af~a)aA(τ^a(g~a)σa(f~a))(g~af~a)2|A|T(d)||ΓΓ||.0\leq\sum_{a\in A}(\hat{\tau}_{a}(\tilde{f}_{a})-\sigma_{a}(\tilde{g}_{a}))\cdot(\tilde{g}_{a}-\tilde{f}_{a})\leq\sum_{a\in A}(\hat{\tau}_{a}(\tilde{g}_{a})-\sigma_{a}(\tilde{f}_{a}))\cdot(\tilde{g}_{a}-\tilde{f}_{a})\leq 2\cdot|A|\cdot T(d^{\prime})\cdot|\!|\Gamma^{\prime}-\Gamma^{\prime\prime}|\!|. (4.39)

Inequalities (4.39) and (4.36) imply that

0mΓaA|g~af~a|2aA(τ^a(g~a)τ^a(f~a))(g~af~a)aA(τ^a(g~a)σa(g~a))(g~af~a)||ΓΓ||aA|g~af~a|(1+|𝒦|MΓ)||ΓΓ||aA|g~af~a|(1+|𝒦|MΓ)||ΓΓ||aA|g~af~a|2.\begin{split}0\leq m_{\Gamma}\cdot\sum_{a\in A}|\tilde{g}_{a}-\tilde{f}_{a}|^{2}&\leq\sum_{a\in A}\big{(}\hat{\tau}_{a}(\tilde{g}_{a})-\hat{\tau}_{a}(\tilde{f}_{a})\big{)}\cdot\big{(}\tilde{g}_{a}-\tilde{f}_{a}\big{)}\\ &\leq\sum_{a\in A}(\hat{\tau}_{a}(\tilde{g}_{a})-\sigma_{a}(\tilde{g}_{a}))\cdot(\tilde{g}_{a}-\tilde{f}_{a})\leq|\!|\Gamma^{\prime}-\Gamma^{\prime\prime}|\!|\cdot\sum_{a\in A}|\tilde{g}_{a}-\tilde{f}_{a}|\\ &\leq(1+|\mathcal{K}|\cdot M_{\Gamma})\cdot|\!|\Gamma^{\prime}-\Gamma|\!|\cdot\sum_{a\in A}|\tilde{g}_{a}-\tilde{f}_{a}|\\ &\leq(1+|\mathcal{K}|\cdot M_{\Gamma})\cdot|\!|\Gamma^{\prime}-\Gamma|\!|\cdot\sqrt{\sum_{a\in A}|\tilde{g}_{a}-\tilde{f}_{a}|^{2}}.\end{split} (4.40)

Inequality (4.40) yields immediately that

|τ^a(f~a)τ^a(g~a)|MΓ(1+|𝒦|MΓ)mΓ||ΓΓ||aA.|\hat{\tau}_{a}(\tilde{f}_{a})-\hat{\tau}_{a}(\tilde{g}_{a})|\leq\frac{M_{\Gamma}\cdot(1+|\mathcal{K}|\cdot M_{\Gamma})}{m_{\Gamma}}\cdot|\!|\Gamma^{\prime}-\Gamma|\!|\quad\forall a\in A. (4.41)

We then obtain by (4.36), (4.38) and (4.41) that

|𝒞~(Γ)𝒞~(Γ)|=|aA(σa(g~a)g~aτ^a(f~a)f~a)|=|aA(σa(g~a)τ^a(g~a))g~a+aA(τ^a(g~a)g~aτ^a(f~a)f~a)|=aA|σa(g~a)τ^a(g~a)|g~a+aA|τ^a(g~a)τ^a(f~a)|g~a+aAτ^a(f~a)(g~af~a)2|A|T(d)||ΓΓ||+|A|T(d)MΓ(1+|𝒦|MΓ)mΓ||ΓΓ||<2(2+MΓmΓ)|A|T(d)(1+|𝒦|MΓ)||ΓΓ||\begin{split}\big{|}\tilde{\mathcal{C}}(\Gamma^{\prime})-&\tilde{\mathcal{C}}(\Gamma^{\prime\prime})\big{|}=\big{|}\sum_{a\in A}\big{(}\sigma_{a}(\tilde{g}_{a})\cdot\tilde{g}_{a}-\hat{\tau}_{a}(\tilde{f}_{a})\cdot\tilde{f}_{a}\big{)}\big{|}\\ &=\big{|}\sum_{a\in A}\big{(}\sigma_{a}(\tilde{g}_{a})-\hat{\tau}_{a}(\tilde{g}_{a})\big{)}\cdot\tilde{g}_{a}+\sum_{a\in A}\big{(}\hat{\tau}_{a}(\tilde{g}_{a})\cdot\tilde{g}_{a}-\hat{\tau}_{a}(\tilde{f}_{a})\cdot\tilde{f}_{a}\big{)}\big{|}\\ &=\sum_{a\in A}\big{|}\sigma_{a}(\tilde{g}_{a})-\hat{\tau}_{a}(\tilde{g}_{a})\big{|}\cdot\tilde{g}_{a}+\sum_{a\in A}\big{|}\hat{\tau}_{a}(\tilde{g}_{a})-\hat{\tau}_{a}(\tilde{f}_{a})\big{|}\cdot\tilde{g}_{a}+\sum_{a\in A}\hat{\tau}_{a}(\tilde{f}_{a})\cdot\big{(}\tilde{g}_{a}-\tilde{f}_{a}\big{)}\\ &\leq 2\cdot|A|\cdot T(d^{\prime})\cdot|\!|\Gamma^{\prime}-\Gamma^{\prime\prime}|\!|+|A|\cdot T(d^{\prime})\cdot\frac{M_{\Gamma}\cdot(1+|\mathcal{K}|\cdot M_{\Gamma})}{m_{\Gamma}}\cdot|\!|\Gamma^{\prime}-\Gamma|\!|\\ &<2\cdot\left(2+\frac{M_{\Gamma}}{m_{\Gamma}}\right)\cdot|A|\cdot T(d)\cdot(1+|\mathcal{K}|\cdot M_{\Gamma})\cdot|\!|\Gamma^{\prime}-\Gamma|\!|\end{split} (4.42)

when ||ΓΓ|||\!|\Gamma-\Gamma^{\prime}|\!| is small.

Inequalities (4.37) and (4.42) together then give

|ρ(Γ)ρ(Γ)|<4+4(2+MΓmΓ)ρ(Γ)𝒞(Γ)|A|T(d)(1+|𝒦|MΓ)||ΓΓ|||\rho(\Gamma^{\prime})-\rho(\Gamma^{\prime\prime})|<\frac{4+4\cdot\big{(}2+\frac{M_{\Gamma}}{m_{\Gamma}}\big{)}\cdot\rho(\Gamma)}{\mathcal{C}^{*}(\Gamma)}\cdot|A|\cdot T(d)\cdot(1+|\mathcal{K}|\cdot M_{\Gamma})\cdot|\!|\Gamma^{\prime}-\Gamma|\!| (4.43)

when ||ΓΓ|||\!|\Gamma-\Gamma^{\prime}|\!| is small.

By an argument similar with that for Lemma 11, we obtain that

|𝒞~(Γ)𝒞~(Γ)|2(2+MΓmΓ)|A|T(d)(1+|𝒦|MΓ)||ΓΓ||+2|A|τmax(T(d))|𝒦|(1+|𝒦|MΓ)||ΓΓ||,\begin{split}|\tilde{\mathcal{C}}(\Gamma)-\tilde{\mathcal{C}}(\Gamma^{\prime\prime})|&\leq 2\cdot\left(2+\frac{M_{\Gamma}}{m_{\Gamma}}\right)\cdot|A|\cdot T(d)\cdot(1+|\mathcal{K}|\cdot M_{\Gamma})\cdot|\!|\Gamma^{\prime}-\Gamma|\!|\\ &\hskip 56.9055pt+2\cdot|A|\cdot\tau_{max}(T(d))\cdot|\mathcal{K}|\cdot(1+|\mathcal{K}|\cdot M_{\Gamma})\cdot|\!|\Gamma^{\prime}-\Gamma|\!|,\end{split} (4.44)

when ||ΓΓ|||\!|\Gamma-\Gamma^{\prime}|\!| is small, where τmax(T(d)):=maxaτa(T(d)).\tau_{\max}(T(d)):=\max_{a}\tau_{a}(T(d)). Here, we use (4.42) instead of (4.15). Moreover, Lemma 11b) yields that

|𝒞(Γ)𝒞(Γ)|<4(|A||𝒦|τmax(T(d))+|A|T(d)MΓ1)(1+|𝒦|MΓ)||ΓΓ|||\mathcal{C}^{*}(\Gamma)-\mathcal{C}^{*}(\Gamma^{\prime\prime})|<4\cdot\big{(}|A|\cdot|\mathcal{K}|\cdot\tau_{max}(T(d))+|A|\cdot T(d)\cdot M_{\Gamma_{1}}\big{)}\cdot(1+|\mathcal{K}|\cdot M_{\Gamma})\cdot|\!|\Gamma^{\prime}-\Gamma|\!| (4.45)

when ||ΓΓ|||\!|\Gamma-\Gamma^{\prime}|\!| is small. Inequalities (4.44) and (4.45) imply that

|ρ(Γ)ρ(Γ)|<νΓ(MΓ,mΓ,T(d),τmax(T(d)))||ΓΓ|||\rho(\Gamma)-\rho(\Gamma^{\prime\prime})|<\nu_{\Gamma}(M_{\Gamma},m_{\Gamma},T(d),\tau_{\max}(T(d)))\cdot|\!|\Gamma-\Gamma^{\prime}|\!| (4.46)

when ||ΓΓ|||\!|\Gamma-\Gamma^{\prime}|\!| is small. Here νΓ(MΓ,mΓ,T(d),τmax(T(d)))>0\nu_{\Gamma}(M_{\Gamma},m_{\Gamma},T(d),\tau_{\max}(T(d)))>0 is a constant depending only on MΓ,M_{\Gamma}, mΓ,m_{\Gamma}, T(d),T(d), and τmax(T(d))\tau_{\max}(T(d)) of Γ.\Gamma.

Inequalities (4.43) and (4.46) together imply Theorem 12b).

This completes the proof of Theorem 12.

\square

We have been able to obtain a stronger Hölder exponent in Theorem 12a)–b) since we no longer need Lemma 1c) in the Hölder continuity analysis of the PoA on demand slices when the cost functions are constants or have strictly positive first-order derivatives. For constant cost functions, we have used inequality (4.32) instead of Lemma 1c). For cost functions with strictly positive first-order derivatives, we have used inequalities (4.39)–(4.42) instead. Unfortunately, these inequalities do not hold for arbitrary Lipschitz continuous cost functions on [0,T(d)].[0,T(d)]. So far, we are still lack a unified technique to replace Lemma 1c) in the Hölder analysis of the PoA on demand slices with arbitrary Lipschtiz continuous cost functions on [0,T(d)].[0,T(d)].

4.4 Open questions

We have shown for Lipschitz continuous cost functions on [0,T(w)][0,T(w)] that the PoA is pointwise Hölder continuous with Hölder exponent 12\frac{1}{2} on a demand slice 𝒢(G,𝒦,𝒮)|w,\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|w}, see Lemma 10d). At present, we are unable to improve this exponent, since we need Lemma 1c) to bound |C(Γ1,f~)C(Γ2,g~)||C(\Gamma_{1},\tilde{f})-C(\Gamma_{2},\tilde{g})| for two games Γ1=(π(1),w)\Gamma_{1}=(\pi^{(1)},w) and Γ2=(π(2),w)\Gamma_{2}=(\pi^{(2)},w) from the same demand slice 𝒢(G,𝒦,𝒮)|w,\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|w}, see (4.9), and since both Lemma 1c) and the upper bound on the approximation threshold in Lemma 10c) are tight, see Example 2 and Remark 4.2. Hence, it is unclear to us if the exponent 12\frac{1}{2} is tight for arbitrary Lipschitz continuous cost functions on [0,T(w)].[0,T(w)]. We leave this as an open question.

Open Question 4.1.

Is the Hölder exponent 12\frac{1}{2} in Lemma 10d) tight for Lipschitz continuous cost functions on [0,T(w)][0,T(w)]?

Although we can neither affirm nor negate this question, there is some evidence for a negative answer. Both Example 2 and Remark 4.2 have only used constant cost functions or polynomial cost functions with strictly positive first-order derivatives, for which we have shown an improved Hölder exponent in Theorem 12. Hence, cost functions making Lemma 1c) tight need not make the approximation threshold upper bound in Lemma 10c) tight, and vice versa. Nonetheless, it will be challenging to show an improved Hölder exponent for arbitrary Lipschitz continuous cost functions on [0,T(w)].[0,T(w)]. This will require an alternative approach to bound |C(Γ1,f~)C(Γ2,g~)||C(\Gamma_{1},\tilde{f})-C(\Gamma_{2},\tilde{g})|, even different from our special approaches for cost functions that are constants or have strictly positive first-order derivatives in Section 4.3.

Theorem 12 implies directly that the upper derivative

D¯ρ(Γ):=lim¯ΓΓ|ρ(Γ)ρ(Γ)|||ΓΓ||\overline{D}_{\rho}(\Gamma):=\overline{\lim}_{\Gamma^{\prime}\to\Gamma}\frac{|\rho(\Gamma^{\prime})-\rho(\Gamma)|}{|\!|\Gamma^{\prime}-\Gamma|\!|} (4.47)

of the PoA map ρ()\rho(\cdot) at a game Γ=(τ,d)\Gamma=(\tau,d) exists and is bounded from above by a finite Hölder constant ϰΓ>0\varkappa_{\Gamma}>0 when the cost functions τa()\tau_{a}(\cdot) have strictly positive derivatives on the compact interval [0,T(d)].[0,T(d)]. While Cominetti et al. [2020] obtained a stronger differentiability result of the resulting PoA function on a cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} for networks with one O/D pair, we are unable to generalize their result to the whole game space 𝒢(G,𝒦,𝒮)\mathcal{G}(G,\mathcal{K},\mathcal{S}) on networks with multiple O/D pairs, as our PoA map ρ()\rho(\cdot) is not an ordinary real-valued function, but a functional on the game space 𝒢(G,𝒦,𝒮)\mathcal{G}(G,\mathcal{K},\mathcal{S}). In particular, it is also unclear to us under which conditions the upper derivative D¯ρ(Γ)\overline{D}_{\rho}(\Gamma) coincides with the lower derivative

D¯ρ(Γ):=lim¯ΓΓ|ρ(Γ)ρ(Γ)|||ΓΓ||,\underline{D}_{\rho}(\Gamma):=\underline{\lim}_{\Gamma^{\prime}\to\Gamma}\frac{|\rho(\Gamma^{\prime})-\rho(\Gamma)|}{|\!|\Gamma^{\prime}-\Gamma|\!|}, (4.48)

although it is clear that 0D¯ρ(Γ)D¯ρ(Γ).0\leq\underline{D}_{\rho}(\Gamma)\leq\overline{D}_{\rho}(\Gamma). We leave this also as an open question.

Open Question 4.2.

Which condition on Γ=(τ,d)\Gamma=(\tau,d) implies that D¯ρ(Γ)=D¯ρ(Γ)\underline{D}_{\rho}(\Gamma)=\overline{D}_{\rho}(\Gamma)?

Addressing Open Question 4.2 on the game space 𝒢(G,𝒦,𝒮)\mathcal{G}(G,\mathcal{K},\mathcal{S}) might be too ambitious due to the extremely complicated structure of its topology. Inspired by Patriksson [2004], Josefsson and Patriksson [2007], Lu [2008], and Klimm and Warode [2021], a promising first step is to consider the differentiability of the PoA on a particular subspace of 𝒢(G,𝒦,𝒮)\mathcal{G}(G,\mathcal{K},\mathcal{S}) that can be parameterized and is homomorphic to some finite dimensional Euclidean space. Then classic Calculus techniques for finite dimensional Euclidean spaces apply. This may facilitate the differentiability analysis of the PoA. Nevertheless, we will not continue this direction in the current paper, and would like to leave it for future work.

5 An application to the convergence rate of the PoA

As an application of our Hölder continuity results, we now demonstrate that they help to analyze the convergence rate of the PoA in non-atomic congestion games for both growing and decreasing demands.

The convergence analysis of the PoA investigates the limit of the PoA sequence (ρ(Γn))n\big{(}\rho(\Gamma_{n})\big{)}_{n\in\mathbb{N}} of a sequence (Γn)n(\Gamma_{n})_{n\in\mathbb{N}} of games when all components Γn=(τ,d(n))\Gamma_{n}=(\tau,d^{(n)}) belong to the same cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} and the total demand T(d(n))T(d^{(n)}) tends to zero or infinity, i.e., limnT(d(n))=0\lim_{n\to\infty}T(d^{(n)})=0 or .\infty. When limnρ(Γn)=1\lim_{n\to\infty}\rho(\Gamma_{n})=1 for an arbitrary sequence (Γn)n𝒢(G,𝒦,𝒮)|τ(\Gamma_{n})_{n\in\mathbb{N}}\in\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau}^{\mathbb{N}} with limnT(d(n))=,\lim_{n\to\infty}T(d^{(n)})=\infty, then we say that the cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} behaves well for growing demands. Notice that this notion has been introduced by Colini-Baldeschi et al. [2016] and corresponds to the notion of asymptotically well designed introduced by Wu et al. [2021].

Similarly, when limnρ(Γn)=1\lim_{n\to\infty}\rho(\Gamma_{n})=1 for an arbitrary sequence (Γn)n𝒢(G,𝒦,𝒮)|τ(\Gamma_{n})_{n\in\mathbb{N}}\in\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau}^{\mathbb{N}} with limnT(d(n))=0,\lim_{n\to\infty}T(d^{(n)})=0, then we say that the cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} behaves well for decreasing demands. In particular, the cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} is said to behave well in limits when it behaves well for both decreasing and growing demands. When a cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} behaves well in limits, then the PoA map ρ()\rho(\cdot) has a tight and finite upper bound on 𝒢(G,𝒦,𝒮)|τ,\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau}, although the cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} is not compact w.r.t. the topology from Section 3. This follows since the PoA is continuous and thus has a tight and finite upper bound on each compact subspace {Γ=(τ,d)𝒢(G,𝒦,𝒮)|τ:N1||d||N2}\{\Gamma^{\prime}=(\tau,d)\in\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau}:N_{1}\leq|\!|d|\!|_{\infty}\leq N_{2}\} of the cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} for two arbitrary positive reals N1N_{1} and N2N_{2} with N1<N2.N_{1}<N_{2}.

5.1 The state of the art

Recent results by Colini-Baldeschi et al. [2017, 2020] and Wu et al. [2021] have already shown that every cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} with regularly varying (Bingham et al. [1987]) cost functions behaves well for growing demands. This applies directly to cost slices 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} with cost functions τa()\tau_{a}(\cdot) that are arbitrary polynomials, arbitrary logarithms, or products of polynomials and logarithms, and thus confirms the earlier observed convergence of the empirical PoA for growing demands by Youn et al. [2008]; Harks et al. [2015]; O’Hare et al. [2016] and Monnot et al. [2017].

When the cost functions τa()\tau_{a}(\cdot) are of the form τa(x)=nξa,nxn\tau_{a}(x)=\sum_{n\in\mathbb{N}}\xi_{a,n}\cdot x^{n} for each aAa\in A and each x[0,)x\in[0,\infty), then Colini-Baldeschi et al. [2017, 2020] have shown that limnρ(Γn)=1\lim_{n\to\infty}\rho(\Gamma_{n})=1 for each sequence (Γn)n𝒢(G,𝒦,𝒮)|τ(\Gamma_{n})_{n\in\mathbb{N}}\in\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau}^{\mathbb{N}} with a demand sequence (d(n))n(d^{(n)})_{n\in\mathbb{N}} satisfying limnT(d(n))=0\lim_{n\to\infty}T(d^{(n)})=0 and the condition that

lim¯ndk(n)T(d(n))>0k𝒦.\varliminf_{n\to\infty}\frac{d_{k}^{(n)}}{T(d^{(n)})}>0\quad\forall k\in\mathcal{K}. (5.1)

This is presently the only known convergence result for decreasing demands.

For polynomial cost functions τa()\tau_{a}(\cdot) and a sequence (Γn)n(\Gamma_{n})_{n\in\mathbb{N}} of games Γn=(τ=(τa)aA,d(n))\Gamma_{n}=(\tau=(\tau_{a})_{a\in A},d^{(n)}) satisfying the condition that

dk(n)T(d(n))=:dk>0k𝒦n,\frac{d_{k}^{(n)}}{T(d^{(n)})}=:d_{k}>0\quad\forall k\in\mathcal{K}\ \forall n\in\mathbb{N}, (5.2)

Colini-Baldeschi et al. [2017, 2020] have shown further that ρ(Γn)=1+O(1T(d(n)))\rho(\Gamma_{n})=1+O(\frac{1}{T(d^{(n)})}) when limnT(d(n))=,\lim_{n\to\infty}T(d^{(n)})=\infty, and that ρ(Γn)=1+O(T(d(n)))\rho(\Gamma_{n})=1+O(T(d^{(n)})) when limnT(d(n))=0.\lim_{n\to\infty}T(d^{(n)})=0. Here, the rate dkd_{k} in (5.2) is a constant independent of nn for each O/D pair k𝒦.k\in\mathcal{K}.

With a different technique, Wu et al. [2021] have shown that condition (5.2) can be removed when the cost functions possess certain good properties. In particular, they proved that ρ(Γ)=1+o(1T(d)β)\rho(\Gamma)=1+o(\frac{1}{T(d)^{\beta}}) for each game Γ=(τ,d)𝒢(G,𝒦,𝒮)|τ\Gamma=(\tau,d)\in\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} when the total demand T(d)T(d) of the game Γ\Gamma is large and the cost functions τa()\tau_{a}(\cdot) are of the BPR type (Bureau of Public Roads [1964]) and have the same degree β>0.\beta>0. Moreover, Wu et al. [2021] have illustrated on an example game Γ=(τ,d)\Gamma=(\tau,d) with BPR cost functions that the rate at which ρ(Γ)\rho(\Gamma) converges to 11 depends crucially on the growth pattern of the demands. For each constant exponent θ<β+1,2β>,\theta\in<\beta+1,2\cdot\beta>, there is a game sequence (Γn)n𝒢(G,𝒦,𝒮)|τ(\Gamma_{n})_{n\in\mathbb{N}}\in\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau}^{\mathbb{N}} such that limnT(d(n))=\lim_{n\to\infty}T(d^{(n)})=\infty and ρ(Γn)=1+Θ(1T(d(n))θ)\rho(\Gamma_{n})=1+\Theta(\frac{1}{T(d^{(n)})^{\theta}}) for large n.n. Here, <v1,v2><v_{1},v_{2}> denotes the closed interval [min{v1,v2},max{v1,v2}][\min\{v_{1},v_{2}\},\max\{v_{1},v_{2}\}] for arbitrary two real numbers v1v_{1} and v2.v_{2}. This negates a conjecture proposed by O’Hare et al. [2016].

5.2 Convergence rates of the PoA for decreasing demands

Consider now a vector τ=(τa)aA\tau=(\tau_{a})_{a\in A} of cost functions τa()\tau_{a}(\cdot) that are defined on the unbounded interval [0,),[0,\infty), strictly positive and Lipschitz continuous on a compact interval [0,b][0,b] with Lipschitz constant Mτ>0M_{\tau}>0 for a constant b(0,).b\in(0,\infty). We now show with Theorem 12a) that the cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} of τ\tau behaves well for decreasing demands.

Let σa(x)τa(0)>0\sigma_{a}(x)\equiv\tau_{a}(0)>0 for each aAa\in A and x[0,).x\in[0,\infty). For each game Γ=(τ,d)𝒢(G,𝒦,𝒮)|τ,\Gamma=(\tau,d)\in\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau}, we define an auxiliary game Γ:=(σ,dT(d))𝒢(G,𝒦,𝒮).\Gamma^{\prime}:=(\sigma,\frac{d}{T(d)})\in\mathcal{G}(G,\mathcal{K},\mathcal{S}). Clearly, ρ(Γ)1\rho(\Gamma^{\prime})\equiv 1 for an arbitrary demand vector d,d, since the cost functions σa()\sigma_{a}(\cdot) are constant. By (3.3), we obtain that

||ΓΛT(d)(Γ)||=maxaA,x[0,1]|τa(T(d)x)σa(x)|=maxaA|τa(T(d))τa(0)|MτT(d)|\!|\Gamma^{\prime}-\Lambda_{T(d)}(\Gamma)|\!|=\max_{a\in A,x\in[0,1]}|\tau_{a}(T(d)\cdot x)-\sigma_{a}(x)|=\max_{a\in A}\ |\tau_{a}(T(d))-\tau_{a}(0)|\leq M_{\tau}\cdot T(d)

when T(d)b.T(d)\leq b. Here, we use that ΛT(d)(Γ)𝒢(G,𝒦,𝒮)\Lambda_{T(d)}(\Gamma)\in\mathcal{G}(G,\mathcal{K},\mathcal{S}) is the game resulting from the demand normalization to Γ\Gamma with factor T(d),T(d), and that both ΛT(d)(Γ)\Lambda_{T(d)}(\Gamma) and Γ\Gamma^{\prime} are from the same demand slice 𝒢(G,𝒦,𝒮)|dT(d)\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\frac{d}{T(d)}} and so have the same total demand T(dT(d))=1.T(\frac{d}{T(d)})=1.

Let τmin(0):=minaAτa(0)>0.\tau_{\min}(0):=\min_{a\in A}\tau_{a}(0)>0. Then 𝒞(Γ)τmin(0)\mathcal{C}^{*}(\Gamma^{\prime})\geq\tau_{\min}(0) for every demand vector dd of Γ,\Gamma, since Γ\Gamma^{\prime} has the total demand T(dT(d))=1.T(\frac{d}{T(d)})=1. Inequality (4.33) then applies and yields

|ρ(Γ)ρ(Γ)|=|ρ(ΛT(d)(Γ))ρ(Γ)|=ρ(ΛT(d)(Γ))18|A|(|𝒦|+1)τmin(0)MτT(d)\begin{split}|\rho(\Gamma^{\prime})-\rho(\Gamma)|=|\rho(\Lambda_{T(d)}(\Gamma))-\rho(\Gamma^{\prime})|=\rho(\Lambda_{T(d)}(\Gamma))-1\leq\frac{8\cdot|A|\cdot(|\mathcal{K}|+1)}{\tau_{\min}(0)}\cdot M_{\tau}\cdot T(d)\end{split}

when T(d)T(d) is small, which tends to 0 as T(d)0.T(d)\to 0. Here, we use that both the lower bound τmin(0)\tau_{\min}(0) of the SO cost 𝒞(Γ)\mathcal{C}^{*}(\Gamma^{\prime}) and the total demand T(dT(d))T(\frac{d}{T(d)}) (1\equiv 1) of the game Γ\Gamma^{\prime} do not depend on the demand vector dd of Γ\Gamma though the game Γ\Gamma^{\prime} does, and that the pointwise Hölder constant in (4.33) depends only on the SO cost and the total demand, and is thus bounded from above by the constant 8|A|(|𝒦|+1)τmin(0)\frac{8\cdot|A|\cdot(|\mathcal{K}|+1)}{\tau_{\min}(0)} independent of the demand vector dd of Γ\Gamma.

Hence, 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} behaves well for decreasing demands. We summarize in Corollary 5.1.

Corollary 5.1.

Consider an arbitrary vector τ=(τa)aA\tau=(\tau_{a})_{a\in A} of cost functions τa()\tau_{a}(\cdot) that are defined on the unbounded interval [0,),[0,\infty), strictly positive and Lipschitz continuous on a compact interval [0,b][0,b] with Lipschitz constant Mτ>0M_{\tau}>0 for a constant b(0,).b\in(0,\infty). Then ρ(Γ)1+8Mτ|A|(|𝒦|+1)τmin(0)MτT(d)\rho(\Gamma)\leq 1+\frac{8\cdot M_{\tau}\cdot|A|\cdot(|\mathcal{K}|+1)}{\tau_{\min}(0)}\cdot M_{\tau}\cdot T(d) for every game Γ=(τ,d)𝒢(G,𝒦,𝒮)|τ\Gamma=(\tau,d)\in\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} when the total demand T(d)T(d) is small and τmin(0):=minaAτa(0)>0.\tau_{\min}(0):=\min_{a\in A}\tau_{a}(0)>0. In particular, the cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} behaves well for decreasing demands.

Corollary 5.1 and the convergence results in Wu et al. [2021] imply immediately for every slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} whose cost functions τa()\tau_{a}(\cdot) are strictly positive, differentiable and regularly varying that the PoA behaves well in limits. Theorem 6 then implies that the PoA has a finite and tight upper bound on each of these cost slices 𝒢(G,𝒦,𝒮)|τ.\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau}. In particular, Corollary 5.1 and Wu et al. [2021] together confirm the observed behavior of the empirical PoA for Beijing traffic data, which starts at 11, then increases at a relatively mild rate with growing total demand, and eventually decays rapidly to 11 after the total demand reaches a certain threshold, see Wu et al. [2021].

5.3 Convergence rates of the PoA for growing demands

Colini-Baldeschi et al. [2017, 2020] have demonstrated that the PoA may diverge for growing total demand when the cost functions are not regularly varying. Wu et al. [2021] showed the convergence of the PoA to 11 for growing total demand for arbitrary regularly varying cost functions, but concrete convergence rates of the PoA for arbitrary regularly varying cost functions (other than polynomials) are still missing. Our Hölder continuity results prove to be very helpful here and yield the first convergence rate for growing total demand for certain regularly varying cost functions that are not polynomials.

Consider a vector τ=(τa)aA\tau=(\tau_{a})_{a\in A} of regularly varying cost functions τa()\tau_{a}(\cdot) that have the same regular variation index β>0\beta>0 and satisfy the condition

limTτa(T)τb(T)=:λa,b(0,)a,bA.\begin{split}\lim_{T\to\infty}\frac{\tau_{a}(T)}{\tau_{b}(T)}=:\lambda_{a,b}\in(0,\infty)\quad\forall a,b\in A.\end{split} (5.3)

Karamata’s Characterization Theorem (Bingham et al. [1987]) yields that these cost functions τa()\tau_{a}(\cdot) have the form τa(x)=Qa(x)xβ\tau_{a}(x)=Q_{a}(x)\cdot x^{\beta} for a (slowly varying) function Qa(x)Q_{a}(x) satisfying the condition that limTQa(Tx)Qa(T)=1\lim_{T\to\infty}\frac{Q_{a}(T\cdot x)}{Q_{a}(T)}=1 for all x>0.x>0. Condition (5.3) then implies that limTQ(T)Qb(T)=λa,b\lim_{T\to\infty}\frac{Q(T)}{Q_{b}(T)}=\lambda_{a,b} for all a,bA.a,b\in A.

Consider now an arbitrary game Γ=(τ,d)\Gamma=(\tau,d) from the cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} of τ,\tau, and an arbitrary arc bA.b\in A. Then ρ(Γ)=ρ(Γ^)\rho(\Gamma)=\rho(\hat{\Gamma}) when Γ^:=(τ^,dT(d))\hat{\Gamma}:=(\hat{\tau},\frac{d}{T(d)}) with τ^a(x):=τa(T(d)x)τb(T(d))\hat{\tau}_{a}(x):=\frac{\tau_{a}(T(d)\cdot x)}{\tau_{b}(T(d))} for all aAa\in A and all x[0,1].x\in[0,1]. This holds since Γ^=Ψτb(T(d))ΛT(d)(Γ).\hat{\Gamma}=\Psi_{\tau_{b}(T(d))}\circ\Lambda_{T(d)}(\Gamma).

Consider the auxiliary cost function vector σ=(σa)aA\sigma=(\sigma_{a})_{a\in A} with

σa(x):=λa,bxβ(a,x)A×[0,1],\sigma_{a}(x):=\lambda_{a,b}\cdot x^{\beta}\quad\forall(a,x)\in A\times[0,1],

and put Γ:=(σ,dT(d)).\Gamma^{\prime}:=(\sigma,\frac{d}{T(d)}). Then ρ(Γ)1,\rho(\Gamma^{\prime})\equiv 1, as the cost functions of Γ\Gamma^{\prime} are monomials of the same degree β\beta, see, e.g., Wu et al. [2021] and Roughgarden and Tardos [2002].

Both Γ\Gamma^{\prime} and Γ^\hat{\Gamma} belong to the same cost slice 𝒢(G,𝒦,𝒮)|dT(d)\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\frac{d}{T(d)}} and have the same total demand T(dT(d))1.T(\frac{d}{T(d)})\equiv 1. Moreover, 𝒞(Γ)minaAλa,b>0\mathcal{C}^{*}(\Gamma^{\prime})\geq\min_{a\in A}\lambda_{a,b}>0 and σa(x)\sigma_{a^{\prime}}(x) is Lipschitz continuous on the compact interval [0,1][0,1] with Lipschitz constant βmaxaAλa,b\beta\cdot\max_{a\in A}\lambda_{a,b} for all aA.a^{\prime}\in A.

Lemma 10d) then applies and yields that

|ρ(Γ)ρ(Γ)|=|ρ(Γ^)ρ(Γ)|2ρ(Γ)+|A|T(dT(d))βmaxaAλa,b+2𝒞(Γ)|A|T(dT(d))||Γ^Γ||21+|A|βmaxaAλa,b+2minaAλa,b|A|||Γ^Γ||\begin{split}|\rho(\Gamma)-&\rho(\Gamma^{\prime})|=|\rho(\hat{\Gamma})-\rho(\Gamma^{\prime})|\\ &\leq 2\cdot\frac{\rho(\Gamma^{\prime})+\sqrt{|A|\cdot T(\frac{d}{T(d)})\cdot\beta\cdot\max_{a\in A}\lambda_{a,b}}+2}{\mathcal{C}^{*}(\Gamma^{\prime})}\cdot|A|\cdot T(\frac{d}{T(d)})\cdot\sqrt{|\!|\hat{\Gamma}-\Gamma^{\prime}|\!|}\\ &\leq 2\cdot\frac{1+\sqrt{|A|\cdot\beta\cdot\max_{a\in A}\lambda_{a,b}}+2}{\min_{a\in A}\lambda_{a,b}}\cdot|A|\cdot\sqrt{|\!|\hat{\Gamma}-\Gamma^{\prime}|\!|}\end{split} (5.4)

when ||Γ^Γ||𝒞(Γ)2|A|.|\!|\hat{\Gamma}-\Gamma^{\prime}|\!|\leq\frac{\mathcal{C}^{*}(\Gamma^{\prime})}{2\cdot|A|}. Here, we have used that

||Γ^Γ||\displaystyle|\!|\hat{\Gamma}-\Gamma^{\prime}|\!| =||τ^|1σ|1||=maxaA,x[0,1]|τ^a(x)σa(x)|\displaystyle=|\!|\hat{\tau}_{|1}-\sigma_{|1}|\!|_{\infty}=\max_{a\in A,x\in[0,1]}|\hat{\tau}_{a}(x)-\sigma_{a}(x)|
=maxaA,x[0,1]|Qa(T(d)x)Qb(T(d))xβλa,bxβ|=:w(T(d))\displaystyle=\max_{a\in A,x\in[0,1]}\big{|}\frac{Q_{a}(T(d)\cdot x)}{Q_{b}(T(d))}\cdot x^{\beta}-\lambda_{a,b}\cdot x^{\beta}\big{|}=:w(T(d)) (5.5)
=maxaA,x[0,1]|Qa(T(d))Qb(T(d))Qa(T(d)x)Qa(T(d))λa,b|xβ\displaystyle=\max_{a\in A,x\in[0,1]}\big{|}\frac{Q_{a}(T(d))}{Q_{b}(T(d))}\cdot\frac{Q_{a}(T(d)\cdot x)}{Q_{a}(T(d))}-\lambda_{a,b}\big{|}\cdot x^{\beta}
maxaA,x[0,1]|Qa(T(d))Qb(T(d))λa,b|Qa(T(d)x)Qa(T(d))xβ+maxaA,x[0,1]|Qa(T(d)x)Qa(T(d))1|λa,bxβ,\displaystyle\leq\max_{a\in A,x\in[0,1]}\big{|}\frac{Q_{a}(T(d))}{Q_{b}(T(d))}-\lambda_{a,b}\big{|}\cdot\frac{Q_{a}(T(d)\cdot x)}{Q_{a}(T(d))}\cdot x^{\beta}+\max_{a\in A,x\in[0,1]}\big{|}\frac{Q_{a}(T(d)\cdot x)}{Q_{a}(T(d))}-1\big{|}\cdot\lambda_{a,b}\cdot x^{\beta},

which tends to 0 as T(d).T(d)\to\infty. Hence, ρ(Γ)1\rho(\Gamma)\to 1 as T(d),T(d)\to\infty, since ρ(Γ)1.\rho(\Gamma^{\prime})\equiv 1.

Inequalities (5.4) and (5.5) show that the convergence rate of ρ(Γ)\rho(\Gamma) depends heavily on the properties of the functions Qa(x)Q_{a}(x). As an example, we assume now that these factors Qa(x)Q_{a}(x) are of the form

ζalnα(x+1),ζa>0,α0.\zeta_{a}\cdot\ln^{\alpha}(x+1),\quad\zeta_{a}>0,\ \alpha\geq 0. (5.6)

Then we obtain by (5.4) that

||ΓΓ^||=w(T(d))=maxaA,x[0,1]|Qa(T(d)x)Qb(T(d))xβλa,bxβ|=maxaA,x[0,1]λa,b(xβ(ln(T(d)x+1)ln(T(d)+1))αxβ)maxaA,x[0,1]λa,bαβT(d)xβ+1T(d)x+1lnα1(T(d)x+1)lnα(T(d)+1)αβT(d)T(d)+11ln(T(d)+1)maxaAλa,b\begin{split}|\!|\Gamma^{\prime}-\hat{\Gamma}|\!|&=w(T(d))=\max_{a\in A,x\in[0,1]}\big{|}\frac{Q_{a}(T(d)\cdot x)}{Q_{b}(T(d))}\cdot x^{\beta}-\lambda_{a,b}\cdot x^{\beta}\big{|}\\ &=\max_{a\in A,x\in[0,1]}\lambda_{a,b}\cdot\Big{(}x^{\beta}-\big{(}\frac{\ln(T(d)\cdot x+1)}{\ln(T(d)+1)}\big{)}^{\alpha}\cdot x^{\beta}\Big{)}\\ &\leq\max_{a\in A,x\in[0,1]}\lambda_{a,b}\cdot\frac{\alpha}{\beta}\cdot\frac{T(d)\cdot x^{\beta+1}}{T(d)\cdot x+1}\cdot\frac{\ln^{\alpha-1}(T(d)\cdot x+1)}{\ln^{\alpha}(T(d)+1)}\\ &\leq\frac{\alpha}{\beta}\cdot\frac{T(d)}{T(d)+1}\cdot\frac{1}{\ln(T(d)+1)}\cdot\max_{a\in A}\ \lambda_{a,b}\end{split}

and thus ρ(Γ)=1+O(1ln(T(d)+1)).\rho(\Gamma)=1+O(\sqrt{\frac{1}{\ln(T(d)+1)}}). This follows since β>0\beta>0 and the function

xβ(ln(T(d)x+1)ln(T(d)+1))αxβ,x[0,1],x^{\beta}-\Big{(}\frac{\ln(T(d)\cdot x+1)}{\ln(T(d)+1)}\Big{)}^{\alpha}\cdot x^{\beta},\quad x\in[0,1],

reaches its maximum at a point x[0,1]x\in[0,1] satisfying the equation

lnα(T(d)+1)lnα(T(d)x+1)=αβT(d)xT(d)x+1lnα1(T(d)x+1).\ln^{\alpha}(T(d)+1)-\ln^{\alpha}(T(d)\cdot x+1)=\frac{\alpha}{\beta}\cdot\frac{T(d)\cdot x}{T(d)\cdot x+1}\cdot\ln^{\alpha-1}(T(d)\cdot x+1).

We summarize this in Corollary 5.2.

Corollary 5.2.

Consider an arbitrary cost function vector τ=(τa)aA\tau=(\tau_{a})_{a\in A} s.t. all cost functions τa()\tau_{a}(\cdot) are regularly varying with the same regular variation index β>0\beta>0 and satisfy the condition in equation (5.3).

  • a)

    The resulting cost slice 𝒢(G,𝒦,𝒮)|τ\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} behaves well for growing demands, and ρ(Γ)=1+O(w(T(d)))\rho(\Gamma)=1+O(\sqrt{w(T(d))}) with the upper bound w(T(d))w(T(d)) defined in (5.5) for each game Γ=(τ,d)𝒢(G,𝒦,𝒮)|τ\Gamma=(\tau,d)\in\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} with a large total demand T(d).T(d).

  • b)

    For cost functions τa(x)=ζaxβlnα(x+1),\tau_{a}(x)=\zeta_{a}\cdot x^{\beta}\cdot\ln^{\alpha}(x+1), aA,a\in A, the PoA is ρ(Γ)=1+O(1ln(T(d)+1))\rho(\Gamma)=1+O(\sqrt{\frac{1}{\ln(T(d)+1)}}) for each game Γ=(τ,d)𝒢(G,𝒦,𝒮)|τ\Gamma=(\tau,d)\in\mathcal{G}(G,\mathcal{K},\mathcal{S})_{|\tau} with a large total demand T(d).T(d). Here, α>0\alpha>0 and ζa>0\zeta_{a}>0 are arbitrary constants independent of the demand vector dd.

Condition (5.5) actually means that both the regularly varying cost functions τa()\tau_{a}(\cdot) and the slowly varying functions Qa()Q_{a}(\cdot) are mutually comparable, i.e., both limTτa(T)τb(T)\lim_{T\to\infty}\frac{\tau_{a}(T)}{\tau_{b}(T)} and limTQa(T)Qb(T)\lim_{T\to\infty}\frac{Q_{a}(T)}{Q_{b}(T)} exist for any two a,bAa,b\in A. These cost functions in Corollary 5.2 can then be thought of as a generalization of the BPR cost functions qaxβ+paq_{a}\cdot x^{\beta}+p_{a}, when we substitute the positive constant factors qaq_{a} by mutually comparable and slowing varying functions Qa()Q_{a}(\cdot). However, this generalization comes at the price of a weaker convergence rate of the PoA.

6 Summary and future work

This paper presents the first sensitivity analysis for the PoA in non-atomic congestion games when both the demands and cost functions may change. To achieve this, we have introduced a topology and a metric on the class of games with the same combinatorial structure, which may also be of use for other research purposes. The PoA, the SO cost, and the WE cost turned out to be continuous maps w.r.t. that topology, see Theorem 6. Their dependence on a small variation of the demands and/or Lipschitz continuous cost functions is thus small. With the metric, we have quantified the variation of the PoA when demands and cost functions change simultaneously.

This has led to an analysis of the Hölder continuity of the PoA map ρ()\rho(\cdot) on the game space. We have shown first that the map ρ()\rho(\cdot) is not uniformly Hölder continuous on the whole game space, and that the Hölder neighborhood of each game is a proper subset of the game space, see Theorem 8. So the PoA can in general only be pointwise and locally Hölder continuous. For each game Γ=(τ,d)\Gamma=(\tau,d) with cost functions τa()\tau_{a}(\cdot) that are Lipschitz continuous on the compact interval [0,T(d)],[0,T(d)], we have shown in Theorem 9 that the PoA map ρ()\rho(\cdot) is pointwise Hölder continuous at Γ\Gamma with Hölder exponent 12,\frac{1}{2}, i.e., there are constants ϰΓ,ϵΓ>0\varkappa_{\Gamma},\epsilon_{\Gamma}>0 depending only on Γ\Gamma such that |ρ(Γ)ρ(Γ)|<ϰΓ||ΓΓ|||\rho(\Gamma^{\prime})-\rho(\Gamma)|<\varkappa_{\Gamma}\cdot\sqrt{|\!|\Gamma^{\prime}-\Gamma|\!|} for each game Γ\Gamma^{\prime} with ||ΓΓ||<ϵΓ.|\!|\Gamma^{\prime}-\Gamma|\!|<\epsilon_{\Gamma}. When the cost functions τa()\tau_{a}(\cdot) have stronger properties, e.g., when they are constant or have strictly positive derivatives on the compact interval [0,T(d)][0,T(d)], we have shown in Theorem 12a)–b) that the PoA map ρ()\rho(\cdot) then has the stronger Hölder exponent 11 at the game Γ\Gamma, which is of the same order as in the recent results by Englert et al. [2010], Takalloo and Kwon [2020] and Cominetti et al. [2020] for demand changes.

Finally, we have applied our results to analyze the convergence behavior of the PoA when the total demand T(d)T(d) tends to 0 or .\infty. We showed that the PoA tends to 11 at a rate of O(T(d))O(T(d)) for decreasing total demand T(d)T(d) when the cost functions are strictly positive and Lipschitz continuous within a small neighborhood around the origin 0,0, and identified a class of non-polynomial regularly varying cost functions for which the PoA tends to 11 at a rate of O(1/ln(T(d)+1))O(\sqrt{1/\ln(T(d)+1)}) for growing total demand. These complement recent results on the convergence rate by Colini-Baldeschi et al. [2017, 2020] and Wu et al. [2021].

Theorem 9 yields the Hölder exponent 12\frac{1}{2} for the PoA at games Γ=(τ,d)\Gamma=(\tau,d) whose cost functions are Lipschitz continuous on the compact interval [0,T(d)].[0,T(d)]. We conjecture that this exponent should be 1,1, but could only confirm this in Theorem 12 for games Γ=(τ,d)\Gamma=(\tau,d) with cost functions τa()\tau_{a}(\cdot) that are constant or have strictly positive derivatives on [0,T(d)].[0,T(d)]. We have neither been able to confirm this, nor to provide a counterexample for games Γ=(τ,d)\Gamma=(\tau,d) with cost functions τa()\tau_{a}(\cdot) that are Lipschitz continuous on [0,T(d)].[0,T(d)].

This is closely related to Open Question 4.1, which concerns the tightness of the Hölder exponent 12\frac{1}{2} for Lemma 10d). We guess that the exponent 12\frac{1}{2} is not tight for Lemma 10d). However, we cannot confirm this at present, and thus leave it as a topic for future research.

Open Question 4.2 is another interesting topic for future work, which may further develop the results obtained by Cominetti et al. [2020], Patriksson [2004]; Josefsson and Patriksson [2007], Lu [2008] and Klimm and Warode [2021] for the differentiability of the PoA and Wardrop equilibria in non-atomic congestion games.

Acknowledgement

We thank the two anonymous referees for their very deep and constructive suggestions that have greatly helped to improve the paper.

Moreover, the first author acknowledges support from the National Science Foundation of China with grant No. 61906062, support from the Science Foundation of Anhui Science and Technology Department with grant No. 1908085QF262, and support from the Talent Foundation of Hefei University with grant No. 1819RC29; The first and second authors acknowledge support from the Science Foundation of the Anhui Education Department with grant No. KJ2019A0834.

References

  • Beckmann et al. [1956] M. J. Beckmann, C. McGuire, and C. B. Winsten. Studies in the economics of transportation. Yale Univ. Press, New Haven, CT, 1956.
  • Bingham et al. [1987] N. H. Bingham, C. M. Goldie, and J. L. Teugels. Regular variation. Cambrudge University Press, 1987.
  • Bureau of Public Roads [1964] Bureau of Public Roads. Traffic assignment manual. U.S. Department of Commerce, Urban Planning Division, Washington, D.C., 1964.
  • Colini-Baldeschi et al. [2020] R Colini-Baldeschi, R Cominetti, P Mertikopoulos, and M Scarsini. When is selfish routing bad? the price of anarchy in light and heavy traffic. Operations Research, 68(2):411–434, 2020. doi: https://doi.org/10.1287/opre.2019.1894.
  • Colini-Baldeschi et al. [2016] Riccardo Colini-Baldeschi, Roberto Cominetti, and Marco Scarsini. On the price of anarchy of highly congested nonatomic network games. In International Symposium on Algorithmic Game Theory, pages 117–128. Springer, Lecture Notes in Computer Science 9928, 2016.
  • Colini-Baldeschi et al. [2017] Riccardo Colini-Baldeschi, Roberto Cominetti, Panayotis Mertikopoulos, and Marco Scarsini. The asymptotic behavior of the price of anarchy. In WINE 2017, Lecture Notes in Computer Science 10674, pages 133–145, 2017.
  • Cominetti et al. [2020] R. Cominetti, V. Dose, and M. Scarsini. The price of anarchy in routing games as a function of the demand. Technical Report, arXiv:190710101v2 [cs.GT], July 2020.
  • Correa et al. [2004] José R. Correa, Andres S. Schulz, and Nicolás E. Stier Moses. Selfish routing in capacitated networks. Mathematics of Operations Research, 29(4):961–976, 2004.
  • Correa et al. [2005] José R. Correa, Andreas S. Schulz, and Nicolas E. Stier-Moses. On the inefficiency of equilibria in congestion games. extended abstract. In Integer Programming and Combinatorial Optimization, International IPCO Conference, Berlin, Germany, June 8-10, 2005, Proceedings, pages 167–181. Springer, Lecture Notes in Computer Science 3509, 2005.
  • Dafermos [1980] S. Dafermos. Traffic equilibrium and variational inequalities. Transportation Science, 14:42–54, 1980.
  • Dafermos and Sparrow [1969] S. C. Dafermos and F. T. Sparrow. The traffic assignment problem for a general network. Journal of Research of the U.S. National Bureau of Standards 73B, pages 91–118, 1969.
  • Englert et al. [2010] Matthias Englert, Thomas Franke, and Lars Olbrich. Sensitivity of Wardrop equilibria. Theory of Computing Systems, 47(1):3–14, 2010.
  • Hall [1978] Michael Hall. Properties of the equilibrium state in transportation networks. Transportation Science, 12(3):208–216, 1978.
  • Harks et al. [2015] Tobias Harks, Ingo Kleinert, Max Klimm, and Rolf H Möhring. Computing network tolls with support constraints. Networks, 65(3):262–285, 2015.
  • Jahn et al. [2005] Olaf Jahn, Rolf H. Möhring, Andreas S. Schulz, and Nicolás E. Stier-Moses. System-optimal routing of traffic flows with user constraints in networks with congestion. Operations Research, 53(4):600–616, 2005.
  • Josefsson and Patriksson [2007] Magnus Josefsson and Michael Patriksson. Sensitivity analysis of separable traffic equilibrium equilibria with application to bilevel optimization in network design. Transportation Research Part B, 41:4–31, 2007.
  • Kelley [1975] J. Kelley. General topology. Springer, 1975.
  • Klimm and Warode [2019] Max Klimm and Philipp Warode. Computing all wardrop equilibria parametrized by the flow demand. In Proceedings of the 30th Annual ACM-SIAM Symposium on Discrete Algorithms, pages 917–934. SIAM, 2019.
  • Klimm and Warode [2021] Max Klimm and Philipp Warode. Parametric computation of minimum-cost flows with piece-wise quadratic costs. Mathematics of Operations Research,, page to appear, 2021.
  • Koutsoupias and Papadimitriou [1999] E. Koutsoupias and C. H. Papadimitriou. Worst-case equilibria. In Proceedings of the 16th Annual Symposium on Theoretical Aspects of Computer Science (STACS), Trier, Germany, Volume 1563 of Lecture Notes in Computer Science, Springer, Heidelberg, pages 404–413, 1999.
  • Lu [2008] Shu Lu. Sensitivity of static traffic user equilibria with perturbations in arc cost function and travel demand. Transportation Science, 42(1):105–123, 2008.
  • Monnot et al. [2017] B. Monnot, F. Benita, and G. Piliouras. How bad is selfish routing in practice? Technical report. arXiv:1703.01599v2, 2017.
  • Nisan et al. [2007] Noam Nisan, Tim Roughgarden, Eva Tardos, and Vijay V. Vaz. Algorithmic game theory. Cambridge University Press, 2007.
  • O’Hare et al. [2016] Steven J. O’Hare, Richard D. Connors, and David P. Watling. Mechanisms that govern how the price of anarchy varies with travel demand. Transportation Research Part B Methodological, 84:55–80, 2016.
  • Papadimitriou [2001] Christos Papadimitriou. Algorithms, games, and the internet. In International Colloquium on Automata, Languages, and Programming, pages 1–3. Springer, Lecture Notes in Computer Science 2076, 2001.
  • Patriksson [2004] Michael Patriksson. Sensitivity analysis of traffic equilibria. Transportation Science, 38(3):258–281, 2004.
  • Perakis [2007] Georgia Perakis. The price of anarchy under nonlinear and asymmetric costs. Mathematics of operations research, 32(3):614–628, Aug. 2007.
  • Pigou [1920] A. C. Pigou. The Economics of Welfare. Macmillan and Co., London, 1st edn edition, 1920.
  • Rosenthal [1973] Robert W. Rosenthal. A class of games possessing pure-strategy nash equilibria. International Journal of Game Theory, 2(1):65–67, 1973.
  • Roughgarden and Tardos [2002] T Roughgarden and E Tardos. How bad is selfish routing? Journal of the ACM, 49:236–259, 2002.
  • Roughgarden [2001] Tim Roughgarden. Designing networks for selfish users is hard. Proceedings of Annual Symposium on Foundations of Computer Science, 72(72):472–481, 2001.
  • Roughgarden [2003] Tim Roughgarden. The price of anarchy is independent of the network topology. Journal of Computer & System Sciences, 67(2):341–364, 2003.
  • Roughgarden [2005] Tim Roughgarden. Selfish Routing and the Price of Anarchy. The MIT Press, 2005.
  • Roughgarden [2015] Tim Roughgarden. Intrinsic robustness of the price of anarchy. Journal of the ACM (JACM), 62(5):32, 2015.
  • Roughgarden and Tardos [2004] Tim Roughgarden and Eva Tardos. Bounding the inefficiency of equilibria in nonatomic congestion games. Games & Economic Behavior, 47(2):389–403, 2004.
  • Roughgarden and Tardos [2007] Tim Roughgarden and Eva Tardos. Introduction to the inefficiency of equilibria. In Noam Nisan, Tim Roughgarden, Éva Tardos, and Vijay V. Vazirani, editors, Algorithmic game theory, pages 443–459. Cambridge University Press, Cambridge, 2007.
  • Sandholm [2001] W. H. Sandholm. Potential games with continuous player sets. Journal of Economic Theory, 97:81–108, 2001.
  • Schmeidler [1973] David Schmeidler. Equilibrium points of nonatomic games. Journal of Statistical Physics, 7(4):295–300, 1973.
  • Smith [1979] M. J. Smith. The existence, uniqueness and stability of traffic equilibria. Transportation Research Part B Methodological, 13(4):295–304, 1979.
  • Takalloo and Kwon [2020] Mahdi Takalloo and Changhyun Kwon. Sensitivity of wardrop equilibria: Revisited. Optimization Letters, 14(3):781–796, 2020.
  • Wardrop [1952] J. G. Wardrop. Some theoretical aspects of road traffic research. Proceedings of the Institution of Civil Engineers, Part II, 1:325–378, 1952.
  • Wu et al. [2021] Zijun Wu, Rolf Möhring, Yanyan Chen, and Dachuan Xu. Selfishness need not be bad. Operations Research, 69(2):410–435, 2021. URL https://doi.org/10.1287/opre.2020.2036.
  • Youn et al. [2008] H Youn, M. T. Gastner, and H Jeong. Price of anarchy in transportation networks: efficiency and optimality control. Physical Review Letters, 101(12):128701, 2008.