Skip to main content
Log in

The Impact of Worst-Case Deviations in Non-Atomic Network Routing Games

  • Published:
Theory of Computing Systems Aims and scope Submit manuscript

Abstract

We introduce a unifying model to study the impact of worst-case latency deviations in non-atomic selfish routing games. In our model, latencies are subject to (bounded) deviations which are taken into account by the players. The quality deterioration caused by such deviations is assessed by the Deviation Ratio, i.e., the worst case ratio of the cost of a Nash flow with respect to deviated latencies and the cost of a Nash flow with respect to the unaltered latencies. This notion is inspired by the Price of Risk Aversion recently studied by Nikolova and Stier-Moses (Nikolova and Stier-Moses 2015). Here we generalize their model and results. In particular, we derive tight bounds on the Deviation Ratio for multi-commodity instances with a common source and arbitrary non-negative and non-decreasing latency functions. These bounds exhibit a linear dependency on the size of the network (besides other parameters). In contrast, we show that for general multi-commodity networks an exponential dependency is inevitable. We also improve recent smoothness results to bound the Price of Risk Aversion.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Notes

  1. We remark that for certain types of (0, β)-deviations, e.g., scaled marginal tolls, better bounds can be obtained; see the section “Relations to network toll problems” in Appendix B for relevant literature.

  2. For example, there are parallel-arc networks for which the Biased Price of Anarchy is unbounded, whereas the Deviation Ratio is a constant.

  3. Meir and Parkes [13] define a function l to be (1, μ)-smooth if xl(y) ≤ μyl(y) + xl(x) for all x, y ≥ 0 (which is slightly different from Roughgarden’s original smoothness definition [19]). Lineas et al. [11] only require local smoothness where y is taken fixed.

  4. The existence of a risk-averse Nash flow is proven in [15].

  5. Note that the values lP(x) + δP(x)are the same for all flow-carrying paths, but this is not necessarily true for the valueslP(x).

  6. Note that ηi ≤⌈(n − 1)/2⌉.

  7. Note that the pathsPlcan overlap, use parts of B, or even be subpaths of each other.

  8. We use the standard notation δ(v) and δ+(v) to refer to the set of outgoing and incoming edges of a node v, respectively.

  9. Note that the value ⌈(n − 1)/2⌉is the same for n ∈{2m, 2m + 1}with \(m \in \mathbb {N}\).The example shows tightness for n = 2m.The tightness for n = 2m + 1then follows trivially by adding a dummy node.

  10. For example \(y_{m}(g) = m(m-1)\beta \max \{0,(g - \frac {1}{m})\}\).That is, we define ymto be zero for 0 ≤ g ≤ 1/m and we let it increase with constant rate toβ in 1/(m − 1).

  11. We assume that the infimum and supremum are attained in the set Δ(𝜃); in particular, this is true for (α, β)-deviations.

References

  1. Beckmann, M., McGuire, B., Winsten, C.: Studies in the Economics of Transportation. Yale University Press, New Haven (1956)

    Google Scholar 

  2. Bonifaci, V., Salek, M., Schäfer, G.: On the efficiency of restricted tolls in network routing games. Lecture Notes in Computer Science (2011)

  3. Chen, P.-A., Kempe, D.: Altruism, selfishness, and spite in traffic routing. In: Proceedings of the 9th ACM conference on electronic commerce, pp 140–149. ACM (2008)

  4. Christodoulou, G., Koutsoupias, E., Spirakis, P. G.: On the performance of approximate equilibria in congestion games. Algorithmica 61(1), 116–140 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  5. Cominetti, R.: Equilibrium routing under uncertainty. Math. Program. 151 (1), 117–151 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  6. Correa, J.R., Schulz, A.S., Stier-Moses, N.E.: A geometric approach to the price of anarchy in nonatomic congestion games. Games and Economic Behavior 64 (2), 457–469 (2008). Special Issue in Honor of Michael B. Maschler

    Article  MathSciNet  MATH  Google Scholar 

  7. Englert, M., Franke, T., Olbrich, L.: Sensitivity of Wardrop Equilibria, pp. 158–169. Springer, Berlin (2008)

    MATH  Google Scholar 

  8. Fotakis, D., Kalimeris, D., Lianeas, T.: Improving selfish routing for risk-averse players. In: Proceedings of Web and Internet Economics - 11th International Conference, WINE 2015, Amsterdam, the netherlands, december 9–12, 2015, pp 328–342 (2015)

  9. Hoefer, M., Olbrich, L., Skopalik, A.: Taxing subnetworks. In: Papadimitriou, C.H., Zhang, S. (eds.) WINE, volume 5385 of Lecture Notes in Computer Science, pp. 286–294. Springer (2008)

  10. Kleer, P., Schäfer, G.: Path Deviations Outperform Approximate Stability in Heterogeneous Congestion Games, pp. 212–224. Springer International Publishing Cham, Berlin (2017)

  11. Lianeas, T., Nikolova, E., Stier-Moses, N.E.: Asymptotically tight bounds for inefficiency in risk-averse selfish routing. CoRR, arXiv:1510.02067(2015)

  12. Lin, H., Roughgarden, T., Tardos, É., Walkover, A.: Stronger bounds on braess’s paradox and the maximum latency of selfish routing. SIAM J. Discret. Math. 25(4), 1667–1686 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  13. Meir, R., Parkes, D.: Playing the wrong game smoothness bounds for congestion games with behavioral biases. SIGMETRICS Perform. Eval. Rev. 43(3), 67–70 (2015)

    Article  Google Scholar 

  14. Meir, R., Parkes, D.C.: Congestion games with distance-based strict uncertainty. CoRR, arXiv:1411.4943 (2014)

  15. Nikolova, E., Stier-Moses, N.E.: A mean-risk model for the traffic assignment problem with stochastic travel times. Oper. Res. 62(2), 366–382 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  16. Nikolova, E., Stier-Moses, N.E.: The Burden of Risk Aversion in Mean-Risk Selfish Routing. In: Proceedings of the 16th ACM Conference on Economics and Computation, EC ’15, pp. 489–506. ACM, New York (2015)

  17. Nisan, N., Roughgarden, T., Tardos, E., Vazirani, V.: Algorithmic Game Theory. Cambridge University Press, New York (2007)

    Book  MATH  Google Scholar 

  18. Roughgarden, T.: On the severity of braess’s paradox: Designing networks for selfish users is hard. J. Comput. Syst. Sci. 72(5), 922–953 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  19. Roughgarden, T.: Intrinsic robustness of the price of anarchy. J. ACM 62(5), 32 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  20. Wardrop, J. G.: Some theoretical aspects of road traffic research. Proc. Inst. Civ. Eng. 1, 325–378 (1952)

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Guido Schäfer.

Additional information

This article is part of the Topical Collection on Special Issue on Algorithmic Game Theory (SAGT 2016)

Pieter Kleer is supported by the NWO Gravitation Project NETWORKS, Grant Number 024.002.003.

Appendices

Appendix A: Proof of Lemma 1

Lemma 1

Let− 1 < α ≤ 0 ≤ βbe fixed. Thenf is inducible with an (α, β)-deviationif and only if it is inducible with a\((0,\frac {\beta - \alpha }{1 + \alpha })\)-deviation.

Proof

Let f be δ-inducible for some αlδβl, and for aA, write δa(fa) = dala(fa). Without loss of generality we may assume that δa(x) = dala(x)(since by definition dala(x) also induces f). From the equilibrium conditions (2), we know that

$$\forall i \in [k], \forall P \in \mathcal{P}_{i}, f_{P} > 0: \quad \sum\limits_{a \in P} l_{a}(f_{a}) + \delta_{a}(f_{a}) \leq \sum\limits_{a \in P^{\prime}} l_{a}(f_{a}) + \delta_{a}(f_{a}) \ \ \forall P^{\prime} \in \mathcal{P}_{i}. $$

This is equivalent to \(\forall i \in [k], \forall P \in \mathcal {P}_{i}, f_{P} > 0:\)

$$\sum\limits_{a \in P} \left(1 + \frac{d_{a} - \alpha}{1 + \alpha}\right)l_{a}(f_{a}) \leq \sum\limits_{a \in P^{\prime}} \left(1 + \frac{d_{a} - \alpha}{1 + \alpha}\right)l_{a}(f_{a}) \ \ \forall P^{\prime} \in \mathcal{P}_{i} $$

which can be seen bywriting

$$l_{a}(f_{a}) + \delta_{a}(f_{a}) = (1 + d_{a})l_{a}(f_{a}) = (1 + \alpha + d_{a} - \alpha)l_{a}(f_{a}),$$

and then dividingthe inequality by 1 + α.We then see that δ,defined by \(\delta _{a}^{\prime }(x) = \frac {d_{a} - \alpha }{1 + \alpha } l_{a}(x)\) for all aA and x ≥ 0, also induces f since

$$\alpha l_{a}(x) \leq d_{a}l_{a}(x) \leq \beta l_{a}(x) \ \ \ \Leftrightarrow \ \ \ 0 \leq \frac{d_{a} - \alpha}{1 + \alpha} l_{a}(x) \leq \frac{\beta - \alpha}{1 + \alpha} l_{a}(x). $$

Appendix B: Computing Optimal Deviations

The bounded deviation model introduced in Section 2.2 naturally gives rise to the following two optimization problems:Footnote 11

  1. 1.

    Best deviation problem: compute a deviation δ∈ Δ(𝜃) such that

    $$\delta^{*} = \arg\inf_{\delta \in {\Delta}(\theta)} \inf \{C(f^{\delta}) \; | \; f^{\delta}\text{ is }\delta\text{-inducible}\}. $$
  2. 2.

    Worst deviation problem: compute a deviation δ∈ Δ(𝜃) such that

    $$\delta^{*} = \arg\sup_{\delta \in {\Delta}(\theta)} \sup \{C(f^{\delta}) \; | \; f^{\delta}\text{ is }\delta\text{-inducible}\}. $$

Recall that the social cost function C only takes into account the latencies but not the deviations. A somewhat subtle point here is that for a fixed deviation δ ∈ Δ(𝜃), the social cost of a δ-inducible flow might not be unique. In particular, in the best deviation problem we seek a feasible deviation δ such that the social cost of the best Nash flow that is δ-inducible is minimized (similar as in [2]). In contrast, in the worst deviation problem we want to determine a feasible deviation δ such that the social cost of the worst Nash flow that is δ-inducible is maximized.

Below we elaborate on relations between the best deviation problem and various network toll problems. As a side result, we also show that the worst deviation problem is NP-hard, even for single-commodity instances with linear latencies (Theorem 12).

2.1 B.1 Relations to Network Toll Problems

The best deviation problem is a direct generalization of the restricted network toll problem introduced by Bonifaci et al. [2]. We obtain this model for 𝜃min = 0. The deviations are interpreted as non-negative tolls on the arcs. The objective minimized in [2] is measured against the social optimum, i.e., the authors are interested in the ratio C(fδ)/C(f), where f is an optimal flow for the instance \(\mathcal {I}\). Also, our definition of (0, β)-deviations is equivalent to the definition of β-restricted tolls in [2].

The work by Fotakis et al. [8] can technically be seen as a variant of the restricted network toll problem in which the tolls are interpreted as risk-averse behavior of players. Here, we have \(\theta ^{\min }_{a} = 0\) and \(\theta ^{\max }_{a} = \gamma l_{a}\) for all aA. The authors consider deviations of the form δa(x) = γala(x) for 0 ≤ γaγ for all aA. In particular, deviations of this form induce an approximate Nash flow as studied by Christodoulou et al. [4]. For example, if all latency functions in the network are polynomials of degree at most d, then we obtain a γd-approximate Nash flow.

Hoefer et al. [9] consider the taxing subnetwork problem, which is a special case of the restricted network toll problem. Here only a designated subset of the arcs can be tolled, which is equivalent to \(\theta ^{\min }_{a} = 0\) and \(\theta ^{\max }_{a} \in \{0,\infty \}\) for all aA. They show that best deviation problem is NP-complete, even for two commodities. To the best of our knowledge, the single-commodity case is still an open problem. On the positive side, Hoefer et al. [9] and Bonifaci et al. [2] give polynomial time algorithms for parallel-arc networks, solving the best deviation problem for their respective definitions of the threshold functions.

Beckmann et al. [1] proved that the social optimum can be induced as a Nash flow using marginal tolls, that is, by setting δa(x) = xla′(x), where \(l^{\prime }_{a}(x)\) is the derivative of la(x) (assuming the existence of \(l_{a}^{\prime }\)). In particular, if these tolls are feasible, i.e., δ ∈ Δ(𝜃), then δ is an optimal solution for the best deviation problem.

There are several models that study deviations in the form of scaled marginal tolls, i.e., deviations defined by δa(x) = ρxla′(x) for some \(\rho \in \mathbb {R}\). We elaborate on two such models in more detail:

In the standard non-atomic routing model it is assumed that players are completely selfish in the sense that they want to minimize their own latencies. However, more recently researchers also considered settings where players are (partially) altruistic. Chen et al. [3], for example, model such altruistic behavior by including scaled marginal tolls in the objective of the players. In particular, they study scaled marginal tolls with − 1 ≤ ρ ≤ 1.

Meir and Parkes [13] also study deviations in the form of scaled marginal tolls, which are interpreted as behavioral biases towards the marginal tolls. A conceptual difference here is that the parameter ρ is chosen by the players, instead of the system designer (as, for example, in the restricted network toll model). Here, the authors are also interested in the case ρ ≥ 1 (which is less relevant in the other models). The authors also study this model in [14], where these deviations are interpreted as distance-based strict uncertainty.

2.2 B.2 Hardness of the Worst Deviation Problem

As a side-result, we prove that the problem of determining worst-case deviations is NP-hard.

Theorem 12

It isNP-hard to compute deviationsδ ∈ Δ(𝜃) such thatC(fδ) is maximized, even for single-commodity networks with linear latencies.

Proof

We give a reduction from the Directed Hamiltonians, t-Path problem: We are given a directed graph G = (V, A),and fixed s, tV,and the goal is to decide whether or not there exists a simple directeds, t-pathin G that visits every node exactly once. Let\(\mathcal {J}\) be an instance of Directed Hamiltonians, t-Path problem.

Now, define an instance \(\mathcal {I}\)of the bounded deviation model on the graph G by takingla(x) = x for allaA,\(\theta ^{\min }_{a} = 0\)for allaA, and\(\theta ^{\max }_{a} = n - 1\)for allaA. Furthermore,take r = 1. We claim that G has a Hamiltonian path from s to t if and only if there is a deviationδ ∈ Δ(𝜃)suchthat C(fδ) ≥ n − 1.First, let G have a Hamiltonian path P from s to t, and defineδ byδa = 0ifaP, andδa = n − 1otherwise. Wethen have that fδis given by \(f_{a}^{\delta } = 1\)if aP and\(f^{\delta }_{a} = 0\)otherwise, since the perceived latency along P is then equal tolP(fδ) = n − 1, and any other pathPuses at least one differentarc aP, which gives usthat

$$l_{P^{\prime}}(f^{\delta}) + \delta_{P^{\prime}}(f^{\delta}) \geq l_{a^{\prime}}(f^{\delta}) + \delta_{a^{\prime}}(f^{\delta}) \geq n-1 = l_{P}(f^{\delta}) + \delta_{P}(f^{\delta}). $$

Notethat fδis the unique Nash flow in this case (since all the perceived latenciesla + δaarestrictly increasing).

Conversely, suppose there is a δ ∈ Δ(𝜃)such that C(fδ) ≥ n − 1. For any feasible flow g we have that lP(g) ≤ n − 1, withstrict inequality if \(f^{\delta }_{P} < 1\)(since thenthere will be at least one arc aP with \(f^{\delta }_{a} < 1\)). This meansthat

$$C(g) = \sum\limits_{P \in \mathcal{P}} g_{P}l_{P}(g) \leq \sum\limits_{P \in \mathcal{P}} g_{P}(n-1) = n - 1,$$

using thatr = 1. Again, we havestrict inequality if 0 < gP < 1for some path P, i.e., if not all players use the same path. This means that forfδthere is at mostone path Pwith\(f^{\delta }_{P^{*}} > 0\), which then impliesthat \(f^{\delta }_{P^{*}} = 1\). Furthermore,we can conclude that \(|A(P^{*})| = l_{P^{*}}(f^{\delta }) = C(f^{\delta }) = n - 1\),which implies that Pis a Hamiltonian path from s to t, since it is a simple path by assumption.□

Appendix C: Necessity of Common Source Assumption in Theorem 2

The example below shows that Theorem 2 does not hold if the assumption that all commodities share a common source is dropped.

Example 2

Consider the graph G = (V, A)in Fig. 6 and suppose that r1 = r2 = 1.Then the flow f that routes one unit of flow over both paths(s1, v1, 1, 2, t1)and (s2, v2, 3, 4, t2)is feasible and inducible (take δ = 0).However, looking at the graph \(\hat {G}(f)\),we obtain a negative cost cycle (1, 4, 3, 2, 1)(by using the reversed arcs of (1,2) and (3,4)).

Fig. 6
figure 6

All the values of \(l_{a}, \ \theta ^{\min }_{a}\) and \(\theta ^{\max }_{a}\) that are not explicitly stated are zero

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kleer, P., Schäfer, G. The Impact of Worst-Case Deviations in Non-Atomic Network Routing Games. Theory Comput Syst 63, 54–89 (2019). https://doi.org/10.1007/s00224-017-9829-y

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00224-017-9829-y

Keywords

Navigation