implicit learning dynamics in stackelberg games...

38
000 001 002 003 004 005 006 007 008 009 010 011 012 013 014 015 016 017 018 019 020 021 022 023 024 025 026 027 028 029 030 031 032 033 034 035 036 037 038 039 040 041 042 043 044 045 046 047 048 049 050 051 052 053 054 Implicit Learning Dynamics in Stackelberg Games: Equilibria Characterization, Convergence Analysis, and Empirical Study Anonymous Authors 1 Abstract Contemporary work on learning in continuous games has commonly overlooked the hierarchi- cal decision-making structure present in machine learning problems formulated as games, instead treating them as simultaneous play games and adopting the Nash equilibrium solution concept. We deviate from this paradigm and provide a com- prehensive study of learning in Stackelberg games. This work provides insights into the optimization landscape of zero-sum games by establishing con- nections between Nash and Stackelberg equilibria along with the limit points of simultaneous gra- dient descent. We derive novel gradient-based learning dynamics emulating the natural structure of a Stackelberg game using the Implicit Function Theorem and provide convergence analysis for deterministic and stochastic updates for zero-sum and general-sum games. Notably, in zero-sum games using deterministic updates, we show the only critical points the dynamics converge to are Stackelberg equilibria and provide a local conver- gence rate. Empirically, our learning dynamics mitigate rotational behavior and exhibit benefits for training generative adversarial networks com- pared to simultaneous gradient descent. 1. Introduction The emerging coupling between game theory and machine learning can be credited to the formulation of learning problems as interactions between competing objectives and strategic agents. Indeed, generative adversarial networks (GANs) (Goodfellow et al., 2014), robust supervised learn- ing (Madry et al., 2018), reinforcement and multi-agent reinforcement learning (Dai et al., 2018; Zhang et al., 2019), and hyperparameter optimization (Maclaurin et al., 2015) 1 Anonymous Institution, Anonymous City, Anonymous Region, Anonymous Country. Correspondence to: Anonymous Author <[email protected]>. Preliminary work. Under review by the International Conference on Machine Learning (ICML). Do not distribute. problems can be cast as zero-sum or general-sum continu- ous action games. To obtain solutions in a tractable manner, gradient-based algorithms have gained attention. Given the motivating applications, much of the contempo- rary work on learning in games has focused on zero-sum games with non-convex, non-concave objective functions and seeking stable fixed points or local equilibria. A number of techniques have been proposed including optimistic and extra-gradient algorithms (Daskalakis & Panageas, 2018; Daskalakis et al., 2018; Mertikopoulos et al., 2019), gra- dient adjustments (Balduzzi et al., 2018; Mescheder et al., 2017), and opponent modeling methods (Foerster et al., 2018; Letcher et al., 2019; Sch ¨ afer & Anandkumar, 2019; Zhang & Lesser, 2010). However, only a select number of algorithms can guarantee convergence to stable fixed points satisfying sufficient conditions for a local Nash equilibrium (LNE)(Adolphs et al., 2019; Mazumdar et al., 2019). The dominant perspective in machine learning applications of game theory has been focused on simultaneous play. However, there are many problems exhibiting a hierarchical order of play, and in a game theoretic context, such problems are known as Stackelberg games. The Stackelberg equilib- rium (Von Stackelberg, 2010) solution concept generalizes the min-max solution to general-sum games. In the simplest formulation, one player acts as the leader who is endowed with the power to select an action knowing the other player (follower) plays a best-response. This viewpoint has long been researched from a control perspective on games (Basar & Olsder, 1998) and in the bilevel optimization commu- nity (Danskin, 1966; 1967; Zaslavski, 2012). The work from a machine learning perspective on games with a hierarchical decision-making structure is sparse and exclusively focuses on zero-sum games (Jin et al., 2019; Metz et al., 2017). In the work of Metz et al. (2017) on unrolled GANs, the follower (discriminator) is allowed to ‘roll out’ parameter updates until reaching an approximate local optimum between each update of the leader (genera- tor). This behavior intuitively approximates a hierarchical order of play. The success of the unrolling method as a training mechanism provides some evidence supporting the efficacy of local Stackelberg equilibrium (LSE) as a solution concept. However, much remains to be discovered on the

Upload: others

Post on 02-Jun-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

000001002003004005006007008009010011012013014015016017018019020021022023024025026027028029030031032033034035036037038039040041042043044045046047048049050051052053054

Implicit Learning Dynamics in Stackelberg Games:Equilibria Characterization, Convergence Analysis, and Empirical Study

Anonymous Authors1

AbstractContemporary work on learning in continuousgames has commonly overlooked the hierarchi-cal decision-making structure present in machinelearning problems formulated as games, insteadtreating them as simultaneous play games andadopting the Nash equilibrium solution concept.We deviate from this paradigm and provide a com-prehensive study of learning in Stackelberg games.This work provides insights into the optimizationlandscape of zero-sum games by establishing con-nections between Nash and Stackelberg equilibriaalong with the limit points of simultaneous gra-dient descent. We derive novel gradient-basedlearning dynamics emulating the natural structureof a Stackelberg game using the Implicit FunctionTheorem and provide convergence analysis fordeterministic and stochastic updates for zero-sumand general-sum games. Notably, in zero-sumgames using deterministic updates, we show theonly critical points the dynamics converge to areStackelberg equilibria and provide a local conver-gence rate. Empirically, our learning dynamicsmitigate rotational behavior and exhibit benefitsfor training generative adversarial networks com-pared to simultaneous gradient descent.

1. IntroductionThe emerging coupling between game theory and machinelearning can be credited to the formulation of learningproblems as interactions between competing objectives andstrategic agents. Indeed, generative adversarial networks(GANs) (Goodfellow et al., 2014), robust supervised learn-ing (Madry et al., 2018), reinforcement and multi-agentreinforcement learning (Dai et al., 2018; Zhang et al., 2019),and hyperparameter optimization (Maclaurin et al., 2015)

1Anonymous Institution, Anonymous City, Anonymous Region,Anonymous Country. Correspondence to: Anonymous Author<[email protected]>.

Preliminary work. Under review by the International Conferenceon Machine Learning (ICML). Do not distribute.

problems can be cast as zero-sum or general-sum continu-ous action games. To obtain solutions in a tractable manner,gradient-based algorithms have gained attention.

Given the motivating applications, much of the contempo-rary work on learning in games has focused on zero-sumgames with non-convex, non-concave objective functionsand seeking stable fixed points or local equilibria. A numberof techniques have been proposed including optimistic andextra-gradient algorithms (Daskalakis & Panageas, 2018;Daskalakis et al., 2018; Mertikopoulos et al., 2019), gra-dient adjustments (Balduzzi et al., 2018; Mescheder et al.,2017), and opponent modeling methods (Foerster et al.,2018; Letcher et al., 2019; Schafer & Anandkumar, 2019;Zhang & Lesser, 2010). However, only a select number ofalgorithms can guarantee convergence to stable fixed pointssatisfying sufficient conditions for a local Nash equilibrium(LNE) (Adolphs et al., 2019; Mazumdar et al., 2019).

The dominant perspective in machine learning applicationsof game theory has been focused on simultaneous play.However, there are many problems exhibiting a hierarchicalorder of play, and in a game theoretic context, such problemsare known as Stackelberg games. The Stackelberg equilib-rium (Von Stackelberg, 2010) solution concept generalizesthe min-max solution to general-sum games. In the simplestformulation, one player acts as the leader who is endowedwith the power to select an action knowing the other player(follower) plays a best-response. This viewpoint has longbeen researched from a control perspective on games (Basar& Olsder, 1998) and in the bilevel optimization commu-nity (Danskin, 1966; 1967; Zaslavski, 2012).

The work from a machine learning perspective on gameswith a hierarchical decision-making structure is sparse andexclusively focuses on zero-sum games (Jin et al., 2019;Metz et al., 2017). In the work of Metz et al. (2017) onunrolled GANs, the follower (discriminator) is allowed to‘roll out’ parameter updates until reaching an approximatelocal optimum between each update of the leader (genera-tor). This behavior intuitively approximates a hierarchicalorder of play. The success of the unrolling method as atraining mechanism provides some evidence supporting theefficacy of local Stackelberg equilibrium (LSE) as a solutionconcept. However, much remains to be discovered on the

Page 2: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

055056057058059060061062063064065066067068069070071072073074075076077078079080081082083084085086087088089090091092093094095096097098099100101102103104105106107108109

Implicit Learning Dynamics in Stackelberg Games

role of such equilibria in the game optimization landscape.The methods of Metz et al. (2017) are formalized by Jinet al. (2019), who show in zero-sum games the stable fixedpoints of simultaneous gradient descent with a timescaleseparation between players approaching infinity satisfy suf-ficient conditions for a LSE. It remains an open question todevelop implementable algorithms with similar guarantees.

Contributions. Motivated by the lack of algorithms focus-ing on games exhibiting an order of play, we provide a studyof learning in Stackelberg games including equilibria char-acterization, novel learning dynamics and convergence anal-ysis, and an illustrative empirical study. The primary bene-fits of this work to the community include an enlightenedperspective on the consideration of equilibrium conceptsreflecting the underlying optimization problems present inmachine learning applications formulated as games and analgorithm that provably converges to stable fixed points sat-isfying sufficient conditions for a LSE in zero-sum games.

We provide a characterization of LSE via sufficient condi-tions on player objectives and term points satisfying theconditions differential Stackelberg equilibria (DSE). Weshow DSE are generic amongst LSE in zero-sum games.This means except on a set of measure zero in the classof zero-sum continuous games, DSE and LSE are equivalent.While the placement of differential Nash equilibria (DNE)amongst critical points in continuous games is reasonablywell understood, an equivalent statement cannot be maderegarding DSE. Accordingly, we draw connections betweenthe solution concepts in the class of zero-sum games. Weshow that DNE are DSE, which indicates the solution conceptin hierarchical play games is not as restrictive as the solu-tion concept in simultaneous play games. Furthermore, wereveal that there exist stable fixed points of simultaneousgradient descent dynamics that are DSE and not DNE. Thisinsight gives meaning to a broad class of fixed points pre-viously thought to lack game-theoretic meaning and maygive some explanation for the adequacy of solutions notsatisfying sufficient conditions for LNE in GANs. To charac-terize this phenomenon, we provide necessary and sufficientconditions for when such points exist.

We derive novel gradient-based learning dynamics emulat-ing the natural structure of a Stackelberg game from thesufficient conditions for a LSE and the implicit functiontheorem. The dynamics can be viewed as an analogue tosimultaneous gradient descent incorporating the structureof hierarchical play games. In stark contrast to the simul-taneous play counterpart, we show in zero-sum games theonly stable fixed points of the dynamics are DSE and suchequilibria must be stable fixed points of the dynamics. Us-ing this fact and saddle avoidance results, we show the onlycritical points the discrete time algorithm converges to givendeterministic gradients are DSE and provide a local conver-

gence rate. In general-sum games, we cannot guarantee theonly critical point attractors of the deterministic learningalgorithms are DSE. However, we give a local convergencerate to critical points which are DSE. For stochastic gradi-ent updates, we obtain analogous convergence guaranteesasymptotically for each game class.

This work finishes with an empirical study. We demonstrateour proposed dynamics mitigate rotational dynamics whentraining GANs and provide an example in which simulta-neous gradient descent avoids DNE and converges to theneighborhood of a DSE. Given recent evidence that stan-dard training techniques do not converge to DNE but insteadto stable points that are locally optimal for the discrimina-tor (Berard et al., 2019), this motivates future investigationto determine if successful training methods often reach DSE.

2. PreliminariesWe now formalize the games we study, present equilibriumconcepts accompanied by sufficient condition characteriza-tions, and formulate Stackelberg learning dynamics.

2.1. Game Formalisms

Consider a non-cooperative game between two agents whereplayer 1 is deemed the leader and player 2 the follower. Theleader has cost f1 : X → R and the follower has costf2 : X → R, where X = X1 ×X2 with X1 and X2 denot-ing the action spaces of the leader and follower, respectively.Here X is a subset of an m-dimensional Euclidean spaceand similarly, each Xi is a subset of an mi-dimensionalEuclidean space. We assume throughout that each fi is suf-ficiently smooth: fi ∈ Cq(X,R) for some q ≥ 2. For zero-sum games, the game is defined by costs (f1, f2) = (f,−f).In words, we the class of two-player smooth games on con-tinuous, unconstrained actions spaces. The designation of‘leader’ and ‘follower’ indicates the order of play betweenthe agents, meaning the leader plays first and the followersecond. This game is called a Stackelberg game.

In a Stackelberg game, the leader aims to solve the opti-mization problem given by minx1∈X1

{f1(x1, x2)∣∣ x2 ∈

arg miny∈X2 f2(x1, y)} and the follower aims to solve theoptimization problem minx2∈X2 f2(x1, x2). We can com-pare the structure of the Stackelberg game with a simulta-neous play game in which each player is faced with theoptimization problem minxi∈Xi fi(xi, x−i). The learningalgorithms we formulate are such that the agents followmyopic update rules which take steps in the direction ofsteepest descent for the respective optimizations problems.

2.2. Equilibria Concepts and Characterizations

Before formalizing learning rules, let us first discuss theequilibrium concept studied for simultaneous play games

Page 3: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

110111112113114115116117118119120121122123124125126127128129130131132133134135136137138139140141142143144145146147148149150151152153154155156157158159160161162163164

Implicit Learning Dynamics in Stackelberg Games

and contrast it with that which is studied in the hierarchi-cal play counterpart. The typical equilibrium notion incontinuous games is the pure strategy Nash equilibrium insimultaneous play games and the Stackelberg equilibriumin hierarchical play games. Each notion of equilibria canbe characterized as the intersection points of the reactioncurves of the players (Basar & Olsder, 1998). We focus ourattention on local notions of the equilibrium concepts as isstandard in learning in games since the objective functionswe consider need not be convex or concave.Definition 1 (Local Nash (LNE)). The joint strategy x∗ ∈ Xis a local Nash equilibrium on U1 × U2 ⊂ X1 ×X2 if foreach i ∈ {1, 2}, fi(x∗) ≤ fi(xi, x∗−i), ∀ xi ∈ Ui ⊂ Xi.

Definition 2 (Local Stackelberg (LSE)). Consider Ui ⊂Xi for each i ∈ {1, 2}. The strategy x∗1 ∈ U1 is a localStackelberg solution for the leader if, ∀x1 ∈ U1,

supx2∈RU2(x∗

1)f1(x∗1, x2) ≤ supx2∈RU2

(x1) f1(x1, x2),

where RU2(x1) = {y ∈ U2|f2(x1, y) ≤ f2(x1, x2),∀x2 ∈U2}. Moreover, (x∗1, x

∗2) for any x∗2 ∈ RU2

(x∗1) is a localStackelberg equilibrium on U1 × U2.

While characterizing existence of equilibria is outsidethe scope of this work, we remark that Nash equilibriaexist for convex costs on compact and convex strategyspaces and Stackelberg equilibria exist on compact strategyspaces (Basar & Olsder, 1998, Thm. 4.3, Thm. 4.8, & §4.9).This means the class of games on which Stackelberg equi-libria exist is broader than on which Nash equilibria exist.Existence of local equilibria is guaranteed if the neighbor-hoods and cost functions restricted to those neighborhoodssatisfy the assumptions of the cited results.

Predicated on existence, equilibria can be characterized interms of sufficient conditions on player costs. We denoteDifi as the derivative of fi with respect to xi, Dijfi as thepartial derivative of Difi with respect to xj , and D(·) as thetotal derivative.1 The following gives sufficient conditionsfor a LNE as given in Definition 1.Definition 3 (Differential Nash (DNE) Ratliff et al. (2016)).The joint strategy x∗ ∈ X is a differential Nash equilibriumif Difi(x

∗) = 0 and D2i fi(x

∗) > 0 for each i ∈ {1, 2}.

Analogous sufficient conditions can be stated to characterizea LSE from Definition 2.Definition 4 (Differential Stackelberg (DSE)). The jointstrategy x∗ = (x∗1, x

∗2) ∈ X is a differential Stackelberg

equilibrium if Df1(x∗) = 0, D2f2(x∗) = 0, D2f1(x∗) >0, and D2

2f2(x∗) > 0 where x∗2 = r(x∗1) and r(·) implicitlydefined by D2f2(x∗) = 0.

Game Jacobians play a key role in determining stability ofcritical points. Let ω(x) = (D1f1(x), D2f2(x)) be the vec-

1For example, given f(x, r(x)), Df = D1f +D2f∂r/∂x.

tor of individual gradients for the simultaneous play gameand ωS(x) = (Df1(x), D2f2(x)) as the equivalent for theStackelberg game. Observe that Df1 is the total derivativeof f1 with respect to x1 given x2 is implicitly a function ofx1, capturing the fact that the leader believes the followerwill play a best response to x1. We note that the reactioncurve of the follower to x1 may not be unique. However,under sufficient conditions on a local Stackelberg solution x,locally D2f2(x) = 0 and det(D2

2f2(x)) 6= 0 so that Df1 iswell defined via the implicit mapping theorem (Lee, 2012).

The vector field ω(x) forms the basis of the well-studiedsimultaneous gradient learning dynamics and the Jacobianof the dynamics is given by

J(x) =

[D2

1f1(x) D12f1(x)D21f2(x) D2

2f2(x)

].

Similarly, the vector field ωS(x) serves as the foundationof the learning dynamics we formulate in Section 2.3 andanalyze throughout. The Jacobian of the Stackelberg vectorfield ωS(x) is given by

JS(x) =

[D1(Df1(x)) D2(Df1(x))D21f2(x) D2

2f2(x)

]. (1)

We say a critical point is stable with respect to a vector fieldV (x) if the spectrum of the Jacobian of V is in the openleft-half complex plane; e.g., x such that ω(x) = 0 is stableif spec(−J(x)) ⊂ C◦−. A critical point is non-degenerate ifthe determinant of the vector field Jacobian is non-zero.

Noting that the Schur complement of JS(x) with respect toD2

2f2(x) is identically D2f(x1, r(x1)), we give alternativebut equivalent sufficient conditions as those in Definition 4in terms of JS(x). This will be used to explicitly draw aconnection between the Stackelberg learning dynamics andthe sufficient conditions for a local Stackelberg. Here, S1(·)denotes the Schur complement of (·) with respect to thebottom block matrix in (·).Proposition 1. Consider a game (f1, f2) defined by fi ∈Cq(X,R), i = 1, 2 with q ≥ 2 and player 1 (withoutloss of generality) taken to be the leader. Let x∗ satisfyD2f2(x∗) = 0 and D2

2f2(x∗) > 0. Then Df1(x∗) = 0and S1(JS(x∗)) > 0 if and only if x∗ is a DSE. Moreover,in zero-sum games, S1(JS(x)) = S1(J(x)).

The proof is in Appendix G.1.

Genericity, Structural Stability, and Bilinear Games. Anatural question is how common is it for local equilibria tosatisfy sufficient conditions, meaning in a formal mathemat-ical sense, what is the gap between necessary and sufficientconditions in games. Towards addressing this, it has beenshown that DNE are generic amongst LNE and structurallystable in the classes of zero-sum and general-sum continu-ous games, respectively (Mazumdar & Ratliff, 2019; Ratliff

Page 4: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

165166167168169170171172173174175176177178179180181182183184185186187188189190191192193194195196197198199200201202203204205206207208209210211212213214215216217218219

Implicit Learning Dynamics in Stackelberg Games

et al., 2016). The results say that except on a set of measurezero in each class of games, DNE = LNE and the equilibriapersist under sufficiently smooth perturbations to the costs.We obtain analogous results for DSE in the class of zero-sumgames using similar proof techniques. The formal resultsare stated and proven in Appendix I and allow us to concludefor a generic zero-sum game DSE = LSE.

However, important classes of non-generic games do exist.In games where the cost function of the follower is bilinear,LSE can exist which do not satisfy the sufficient conditionsoutlined in Definition 4. As a simple example, x∗ = (0, 0)is a LSE for the zero-sum game defined by f(x1, x2) =x1x2 and not a DSE since D2

2f2(x) = 0 ∀ x ∈ X . Sincesuch games belong to a degenerate class in the contextof the genericity result we provide, they naturally deservespecial attention and algorithmic methods. While we do notfocus our attention on this class of games, we do proposesome remedies to allow our proposed learning algorithm tosuccessfully seek out equilibria in them. In the experimentssection, we discuss a regularized version of our dynamicsthat injects a small perturbation to cure degeneracy problemsleveraging the fact that DSE are structurally stable. Furtherdetails can be found in Appendix B. Finally, for bimatrixgames with finite actions it is common to reparameterize theproblem using a softmax function to obtain mixed policieson the simplex (Fudenberg et al., 1998). We explore thisviewpoint in Appendix D on a parameterized bilinear game.

2.3. Stackelberg Learning Dynamics

Recall that ωS(x) = (Df1(x), D2f2(x)) is the vector fieldfor Stackelberg games and it, along with its Jacobian JS(x),characterize sufficient conditions for a DSE. Letting ωS,ibe the i–th component of ωS , the leader total derivative isωS,1(x) = D1f1(x) − D21f2(x)>(D2

2f2(x))−1D2f1(x)where Dr ≡ −(D2

2f2(x))−1 ◦D21f2 with r defined by theimplicit function theorem (Lee, 2012) in a neighborhoodof a differential Stackelberg, meaning where ωS,2(x) = 0with det(D2

2f2(x)) 6= 0 (which is also holds generically atcritical points by Lemma 7). The Stackelberg learning rulewe study for each player in discrete time is given by

xi,k+1 = xi,k − γi,khS,i(xk). (2)

In deterministic learning players have oracle gradient accessso hS,i(x) = ωS,i(x). In stochastic learning players haveunbiased gradient estimates and hS,i(xk) = ωS,i(xk) +wk+1,i where {wi,k} is player i’s noise process.

3. Implications for Zero-Sum SettingsBefore presenting convergence analysis of the update in(2), we draw connections between Nash and Stackelbergequilibria in zero-sum games and discuss the relevance toapplications such as adversarial learning. To do so, we evalu-

ate the limiting behavior of the dynamics from a continuoustime viewpoint since the discrete time system closely ap-proximates this behavior for suitably selected learning rates.While we provide the intuition behind the results here, theformal proofs of the results are in Appendix E.

Let us first show that for zero-sum games, all stable criticalpoints of x = −ωS(x) are DSE and vice versa.

Proposition 2. In zero-sum games (f,−f) with f ∈Cq(X,R) for q ≥ 2, a joint strategy x ∈ X is a stablecritical point of x = −ωS(x) if and only if x is a DSE.Moreover, if f is generic, a point x is a stable critical pointof x = −ωS(x) if and only if it is a LSE.

The result follows from the structure of the Jacobian ofωS(x), which is lower block triangular with player 1 and2 as the leader and follower, respectively. Proposition 2implies that with appropriate stepsizes the update rule in (2)will only converge to Stackelberg equilibria and thus, unlikesimultaneous gradient descent, will not converge to spuriouslocally asymptotically stable points that lack game-theoreticmeaning (see, e.g., Mazumdar & Ratliff (2018)).

This previous result begs the question of which stable criticalpoints of the dynamics x = −ω(x) are DSE? The followinggives a partial answer to the question.

Proposition 3. In zero-sum games (f,−f) with f ∈Cq(X,R) for q ≥ 2, DNE are DSE. Moreover, if f is generic,LNE are LSE.

This result is obtained by examining the first Schur comple-ment of the Jacobian of the simultaneous gradient descentdynamics, and noting that it is exactly the second order suffi-cient conditions for the leader in the Stackelberg game. Thesecond statement follows from the fact that non-degenerateDNE are generic amongst LNE within the class of zero-sumgames (Mazumdar & Ratliff, 2019). In the zero-sum set-ting, the fact that Nash equilibria are a subset of Stackelbergequilibria for finite games is well-known (Basar & Olsder,1998). We extend this result locally to continuous actionspace games. Similar to our work and concurrently, Jin et al.(2019) show that LNE are local min-max solutions.

The preceding result indicates that recent works seekingDNE are also seeking DSE. This leaves the question of themeaning of stable points of x = −ω(x) which are not DNE.

Finding Meaning in Spurious Stable Fixed Points. Wefocus on the question of when stable fixed points of x =−ω(x) are DSE and not DNE. It was shown by Jin et al.(2019) that not all stable points of x = −ω(x) are local min-max or local max-min equilibria since one can construct afunction such that D2

1f(x) and −D22f(x) are both not posi-

tive definite but J(x) has positive eigenvalues. It appears tobe much harder to characterize when a stable critical pointof x = −ω(x) is not a DNE but is a DSE since it requires the

Page 5: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

220221222223224225226227228229230231232233234235236237238239240241242243244245246247248249250251252253254255256257258259260261262263264265266267268269270271272273274

Implicit Learning Dynamics in Stackelberg Games

15 10 5 0 5 10 15x1 (player 1's action)

15.0

12.5

10.0

7.5

5.0

2.5

0.0

2.5

5.0

x 2 (p

laye

r 2'

s ac

tion)

x0simgrad-stack

DSE/Non-NashDSE & DNE

Figure 1. Example demonstrating existence of DSE and DSE thatare not DNE: G = (f,−f) where f is defined in (3) with a =0.15, b = 0.25. There are two stable points of simultaneousgradient descent which are DSE, but not DNE.

follower’s individual Hessian to be positive definite. Indeed,it reduces to a fundamental problem in linear algebra inwhich the relationship between the eigenvalues of the sumof two matrices is largely unknown without assumptions onthe structure of the matrices (Knutson & Tao, 2001).

In Appendix F, we provide necessary and sufficient condi-tions for attractors at which the follower’s Hessian is positivedefinite to be DSE. Taking intuition from the expressionS1(J(x)) = D2

1f(x) − D21f(x)>(D22f(x))−1D21f(x),

the conditions are derived from relating spec(D21f) to

spec(D22f) viaD12f . To illustrate this fact, consider the fol-

lowing example in which stable points are DSE and not DNE—meaning points x ∈ X at which D2

1f(x) ≯ 0, −D22f(x) >

0 and spec(−J(x∗)) ⊂ C◦− and S1(J(x)) > 0.

Example: Non-Nash Attractors are Stackelberg. Con-sider the zero-sum game defined by

f(x) = −e−0.01(x21+x

22)((ax21 +x2)2 + (bx22 +x1)2). (3)

Let player 1 be the leader who aims to minimize f with re-spect to x1 taking into consideration that player 2 (follower)aims to minimize −f with respect to x2. In Fig. 1, we showthe trajectories for different initializations for this game; itcan be seen that simultaneous gradient descent can lead tostable critical points which are DSE and not DNE. In fact, itis the case that all stable critical points with −D2

2f(x) > 0are DSE in games on R2 (see Corollary 1, Appendix F).

This example, along with Propositions 8 and 9 in Ap-pendix F, implies some stable fixed points of x = −ω(x)which are not DNE are in fact DSE. This is a meaningfulresult since recent works have proposed schemes to avoidstable critical points which are not DNE as they have beenthought to lack game-theoretic meaning (Adolphs et al.,2019; Mazumdar et al., 2019). Moreover, some recent em-pirical studies show a number of successful approaches totraining GANs do not converge to DNE, but rather to stablefixed points of the dynamics at which the follower is ata local optimum (Berard et al., 2019). This may suggestreaching DSE is desirable in GANs.

The ‘realizable’ assumption in the GAN literature says thediscriminator network is zero near an equilibrium parameterconfiguration (Nagarajan & Kolter, 2017). The assumptionimplies the Jacobian of x = −ω(x) is such that D2

1f(x) =0. Under this assumption, we show stable critical pointswhich are not DNE are DSE given −D2

2f(x) > 0.Proposition 4. Consider a zero-sum GAN satisfying therealizable assumption. Any stable critical point of x =−ω(x) at which−D2

2f(x) > 0 is a DSE and a stable criticalpoint of x = −ωS(x).

4. Convergence AnalysisIn this section, we provide convergence guarantees for boththe deterministic and stochastic settings. In the former, play-ers have oracle access to their gradients at each step while inthe latter, players are assumed to have an unbiased estimatorof the gradient appearing in their update rule. Proofs of thedeterministic results can be found in Appendix G and thestochastic results in Appendix H.

4.1. Deterministic Setting

Consider the deterministic Stackelberg update

xk+1 = xk − γ2ωSτ (xk)

where ωSτ (xk) is the m-dimensional vector with entriesτ−1(D1f1(xk) − D>12f2(xk)(D2

2f2(xk))−1D2f1(xk)) ∈Rm1 and D2f2(xk) ∈ Rm2 , and τ = γ2/γ1 is the“timescale” separation with γ2 > γ1. We refer to thesedynamics as the τ -Stackelberg update.

To get convergence guarantees, we apply well known resultsfrom discrete time dynamical systems. For a dynamical sys-tem xk+1 = F (xk), when the spectral radius ρ(DF (x∗)) ofthe Jacobian at fixed point is less than one, F is a contractionat x∗ so that x∗ is locally asymptotically stable (see Prop. 10,Appendix G). In particular, ρ(DF (x∗)) ≤ c < 1 im-plies that ‖DF‖ ≤ c + ε < 1 for ε > 0 on a neighbor-hood of x∗ (Ortega & Rheinboldt, 1970, 2.2.8). Hence,Prop. 10 implies that if ρ(DF (x∗)) = 1 − α < 1 forsome α, then there exists a ball Bp(x∗) of radius p > 0such that for any x0 ∈ Bp(x∗), and some constant K > 0,‖xk − x∗‖2 ≤ K(1− α/2)k‖x0 − x∗‖2 using ε = α/4.

For a zero-sum setting defined by cost function f ∈Cq(X,R) with q ≥ 2, recall that S1(J(x)) = D2

1f(x) −D21f(x)>(D2

2f(x))−1D21f(x) is the first Schur comple-ment of the Jacobian J(x). Let F = (f,−f) : X → R2

and DF be its Jacobian.Theorem 1 (Zero-Sum Convergence.). Consider a zero-sum game defined by f ∈ Cq(X,R) with q ≥ 2. For a DSE

x∗, with α = min{λmin(−D22f(x∗)), λmin( 1

τ S1(J(x∗)))}and β = ρ(DF (x∗)), the τ–Stackelberg update convergeslocally with a rate of O(1− (α/β)2).

Page 6: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

275276277278279280281282283284285286287288289290291292293294295296297298299300301302303304305306307308309310311312313314315316317318319320321322323324325326327328329

Implicit Learning Dynamics in Stackelberg Games

The proof leverages the structure of the Jacobian JSτ , whichis lower block diagonal, along with the above noted resultfrom dynamical systems theory. The key insight is that at agiven x, the spectrum of JSτ (x) is the union of the spectrumof τ−1S1(J(x)) and −D2

2f(x) for zero-sum settings. InAppendix G, we also provide a local convergence rate to anε–LSE given the agents initialize in a radius p ball Bp(x∗).Proposition 5. Consider a zero sum game defined byf ∈ Cq(X,R), q ≥ 2. Suppose that γ2 < 1/L wheremax{spec( 1

τ S1(J(x))) ∪ spec(−D22f(x))} ≤ L. Then, x

is a stable critical point of τ–Stackelberg update if and onlyif x is a DSE.

The proof follows from Theorem 1 and Proposition 2.Theorem 2 (Almost Sure Avoidance of Saddles). Considera general sum game defined by fi ∈ Cq(X,R), q ≥ 2 fori = 1, 2 and where, without loss of generality, player 1 is theleader. Suppose that ωSτ is L-Lipschitz with τ > 1 and thatγ2 < 1/L. The τ–Stackelberg learning dynamics convergeto saddle points of x = −ωSτ (x) on a set of measure zero.

In the zero-sum case, ωSτ being Lipschitz is equivalent tomax{spec( 1

τ S1(J(x))) ∪ spec(−D22f(x))} ≤ L. In this

case, using the structure of the Jacobian JSτ , we knowthat the eigenvalues are real, and hence the only admissibletypes of critical points are stable, unstable, or saddle points.The above results shows that τ -Stackelberg learning avoidssaddle points almost surely. Moreover, we know that theonly stable critical points are DSE and DSE are never saddlepoints for zero-sum games.

We now provide a convergence guarantee for deterministicgeneral-sum games. However, the convergence guarantee isno longer a global guarantee to the set of attractors of whichcritical points are DSE since there is potentially stable criticalpoints which are not DSE. This can be seen by examiningthe Jacobian which is no longer lower block triangular.

Given a stable DSE x∗, let Bp(x∗) be the largest ballof radius p > 0 contained in the region of attrac-tion on which 1

2 (J>Sτ + JSτ ) is positive definite. De-fine α = minx∈Bp(x∗) λ

2min( 1

2 (J>Sτ + JSτ )) and β =

maxx∈Bp(x∗) λmax(JSτ (x)>JSτ (x)).Theorem 3. Consider a general sum game (f1, f2) withfi ∈ Cq(X,R), q ≥ 2 for i = 1, 2 and where, without lossof generality, player 1 is the leader. Suppose that x∗ ∈ Xis a stable DSE and x0 ∈ Bp(x∗). Further, let γ2 =

√α/β

where α < β. Given ε > 0, the number of iterationsrequired by τ–Stackelberg learning to obtain an ε–DSE isbounded by O(2β log (p/ε) /α).

4.2. Stochastic Setting

In the stochastic setting, players use updates of the form

xi,k+1 = xi,k − γi,k(ωS,i(xk) + wi,k+1) (4)

where γ1,k = o(γ2,k) and {wi,k+1} is a stochastic processfor each i = 1, 2. The results in this section assume thefollowing. The maps Df1 : Rm → Rm1 , D2f2 : Rm →Rm2 are Lipschitz, and ‖Df1‖ < ∞. For each i ∈ {1, 2},the learning rates satisfy

∑k γi,k =∞,

∑k γ

2i,k <∞.

The primary technical machinery we use in this section isstochastic approximation theory (Borkar, 2008) and toolsfrom dynamical systems. The convergence guarantees inthis section are analogous to that for deterministic learningbut asymptotic in nature. We now show the dynamics avoidsaddle points in the stochastic learning regime.

Theorem 4 (Almost Sure Avoidance of Saddles.). Considera game (f1, f2) with fi ∈ Cq(Rm1 × Rm2 ,R), q ≥ 2 fori = 1, 2 and where without loss of generality, player 1 isthe leader. Suppose that for each i = 1, 2, there exists aconstant bi > 0 such that E[(wi,t·)+|Fi,t] ≥ bi for everyunit vector v ∈ Rmi . Then, Stackelberg learning convergesto strict saddle points of the game on a set of measure zero.

We also give asymptotic convergence guarantees. In partic-ular, Theorem 7 in Appendix H.3 provides a global conver-gence guarantee in general-sum games to the set of stableattractors which contains the set of DSE under assumptionson the global asymptotic stability of attractors of the contin-uous time limiting singularly perturbed dynamical system.The set of attractors may contain critical points which arenot DSE. However, Corollary 5 in Appendix H.3 gives aglobal convergence guarantee in zero-sum games to theset of stable attractors of which the only critical points areDSE under identical assumptions using the structure of theJacobian for the dynamics.

Relaxing these assumptions, the following proposition pro-vides a local convergence result which ensures that samplepoints asymptotically converge to locally asymptotic trajec-tories of the continuous time limiting singularly perturbedsystem, and thus to stable DSE.

Proposition 6. Consider a general sum game (f1, f2) withfi ∈ Cq(X,R), q ≥ 2 for i = 1, 2 and where, withoutloss of generality, player 1 is the leader and γ1,k = o(γ2,k).Given a DSE x∗, let Bp(x∗) = Bp1(x∗1) × Bp2(x∗2) withp1, p2 > 0 on which det(D2

2f2(x)) 6= 0. Suppose x0 ∈Bp(x

∗). If the dynamics x2 = −D2f2(x) have a locallyasymptotically stable attractor r(x1) uniformly in x1 onBp2(x∗2) and the dynamics x1 = Df1(x1, r(x1)) have alocally asymptotically stable attractor on Bp1(x∗1), thenxk → x∗ almost surely.

Beyond these asymptotic guarantees, in Appendix H, weleverage results stochastic approximation to provide finite-time, high probability concentration bounds which guarana-tee that the players’ stochastic updates get ‘locked in’ to anε-ball around a DSE.

Page 7: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

330331332333334335336337338339340341342343344345346347348349350351352353354355356357358359360361362363364365366367368369370371372373374375376377378379380381382383384

Implicit Learning Dynamics in Stackelberg Games

0 5000 10000Iterations

10 10

10 6

10 2

102

(a) m = 3

0 5000 10000Iterations

10 10

10 6

10 2

102

(b) m = 9

0 5000 10000Iterations

10 10

10 6

10 2

102simgrad: ||12 (W + W )||2simgrad: || V V ||2stack: ||12 (W + W )||2stack: || V V ||2

(c) m = 25

V0, 0

W0,

0

(d) m = 3

V0, 0

W0,

0

(e) m = 9

V1, 1

W1,

1

(f) m = 25

Figure 2. Stackelberg learning more effectively estimates covariance Σ as compared to simgrad. The generator (leader) and dis-criminator (follower) optimize for V,W ∈ Rm×m, respectively. Using learning rates γ1 = γ2/2 = 0.015 and regularization η = m/5,errors are shown in (a)–(c). Observe the cycling of simgrad (blue) in (d)–(f) for learning trajectory plots of elements of W and V .

(a) Gen. (b) Dis.

0

1

2

J Real Eigenvalues

(c) J

0.0

0.2

0.4

Schur Complement Eigenvalues

(d) S1 ◦ J

0.000

0.025

0.050

D21f Eigenvalues

(e) D21f1

0

1

2

D22f Eigenvalues

(f) D22f2

(g) Gen. (h) Dis. (i) J

0.0

0.2

0.4

Schur Complement Eigenvalues

(j) S1 ◦ J

0.05

0.00

0.05D2

1f Eigenvalues

(k) D21f1

0

1

2

D22f Eigenvalues

(l) D22f2

Figure 3. 2-d MoG GAN: Convergence to a DSE that is not a DNE for simgrad (top) and Stackelberg learning dynamics (bottom).The performance is depicted in (a)–(b) and (g)–(h) for the generator (leader) and discriminator (follower), respectively. Positive definitenessof the Jacobian, its Schur complement and individual Hessians is demonstrated via the six smallest real eigenvalues and six largest realeigenvalues in (c)–(f) and (i)–(l). For both updates, the leader’s Hessian is non-positive while the Schur complement is positive.

5. ExperimentsWe present several illustrative experiments showing the roleof DSE in the optimization landscape of GANs and the em-pirical benefits of training GANs with Stackelberg learningcompared to simultaneous gradient descent (simgrad). Weprovide implementation details of the Stackelberg leader up-date, the techniques to compute the relevant eigenvalues, andhyperparameters in Appendix C and D. In Appendix D.3, weapply the Stackelberg learning to a DCGAN on the MNISTdataset, demonstrating that the update can be implementedfor large-scale problems.

Example 1: Learning a Covariance Matrix. We considera data generating process of x ∼ N (0,Σ), where the co-variance Σ is unknown and the objective is to learn it usinga Wasserstein GAN. The discriminator is configured to bethe set of quadratic functions defined as DW (x) = x>Wxand the generator is a linear function of random input noisez ∼ N (0, I) defined by GV (z) = V z. The matrices W ∈Rm×m and V ∈ Rm×m are the parameters of the discrimi-nator and the generator, respectively. The Wasserstein GANcost for the problem f(V,W ) =

∑mi=1

∑mj=1Wij(Σij −∑m

k=1 VikVjk). We consider the generator to be the leaderminimizing f(V,W ). The discriminator is the followerand it minimizes a regularized cost function defined by

−f(V,W ) + η2 Tr(W>W ), where η ≥ 0 is a tunable regu-

larization parameter. The game is formally defined by thecosts (f1, f2) = (f(V,W ),−f(V,W ) + η

2 Tr(W>W )),where player 1 is the leader and player 2 is the follower. Inequilibrium, the generator picks V ∗ such that V ∗(V ∗)> =Σ and the discriminator selects W ∗ = 0. Further details aregiven in Appendix C from Daskalakis et al. (2018).

We compare the deterministic gradient update for Stack-elberg learning dynamics and simgrad, and analyze thedistance from equilibrium as a function of time. Weplot ‖Σ − V V >‖2 for the generator’s performance and‖ 12 (W + W>)‖2 for the discriminator’s performance inFig. 2 for varying dimensions m with learning rates γ1 =γ2/2 = 0.015 and fixed regularization terms η = m/5.We observe that Stackelberg learning converges to an equi-librium in fewer iterations than simgrad. For zero-sumgames, our theory provides reasoning for this behavior sinceat any critical point the eigenvalues of the game Jacobianare purely real. This is in contrast to simgrad, whose Ja-cobian can admit complex eigenvalues, known to causerotational forces in the dynamics. While there may be imag-inary eigenvalues in the Jacobian of Stackelberg dynamics ingeneral-sum games, this example demonstrates empiricallythat the Stackelberg dynamics rotate less.

Page 8: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

385386387388389390391392393394395396397398399400401402403404405406407408409410411412413414415416417418419420421422423424425426427428429430431432433434435436437438439

Implicit Learning Dynamics in Stackelberg Games

(a) Real (b) 8k (c) 20k (d) 40k (e) 60k (f) 8k (g) 20k (h) 40k (i) 60k

0

10

20J Real Eigenvalues

(j) J

0

20

40

60

Schur Complement Eigenvalues

(k) S1 ◦ J

0

10

20

D21f Eigenvalues

(l) D21f1 (m) D2

2f2

0

1

2

J Real Eigenvalues

(n) J

0

1

2

Schur Complement Eigenvalues

(o) S1 ◦ J

0.2

0.0

0.2

0.4D2

1f Eigenvalues

(p) D21f1

0

1

2

3D2

2f Eigenvalues

(q) D22f2

Figure 4. MoG: Improved Stability on the Learning Path. Convergence to DNE for simgrad in (b)–(e) and to non-Nash DSE forStackelberg learning in (f)–(i). To determine definiteness of the Jacobian, its Schur complement and individual Hessians, we show the sixsmallest and six largest real eigenvalues in (j)–(m) for simgrad and (n)–(q) for Stackelberg. The eigenvalues indicate simgrad convergedto a DNE, yet simgrad exhibits unstable learning (b–e), while Stackelberg learning is stable (f–i) and converges to a non-Nash DSE.

Example 2: Learning a Mixture of Gaussians. We traina GAN to learn a mixture of Gaussian distribution, wherethe generator is the leader and the discriminator is the fol-lower. The generator and discriminator networks have twoand one hidden layers, respectively; each hidden layer has32 neurons. We train using a batch size of 256, a latentdimension of 16, and the default ADAM optimizer configu-ration in PyTorch version 1.2 with decaying learning rates.We ensure the follower’s Hessian to be well-conditioned bymodifying the leader’s update,

x+1 = x1 − γ1(D1f1(x) +Drη(x)>D2f1(x))

where Drη(x)> = −D21f2(x)>(D2

2f2(x) + ηI)−1

andη = 1 is the regularization parameter. We provide details onits derivation in Appendix B. We also employ regularizationin the follower’s implicit map to compute the eigenvalues ofthe Schur complement to determine whether a critical pointis in a neighborhood of a DSE for the zero-sum game.

Diamond configuration. We demonstrate that bothsimgrad and Stackelberg learning can converge to a neigh-borhood of a non-Nash DSE. This experiment uses the satu-rating GAN objective (Goodfellow et al., 2014). In Fig. 3a–3b and Fig. 3g–3h we show a sample of the generator and thediscriminator for simgrad and the Stackelberg dynamics af-ter 40,000 training batches. Each learning rule converges sothat the generator can create a distribution that is close to theground truth and the discriminator is nearly at the optimalprobability throughout the input space. In Fig. 3c–3f andFig. 3i–3l, we show eigenvalues from the game that presenta deeper view of the convergence behavior. We observethat simgrad appears be in a neighborhood of a DSE that is

not a DNE since the individual Hessians for the leader andfollower are indefinite and positive definite, respectively,and the Schur complement is positive definite. Moreover,the eigenvalues of the leader’s individual Hessian are nearlyzero, which reflects the realizable assumption (cf. Sec. 3).The Stackelberg learning dynamics also converge to DSE

of the zero-sum game which is not a DNE, evident fromthe Schur complement of J being positive. This exampledemonstrates that standard GAN training can converge tostable critical points that are DSE and not DNE which pro-duce good generator and discriminator performance. Thisindicates that it may not be necessary to look only for DNE.

Circle configuration. We demonstrate improved perfor-mance when using Stackelberg learning dynamics. We useReLU activation functions and the non-saturating objectiveand show the performance in Fig. 4 along the learning pathfor the simgrad and Stackelberg learning dynamics. Theformer cycles and performs poorly until the learning rateshave decayed enough to stabilize the training process. Thelatter converges quickly to a solution that nearly matchesthe ground truth distribution. In a similar fashion as in thecovariance example, the leader update is able to reduce rota-tions. We show the eigenvalues after training and see thatfor this configuration, simgrad converges to a DNE and theStackelberg dynamics converge again to a DSE that is not aDNE. This provides further evidence that DSE may be easierto reach, and can provide suitable performance.

Page 9: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

440441442443444445446447448449450451452453454455456457458459460461462463464465466467468469470471472473474475476477478479480481482483484485486487488489490491492493494

Implicit Learning Dynamics in Stackelberg Games

6. DiscussionWe study learning dynamics in Stackelberg games. Thisclass of games pertains to any application in which thereis an order of play. However, the problem has not beenextensively analyzed in the way the learning dynamics ofsimultaneous play games have been. Consequently, we areable to give novel convergence results and draw connectionsto existing work focused on learning Nash equilibria.

ReferencesAbraham, R., Marsden, J. E., and Ratiu, T. Manifolds,

Tensor Analysis, and Applications. Springer, 2nd edition,1988. doi: 10.1007/978-1-4612-1029-0.

Adolphs, L., Daneshmand, H., Lucchi, A., and Hofmann,T. Local saddle point optimization: A curvature exploita-tion approach. In International Conference on ArtificialIntelligence and Statistics, pp. 486–495, 2019.

Argyros, I. K. A generalization of ostrowski’s theorem onfixed points. Applied Mathematics Letters, 12:77–79,1999.

Balduzzi, D., Racaniere, S., Martens, J., Foerster, J., Tuyls,K., and Graepel, T. The mechanics of n-player differen-tiable games. In International Conference on MachineLearning, pp. 354–363, 2018.

Basar, T. and Olsder, G. J. Dynamic Noncooperative GameTheory. Society for Industrial and Applied Mathematics,2nd edition, 1998.

Benaım, M. Dynamics of stochastic approximation algo-rithms. In Seminaire de Probabilites XXXIII, pp. 1–68,1999.

Berard, H., Gidel, G., Almahairi, A., Vincent, P., andLacoste-Julien, S. A closer look at the optimization land-scapes of generative adversarial networks. arXiv preprintarxiv:1906.04848, 2019.

Berger, T., Giribet, J., Perıa, F. M., and Trunk, C. On a Classof Non-Hermitian Matrices with Positive Definite SchurComplements. arXiv preprint arxiv:1807.08591, 2018.

Bhatnagar, S., Prasad, H., and Prashanth, L. Stochasticrecursive algorithms for optimization: simultaneous per-turbation methods, volume 434. Springer, 2012.

Borkar, V. S. Stochastic Approximation: A DynamicalSystems Viewpoint. Cambridge University Press, 2008.

Borkar, V. S. and Pattathil, S. Concentration bounds fortwo time scale stochastic approximation. In AllertonConference on Communication, Control, and Computing,pp. 504–511. IEEE, 2018.

Broer, H. and Takens, F. Chapter 1 - preliminaries of dynam-ical systems theory. In Henk Broer, F. T. and Hasselblatt,B. (eds.), Handbook of Dynamical Systems, volume 3 ofHandbook of Dynamical Systems, pp. 1 – 42. ElsevierScience, 2010. doi: 10.1016/S1874-575X(10)00309-7.

Callier, F. and Desoer, C. Linear Systems Theory. Springer,1991.

Dai, B., Shaw, A., Li, L., Xiao, L., He, N., Liu, Z., Chen, J.,and Song, L. Sbeed: Convergent reinforcement learningwith nonlinear function approximation. In InternationalConference on Machine Learning, pp. 1125–1134, 2018.

Daneshmand, H., Kohler, J., Lucchi, A., and Hofmann, T.Escaping saddles with stochastic gradients. In Proceed-ings of the 35th International Conference on MachineLearning, volume 80 of Proceedings of Machine Learn-ing Research, pp. 1155–1164, 10–15 Jul 2018.

Danskin, J. M. The theory of max-min, with applications.SIAM Journal on Applied Mathematics, 14(4):641–664,1966.

Danskin, J. M. The Theory of Max-Min and its Applicationto Weapons Allocation Problems. Springer, 1967.

Daskalakis, C. and Panageas, I. The limit points of (opti-mistic) gradient descent in min-max optimization. arXivpreprint arxiv:1807.03907, 2018.

Daskalakis, C., Ilyas, A., Syrgkanis, V., and Zeng, H. Train-ing gans with optimism. In International Conference onLearning Representations, 2018.

Foerster, J., Chen, R. Y., Al-Shedivat, M., Whiteson, S.,Abbeel, P., and Mordatch, I. Learning with opponent-learning awareness. In International Conference on Au-tonomous Agents and MultiAgent Systems, pp. 122–130,2018.

Fudenberg, D., Drew, F., Levine, D. K., and Levine, D. K.The theory of learning in games, volume 2. MIT press,1998.

Gibson, C. G., Wirthmuller, K., du Plessis, A. A., and Looi-jenga, E. J. N. Topological stability of smooth mappings.In Lecture Notes in Mathematics, volume 552. Springer-Verlag, 1976. doi: 10.1007/BFb0095244.

Golubitsky, M. and Guillemin, V. Stable Mappings andTheir Singularities. Springer-Verlag, 1973. doi: 10.1007/978-1-4615-7904-5.

Goodfellow, I., Pouget-Abadie, J., Mirza, M., Xu, B.,Warde-Farley, D., Ozair, S., Courville, A., and Bengio,Y. Generative adversarial nets. In Advances in NeuralInformation Processing Systems, pp. 2672–2680, 2014.

Page 10: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

495496497498499500501502503504505506507508509510511512513514515516517518519520521522523524525526527528529530531532533534535536537538539540541542543544545546547548549

Implicit Learning Dynamics in Stackelberg Games

Hirsch, M. W. Differential topology. Springer New York,1976.

Horn, R. and Johnson, C. Topics in Matrix Analysis. Cam-bridge University Press, 2011.

Jin, C., Netrapalli, P., and Jordan, M. I. Minmax optimiza-tion: Stable limit points of gradient descent ascent arelocally optimal. arXiv preprint arXiv:1902.00618, 2019.

Kelley, J. General Topology. Van Nostrand Reinhold Com-pany, 1955.

Knutson, A. and Tao, T. Honeycombs and sums of hermitianmatrices. Notices of the American Mathematical Society,2001.

Kushner, H. J. and Yin, G. G. Stochastic approximationand recursive algorithms and applications, volume 35.Springer Science & Business Media, 2003.

Lee, J. Introduction to smooth manifolds. Springer, 2012.

Lee, J. D., Simchowitz, M., Jordan, M. I., and Recht, B.Gradient descent only converges to minimizers. In Pro-ceedings of the 29th Annual Conference on LearningTheory, volume 49, pp. 1246–1257, 2016.

Letcher, A., Foerster, J., Balduzzi, D., Rocktaschel, T., andWhiteson, S. Stable opponent shaping in differentiablegames. In International Conference on Learning Repre-sentations, 2019.

Maclaurin, D., Duvenaud, D., and Adams, R. Gradient-based hyperparameter optimization through reversiblelearning. In International Conference on Machine Learn-ing, pp. 2113–2122, 2015.

Madry, A., Makelov, A., Schmidt, L., Tsipras, D., andVladu, A. Towards deep learning models resistant toadversarial attacks. In International Conference on Learn-ing Representations, 2018.

Mazumdar, E. and Ratliff, L. J. On the convergence ofgradient-based learning in continuous games. arXivpreprint arXiv:1804.05464, 2018.

Mazumdar, E. and Ratliff, L. J. Local nash equilibria areisolated, strict local nash equilibria in ’almost all’ zero-sum continuous games. In IEEE Conference on Decisionand Control, 2019.

Mazumdar, E., Jordan, M., and Sastry, S. S. On findinglocal nash equilibria (and only local nash equilibria) inzero-sum games. arxiv:1901.00838, 2019.

Mertikopoulos, P., Lecouat, B., Zenati, H., Foo, C.-S., Chan-drasekhar, V., and Piliouras, G. Optimistic mirror descentin saddle-point problems: Going the extra (gradient) mile.

In International Conference on Learning Representations,2019.

Mescheder, L., Nowozin, S., and Geiger, A. The numericsof gans. In Advances in Neural Information ProcessingSystems, pp. 1825–1835, 2017.

Metz, L., Poole, B., Pfau, D., and Sohl-Dickstein, J. Un-rolled generative adversarial networks. In InternationalConference on Learning Representations, 2017.

Nagarajan, V. and Kolter, J. Z. Gradient descent gan opti-mization is locally stable. In Advances in Neural Infor-mation Processing Systems, pp. 5585–5595, 2017.

Ortega, J. M. and Rheinboldt, W. C. Iterative Solutionsto Nonlinear Equations in Several Variables. AcademicPress, 1970.

Panageas, I. and Piliouras, G. Gradient Descent Only Con-verges to Minimizers: Non-Isolated Critical Points andInvariant Regions. In Proceedings of teh 8th Innovationsin Theoretical Computer Science Conference, volume 67,pp. 2:1–2:12, 2017. doi: 10.4230/LIPIcs.ITCS.2017.2.

Pemantle, R. Nonconvergence to unstable points in urn mod-els and stochastic approximations. Annuals Probability,18(2):698–712, 04 1990.

Radford, A., Metz, L., and Chintala, S. Unsupervised rep-resentation learning with deep convolutional generativeadversarial networks. arXiv preprint arXiv:1511.06434,2015.

Ratliff, L. J., Burden, S. A., and Sastry, S. S. Genericityand structural stability of non-degenerate differential nashequilibria. In American Control Conference, pp. 3990–3995. IEEE, 2014.

Ratliff, L. J., Burden, S. A., and Sastry, S. S. On the Charac-terization of Local Nash Equilibria in Continuous Games.IEEE Transactions on Automatic Control, 61(8):2301–2307, 2016.

Sastry, S. S. Nonlinear Systems Theory. Springer, 1999.

Schafer, F. and Anandkumar, A. Competitive gradient de-scent. In Advances in Neural Information ProcessingSystems, pp. 7623–7633, 2019.

Shub, M. Global Stability of Dynamical Systems. Springer-Verlag, 1978.

Thoppe, G. and Borkar, V. A concentration bound forstochastic approximation via alekseev’s formula. Stochas-tic Systems, 9(1):1–26, 2019.

Tretter, C. Spectral Theory of Block Operator Matrices andApplications. Imperial College Press, 2008.

Page 11: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

550551552553554555556557558559560561562563564565566567568569570571572573574575576577578579580581582583584585586587588589590591592593594595596597598599600601602603604

Implicit Learning Dynamics in Stackelberg Games

Von Stackelberg, H. Market structure and equilibrium.Springer Science & Business Media, 2010.

Zaslavski, A. J. Necessary optimality conditions for bilevelminimization problems. Nonlinear Analysis: Theory,Methods & Applications, 75(3):1655–1678, 2012.

Zhang, C. and Lesser, V. Multi-agent learning with policyprediction. In AAAI Conference on Artificial Intelligence,2010.

Zhang, K., Yang, Z., and Basar, T. Multi-agent reinforce-ment learning: A selective overview of theories and algo-rithms. arXiv preprint arXiv:1911.10635, 2019.

Page 12: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

605606607608609610611612613614615616617618619620621622623624625626627628629630631632633634635636637638639640641642643644645646647648649650651652653654655656657658659

Implicit Learning Dynamics in Stackelberg Games

A. Guide to the AppendixAppendix B. Derivations for the regularized Stackelberg learning dynamics including necessary and sufficient conditions.

Appendix C. Details on numerically computing the Stackelberg update and Schur complement.

Appendix D. Additional details on numerical simulations for the GANs and bimatrix game, comparing the Stackelbergupdate with simgrad.

Appendix E. Proofs for the main results in Section 3.

Appendix F. Necessary and sufficient conditions for non-Nash attractors such that −D22f(x) > 0 to be differential

Stackelberg equilibria in zero-sum settings.

Appendix G. Convergence proofs of the results in Section 4.1, and additional results for the deterministic setting notcovered in Section 4.1.

Appendix H. Convergence proofs for the results in Section 4.2. In addition, this section includes a number of extensionsfor the stochastic setting not included in the main body. In particular, we provide convergence guarantees for the leaderassuming the follower plays an exact best response, and an extended two timescale analysis for the case where thefollower is performing individual gradient play. We also introduce finite-time, high-probability concentration boundswhich guarantee that the agents’ updates get locked-in to an ε-ball around a differential Stackelberg equilibrium.

Appendix I. Proofs of stability and genericity in zero-sum continuous Stackelberg games.

B. Regularizing the Follower’s Implicit MapThe derivative of the implicit function used in the leader’s update requires the follower’s Hessian to be an isomorphism. Inpractice, this may not always be true along the learning path. Consider the modified update

xk+1,1 = xk,1 − γ1(D1f1(xk)−D21f2(xk)>(D22f2(xk) + ηI)−1D2f1(xk))

xk+1,2 = xk,2 − γ2D2f2(xk),

in which we regularize the inverse of D22f2 term. This update can be derived from the following perspective. Suppose player

1 views player 2 as optimizing a linearized version of its cost with a regularization term which captures the leader’s lack ofconfidence in the local linearization holding globally:

arg miny

(y − x2,k)>D2f2(xk) +η

2‖y − x2,k‖2.

The first-order optimality conditions for this problem are

0 = D2f2(xk) + (y − xk,2)>D22f2(xk) + η(y − xk,2)

= D2f2(xk)−(ηI +D2

2f2(xk))xk,2 + (D2

2f2(xk) + ηI)y.

Hence, if the leader views the follower as updating along the gradient direction determined by these first order conditions,then the follower’s response map is given by

xk+1,2 = xk,2 −(D2

2f2(xk) + ηI)−1

D2f2(xk).

Ignoring higher order terms in the derivative of the response map, the approximate Stackelberg update is given by

xk+1,1 = xk,1 − γ1(D1f1(xk)−D21f2(xk)>(D2

2f2(xk) + ηI)−1

D2f1(xk))

xk+1,2 = xk,2 − γ2D2f2(xk).

In our GAN experiments, we use the regularized update since it is quite common for the discriminator’s Hessian to beill-conditioned if not degenerate.

Proposition 7 (Regularized Stackelberg: Sufficient Conditions). A point x∗ such that the first order conditions D1f1(x)−D21f2(x)>(D2

2f2(x) + ηI)−1D2f1(x) = 0 and D2f2(x) = 0 hold, and such that D(D1f1(x)−D21f2(x)>(D22f2(x) +

ηI)−1D2f1(x)) > 0 and D22f2(x) > 0 is a differential Stackelberg equilibrium with respect to the regularized dynamics.

Page 13: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

660661662663664665666667668669670671672673674675676677678679680681682683684685686687688689690691692693694695696697698699700701702703704705706707708709710711712713714

Implicit Learning Dynamics in Stackelberg Games

This result can be seen by examining first and second order sufficient conditions for the leader’s optimization problem giventhe regularized conjecture about the follower’s update, i.e.

arg minx1

{f1(x1, x2)| x2 ∈ arg min

yf2(x1, y) +

η

2‖y‖2

},

and for the problem follower is actually solving with its update arg minx2f2(x1, x2).

C. Computing the Stackelberg Update and Schur ComplementThe learning rule for the leader involves computing an inverse-Hessian-vector product for the D2

2f2(x) inverse termand Jacobian-vector product for the D12f2(x) term. These operations can be done efficiently in Python by utilizingJacobian-vector products in auto-differentiation libraries combined with the sparse.LinearOperator class in scipy.These objects can also be used to compute their eigenvalues, inverses, or the Schur complement of the game dynamicsusing the scipy.sparse.linalg package. We found that the conjugate gradient method cg can compute the regularizedinverse-Hessian-vector products for the leader update accurately with 5 iterations and a warm start.

The operators required for the leader update can be obtained by the following. Consider the Jacobian of the simultaneousgradient descent learning dynamics x = −ω(x) at a critical point for the general sum game (f1, f2):

J(x) =

[D2

1f1(x) D12f1(x)D21f2(x) D2

2f2(x)

].

Its block components consist of four operators Dijfi(x) : Xj → Xi, i, j ∈ {1, 2} that can be computed using forward-mode or reverse-mode Jacobian-vector products. Instantiating these operators as a linear operator in scipy allows us tocompute the eigenvalues of the two player’s individual Hessians. Properties such as the real eigenvalues of a Hermitianmatrix or complex eigenvalues of a square matrix can be computed using eigsh or eigs respectively. Selecting to computethe smallest or largest k eigenvalues—sorted by either magnitude, real or imaginary values—allows one to examine thepositive-definiteness of the operators.

Operators can be combined to compute other operators relatively efficiently for large scale problems without requiring tocompute their full matrix representation. For an example, take the Schur complement of the Jacobian above at fixed networkparameters x ∈ X1 ×X2, D2

1(x)−D12f1(x)(D22f2)−1(x)D21f2(x). We create an operator S1(x) : X1 → X1 that maps

a vector v to p− q by performing the following four operations: u = D21f2(x)v, w = (D22f2)−1(x)u, q = D12f1(x)w,

and p = D21(x)v. Each of the operations can be computed using a single backward pass through the network except for

computing w, since the inverse-Hessian requires an iterative method which can be computationally expensive. It solvesthe linear equation D2

2f2(x)w = u and there are various available methods: we tested (bi)conjugate gradient methods,residual-based methods, or least-squares methods, and each of them provide varying amounts of error when comparedwith the exact solution. Particularly, when the Hessian is poorly conditioned, some methods may fail to converge. Moreinvestigation is required to determine which method is best suited for specific uses. For example, a fixed iteration methodwith warm start might be appropriate for computing the leader update online, while a residual-based method might bebetter for computing the the eigenvalues of the Schur complement. Specifically, for our mixture of gaussians and MNISTGANs, we found that computing the leader update using the conjugate gradient method with maximum of 5 iterations andwarm-start works well. We compared using the true Hessian for smaller scale problems and found the estimate to be withinnumerical precision.

D. Additional Numerical Simulations and DetailsThis section includes complete details on the training process and hyper-parameters selected in the mixture of Gaussian andMNIST experiments. We also present a numerical experiment of a parameterized bilinear game.

D.1. Mixture of Gaussians GAN

The underlying data distribution for the diamond experiment consists of Gaussian distributions with means given byµ = [1.5 sin(ω), 1.5 cos(ω)] for ω ∈ {kπ/2}3k=0 and each with covariance σ2I where σ2 = 0.15. Each sample of realdata given to the discriminator is selected uniformly at random from the set of Gaussian distributions. The underlyingdata distribution for the circle experiment consists of Gaussian distributions with means given by µ = [sin(ω), cos(ω)] for

Page 14: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

715716717718719720721722723724725726727728729730731732733734735736737738739740741742743744745746747748749750751752753754755756757758759760761762763764765766767768769

Implicit Learning Dynamics in Stackelberg Games

(a) (b) (c) (d)

Figure 5. We demonstrate Stackelberg learning on the MNIST dataset for digits for 0s and 1s in (a)-(b) and for all digits in (c)-(d).

ω ∈ {kπ/4}7k=0 and each with covariance σ2I where σ2 = 0.3. Each sample of real data given to the discriminator isselected uniformly at random from the set of Gaussian distributions.

We train the generator using latent vectors z ∈ R16 sampled from a standard normal distribution in each training batch. Thediscriminator is trained using input vectors x ∈ R2 sampled from the underlying distribution in each training batch. Thebatch size for each player in the game is 256. The network for the generator contains two hidden layers, each of whichcontain 32 neurons. The discriminator network consists of a single hidden layer with 32 neurons and it has a sigmoidactivation following the output layer. We let the activation function following the hidden layers be the Tanh function andthe ReLU function in the diamond and circle experiments, respectively. The initial learning rates for each player and foreach learning rule is 0.0001 and 0.0004 in the diamond and circle experiments, respectively. The objective for the gamein the diamond experiment is the GAN objective and in the circle experiment it is the non-saturating GAN objective. Weupdate the parameters for each player and in each experiment using the ADAM optimizer with the default parameters ofβ1 = 0.9, β2 = 0.999, and ε = 10−8. The learning rate for each player is decayed exponentially such that γi,k = γiν

ki .

We let ν1 = ν2 = 1− 10−7 for simultaneous gradient descent and ν1 = 1− 10−5 and ν1 = 1− 10−7 for the Stackelbergupdate. We regularize the implicit map of the follower as detailed in Appendix B using the parameter η = 1.

D.2. MNIST GAN

To demonstrate that the Stackelberg learning dynamics can scale to high dimensional problems, we train a GAN on theMNIST dataset using the DCGAN architecture adapted to handle 28× 28 images. We train a GAN on an MNIST datasetconsisting of only the digits 0 and 1 from the training images and on an MNIST dataset containing the entire set of trainingimages. We train using a batch size of 256 and a latent dimension of 100 and use the ADAM optimizer with the defaultparameters for the DCGAN network. We regularize the implicit map of the follower as detailed in Appendix B using theparameter η = 5000. If we view the regularization as a linear function of the number of parameters in the discriminator,then this selection of regularization is nearly equal to that from the mixture of Gaussian experiments.

We show the results in Fig. 5 after 2900 batches. For each dataset we show a sample of 16 digits to get a clear view of thegenerator performance and a sample of 256 digits to get a broader view of the generator output. The Stackelberg dynamicsare able to converge to a solution that generates realistic handwritten digits. The primary purpose of this example is to showthat the learning dynamics including second order information and an inverse is not an insurmountable problem for traininglarge scale networks with millions of parameters. We believe the tools we develop for our implementation can be helpfulto researchers working on GANs since a number of theoretical works on this topic require second order information tostrengthen the convergence guarantees.

The underlying data distribution for the MNIST experiments consists of digits 0 and 1 from the MNIST training datasetor each digit from the MNIST training dataset. We scale each image to the range [−1, 1]. Each sample of real data givento the discriminator is selected sequentially from a shuffled version of the dataset. The batch size for each player is 256.We train the generator using latent z ∈ R100 sampled from a standard normal distribution in each training batch. Thediscriminator is trained using input vectorized images x ∈ R28×28 sampled from the underlying distribution in each trainingbatch. We use the DCGAN architecture (Radford et al., 2015) for our generator and discriminator. Since DCGAN wasbuilt for 64× 64 images, we adapt it to handle 28× 28 images in the final layer. We follow the parameter choices fromthe DCGAN paper (Radford et al., 2015). This means we initialize the weights using a zero-centered centered Normaldistribution with standard deviation 0.02, optimize using ADAM with parameters β1 = 0.5, β2 = 0.999, and ε = 10−8, and

Page 15: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

770771772773774775776777778779780781782783784785786787788789790791792793794795796797798799800801802803804805806807808809810811812813814815816817818819820821822823824

Implicit Learning Dynamics in Stackelberg Games

(a) simgrad (γ1 = γ2) (b) simgrad (γ1 = γ2/10) (c) Stackelberg (γ1 = γ2) (d) Stackelberg (γ1 = γ2/10)

Figure 6. Parameterized bilinear game. Parameters: (a1, a2) = (2.5,−2.5) and (b1, b2) = (1,−1). (a)–(b): for simgrad, we observeconvergence to an ε-neighborhood of a DNE at (x∗1, x

∗2) = (−.4,−.4). (c)–(d): for the Stackelberg learning dynamics, we observe

convergence to an ε-neighborhood of a DSE at (x∗1, x∗2) = (−.4,−.16) which corresponds to player 1 choosing the action associated with

the top row with probability 0.5 and player 2 choosing the action associated with the first column with probability 0.77. The effects oftime-scale separation is visualized as the light colored horizontal path, showing a low gradient norm along player 2’s reaction curve.

set the initial learning rates to be 0.0002. The learning rate for each player is decayed exponentially such that γi,k = γiνki

and ν1 = 1− 10−5 and ν1 = 1− 10−7. We regularize the implicit map of the follower as detailed in Appendix B using theparameter η = 5000. If we view the regularization as a linear function of the number of parameters in the discriminator,then this selection of regularization is nearly equal to that from the mixture of Gaussian experiments.

D.3. Parameterized Bilinear Game

In the continuous game framework, player’s actions are continuous. To represent strategies with discrete actions, continuousprobability distributions can be employed as mixed strategies over the discrete actions. The gradient-based methodsdeveloped in this paper thus can be used to solve for equilibria in the parameterized strategy space. Consider the followinggame G = (f1, f2) with costs given by f1(x1, x2) = π(x1)>Aπ(x2) + η1

2 ‖x1‖22 and f2(x1, x2) = π(x1)>Bπ(x2) +

η22 ‖x2‖

22 where

A =

[1 00 1

], B =

[1/2 11 1/2

](5)

are the matrices representing the bimatrix game with player 1 as the row player and player 2 as the column player. Werepresent the mixed policy of two discrete actions with a sigmoid-based probability distribution on the simplex, π : R→ ∆1,given by

π(x) = (ea1x+b1 , ea2x+b2)/(ea1x+b1 + ea2x+b2)

where the parameters ai, bi, i = 1, 2 are constants that scale and shift the parameterization. This parameterization schemecan be extended to d + 1 actions using d variables. For two actions, we require that a1 and a2 have opposite signs. Weemploy a 2-norm regularization of each player’s individual action to regularize each agent towards the interior of the simplex.

The bimatrix game admits a unique mixed Nash equilibrium of (1/2, 1/2) for player 1 and (1/2, 1/2) for player 2. Ifthe game is played sequentially with the leader being player 1, the mixed Stackelberg equilibrium of the game is (π1, π2)with π1 = (1/2, 1/2) and any policy π2 in the simplex for the follower. At this strategy, the cost the leader incurs isindependent of the follower’s strategy. We refer to Basar & Olsder (1998, §3.6) for discussion on the mixed Stackelbergequilibrium of this bimatrix game. For the softmax parameterized policy class we consider, using (a1, a2) = (2.5,−2.5),(b1, b2) = (1,−1), π(−0.4) = (1/2, 1/2). That is, the parameter x = −0.4 corresponds to the policy (1/2, 1/2). On theother hand, if (a1, a2) = (2.5,−2.5), (b1, b2) = (0, 0), then π(0) = (1/2, 1/2).

We plot the the vector field ω and ωS and its norm, along with simulations of their discrete time dynamics in Figures 6 and 7.We use parameters (a1, a2) = (2.5,−2.5), and regularization η = 0.1. For the parameters (b1, b2), we explore two differentpairs: (b1, b2) = (1,−1) and (b1, b2) = (0, 0). The latter is such that the regularization term is penalizing for any deviationfrom the equilibrium parameter values, while the former is such that the regularization is penalizing for any deviation from(0, 0) while the equilibrium is at (0.4, y) for any y ∈ [0, 1].

The shading of the action space indicates the norm of the dynamics: darker has a larger norm. Different parameterization

Page 16: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

825826827828829830831832833834835836837838839840841842843844845846847848849850851852853854855856857858859860861862863864865866867868869870871872873874875876877878879

Implicit Learning Dynamics in Stackelberg Games

(a) simgrad (γ1 = γ2) (b) simgrad (γ1 = γ2/10) (c) Stackelberg (γ1 = γ2) (d) Stackelberg (γ1 = γ2/10)

Figure 7. Parameterized bilinear game. Parameters: (a1, a2) = (2.5,−2.5) and (b1, b2) = (0, 0). (a)–(b): for simgrad, we observeconvergence to an ε-neighborhood of a DNE at (x∗1, x

∗2) = (0, 0). (c)–(d): for the Stackelberg learning dynamics, we observe convergence

to an ε-neighborhood of a DSE at (x∗1, x∗2) = (0, 0) which corresponds to player 1 choosing the action associated with the top row with

probability 0.5 and player 2 choosing the action associated with the first column with probability 0.5. The effects of time-scale separationis visualized as the light colored horizontal path, showing a low gradient norm along player 2’s reaction curve. Note that due to the choiceof parameters (b1, b2), the regularization is penalizing for any deviation from the equilibrium at (1/2, 1/2) in the policy space.

constants or regularization weights will affect the outcome of the gradient-based learning.

The timescale separation improves the convergence properties of the Stackelberg learning dynamics as it encourages thedynamics to converge to the ‘ridge’ of the follower’s best-response curve. We observe the distinctly lighter path the shadedplots of Figure 6 (b) and (d), where the follower’s response curve runs horizontal to the plot. Comparing plots Stackelberglearning in figures (c) and (d) of 6 we observe that with timescale separation, the paths of the learning dynamics convergesfirst to the manifold on the ridge, then towards the stationary point along the manifold. Doing so prevents the trajectoryfrom being perturbed by the ‘cliffs’, visualized by the dark cusps with large gradient norm corresponding to area where thefollower’s Hessian is poorly conditioned. The importance of timescale separation in Stackelberg learning is emphasized inthis numerical simulation.

E. Proofs for Results on Connections and Implications for Zero-Sum SettingsThis appendix contains proofs for results given in Section 3. To be clear, we restate each result before providing the proof.

Proposition 2. In zero-sum games (f,−f) with f ∈ Cq(X,R) for q ≥ 2, a joint strategy x ∈ X is a stable critical pointof x = −ωS(x) if and only if x is a DSE. Moreover, if f is generic, a point x is a stable critical point of x = −ωS(x) if andonly if it is a LSE.

Proof. The Jacobian of the Stackelberg limiting dynamics x = −ωS(x) at a stable critical point is

JS(x) =

[D1(Df)(x) 0−D21f(x) −D2

2f(x)

]> 0. (6)

The structure of the Jacobian JS(x) follows from the fact that

D2(Df)(x) = D12f(x)−D12f(x)(D22f)−1(x)D2

2f(x) = 0.

The eigenvalues of a lower triangular block matrix are the union of the eigenvalues in each of the block diagonal components.This implies that if JS(x) > 0, then necessarily D1(Df)(x) > 0 and −D2

2f(x) > 0. Consequently, any stable criticalpoint of the Stackelberg limiting dynamics must be a differential Stackelberg equilibrium by definition.

Proposition 3. In zero-sum games (f,−f) with f ∈ Cq(X,R) for q ≥ 2, DNE are DSE. Moreover, if f is generic, LNE areLSE.

Proof. Consider an arbitrary sufficiently smooth zero-sum game (f,−f) on continuous strategy spaces. Suppose x is a

Page 17: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

880881882883884885886887888889890891892893894895896897898899900901902903904905906907908909910911912913914915916917918919920921922923924925926927928929930931932933934

Implicit Learning Dynamics in Stackelberg Games

stable differential Nash equilibrium so that by definition D21f(x) > 0, −D2

2f(x) > 0, and

J(x) =

[D2

1f(x) D12f(x)−D21f(x) −D2

2f(x)

]> 0.

Then, the Schur complement of J(x) is also positive definite:

D21f(x)−D21f(x)>(D2

2f(x))−1D21f(x) > 0

Hence, x is a differential Stackelberg equilibrium since the Schur complement of J is exactly the derivative D2f at criticalpoints and −D2

2f(x) > 0 since x is a differential Nash equilibrium.

Proposition 4. Consider a zero-sum GAN satisfying the realizable assumption. Any stable critical point of x = −ω(x) atwhich −D2

2f(x) > 0 is a DSE and a stable critical point of x = −ωS(x).

Proof. Consider an attractor x∗ of x = −ω(x) such that −D22f(x∗) ≥ 0. Note that the realizable assumption implies that

the Jacobian of ω is

J(x) =

[0 D12f(x)

−D21f(x) −D22f(x)

](see, e.g., (Nagarajan & Kolter, 2017)). Hence, since −D2

2f(x) > 0,

−D>21f(x∗)(D22f)−1(x∗)D21f(x∗) > 0.

Since x∗ is an attractor, D1f(x∗) = 0 and D2f(x∗) = 0 so that

Df(x∗) = D1f(x∗) +D2f(x∗)(D22f(x∗))−1D21f(x∗) = 0

Consequently, the necessary conditions for a local Stackelberg equilibrium are satisfied. Moreover, since both −D22f(x∗) >

0 and the Schur complement D21f(x∗) − D>21f(x∗)(D2

2f(x∗))−1D21f(x∗) > 0, the Jacobian of ωS has positive realeigenvalues so the point x∗ is stable.

F. When are Non-Nash Attractors of Simultaneous Gradient Play Stackelberg Equilibria?As alluded to in the main body, an interesting question is when non-Nash attractors of simultaneous gradient play—i.e.,critical points x∗ of x = −ω(x) at which spec(−J(x∗)) ⊂ C◦−—are differential Stackelberg equilibria. Attracting criticalpoints x∗ of the dynamics x = −ω(x) that are non-Nash equilibria are such that either D2

1f(x∗) or −D22f(x∗) are not

positive definite. Without loss of generality, considering player 1 to be the leader, an attractor of the Stackelberg dynamicsx = −ωS(x) requires both −D2

2f(x∗) and D21f(x∗)−D21f(x∗)>(D2

2f)−1(x∗)D21f(x∗) to be positive definite. Hence,if −D2

2f(x∗) is not positive definite at a non-Nash attractor of x = −ω(x), then x∗ will also not be an attractor ofx = −ωS(x). A central question is when can this occur zero-sum games? Towards answering this, let us consider thenon-Nash attractors2 with −D2

2f(x∗) > 0 and determine when the Schur complement above is positive definite, so that x∗

is an attractor of x = ωS(x).

In the following two propositions, we need some addition notion that is common across the two results. Let x1 ∈ Rm andx2 ∈ Rn. For a non-Nash attractor x∗, let spec(D2

1f(x∗)) = {µj , j ∈ {1, . . . ,m}} where

µ1 ≤ · · · ≤ µr < 0 ≤ µr+1 ≤ · · · ≤ µm,

and let spec(−D22f(x∗)) = {λi, i ∈ {1, . . . , n}} where λ1 ≥ · · · ≥ λn > 0, and define p = dim(ker(D2

1f(x∗))).

Proposition 8 (Necessary conditions). Consider a non-Nash attractor x∗ of the individual gradient dynamics x = −ω(x)such that−D2

2f(x∗) > 0. Given κ > 0 such that ‖D21f(x∗)‖ ≤ κ, ifD21f(x∗)−D21f(x∗)>(D2

2f(x∗))−1D21f(x∗) > 0,then r ≤ n and κ2λi + µi > 0 for all i ∈ {1, . . . , r − p}.

For a matrix W , let W † denote the conjugate transpose.

2We note that we could study an analogous setup in which we characterize when non-Nash attractors with D21f(x∗) > 0 are such that

−D22f(x∗) +D12f(x∗)>(D2

1f(x∗))−1D12f(x∗) > 0, thereby switching the roles of leader and follower.

Page 18: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

935936937938939940941942943944945946947948949950951952953954955956957958959960961962963964965966967968969970971972973974975976977978979980981982983984985986987988989

Implicit Learning Dynamics in Stackelberg Games

Proposition 9 (Sufficient conditions). Let x∗ be a stable non-Nash attractor of the individual gradient dynamics x = −ω(x)such that D2

1f(x∗) and −D22f(x∗) are Hermitian, and −D2

2f(x∗) > 0. Suppose that there exists a diagonal matrix (notnecessarily positive) Σ ∈ Cm×n with non-zero entries such that D12f(x∗) = W1ΣW †2 where W1 are the orthonormaleigenvectors of D2

1f(x∗) and W2 are orthonormal eigenvectors of −D22f(x∗). Given κ > 0 such that ‖D21f(x∗)‖ ≤ κ, if

r ≤ n and κ2λi + µi > 0 for each i ∈ {1, . . . , r − p}, then x∗ is a differential Stackelberg equilibrium and an attractor ofx = −ωS(x).

The proofs of the propositions follow from linear algebra results. Before diving into the proofs, we provide some commentary.Essentially, this says that if D2

1f(x∗) = W1MW †1 with W1W†1 = In×n and M diagonal, and −D2

2f(x∗) = W2ΛW †2with W2W

†2 = Im×m and Λ diagonal, then D12f(x∗) can be written as W1ΣW †2 for some diagonal matrix Σ ∈ Rn×m

(not necessarily positive). Note that since Σ does not necessarily have positive values, W1ΣW †2 is not the singular valuedecomposition of D12f(x∗). In turn, this means that each eigenvector of D2

1f(x∗) gets mapped onto a single eigenvector of−D2

2f(x∗) through the transformation D12f(x∗) which describes how player 1’s variation D1f(x) changes as a function ofplayer 2’s choice. With this structure for D12f(x∗), we can show that D2

1f(x∗)−D21f(x∗)>(D22f(x∗))−1D21f(x∗) > 0.

Note that if we remove the assumption that Σ has non-zero entries, then the remaining assumptions are still sufficient toguarantee that

D21f(x∗)−D21f(x∗)>(D2

2f)−1(x∗)D21f(x∗) ≥ 0.

This means that x∗ does not satisfy the conditions for a differential Stackelberg, however, the point does satisfy necessaryconditions for a local Stackelberg equilibrium and the point is a marginally stable attractor of the dynamics.

The results in this subsection follow from the theory of block operator matrices and indefinite linear algebra (Tretter, 2008).

The following lemma is a very well-known result in linear algebra and can be found in nearly any advanced linear algebratext such as (Horn & Johnson, 2011).

Lemma 1. Let W ∈ Cn×n be Hermitian with k positive eigenvalues (counted with multiplicities) and let U ∈ Cm×n. Then

λj(UWU†) ≤ ‖U‖2λj(W )

for j = 1, . . . ,min{k,m, rank(UWU†)}.

Let us define |M | = (MM>)1/2 for a matrix M . Recall also that for Propositions 8 and 9, we have definedspec(D2

1f(x∗)) = {µj , j ∈ {1, . . . ,m}} where µ1 ≤ · · · ≤ µr < 0 ≤ µr+1 ≤ · · · ≤ µm, and spec(−D22f(x∗)) =

{λi, i ∈ {1, . . . , n}} where λ1 ≥ · · · ≥ λn > 0, given an attractor x∗.

We can now use the above Lemma to prove Proposition 8. The proof follows the main arguments in the proof of Lemma 3.2in the work by Berger et al. (2018) with some minor changes due to the nature of our problem.

Proof of Proposition 8. Let x∗ be a stable attractor of x = −ωS(x) such that −D22f(x∗) > 0. For the sake of presentation,

define A = D21f(x∗), B = D12f(x∗), and C = D2

2f(x∗). Recall that x1 ∈ Rn and x2 ∈ Rm. Suppose that A −BC−1B> > 0.

Claim: r ≤ n is necessary. We argue by contradiction. Suppose not—i.e., assume that r > n. Note that if m < n,then this is not possible. In this case, we automatically satisfy that r ≤ n. Otherwise, r ≥ m > n. Let S1 =ker(B(−C−1 + |C−1|)B>) and consider the subspace S2 of Cm spanned by the all the eigenvectors of A corresponding tonon-positive eigenvalues. Note that

dimS1 = m− rank(B(−C−1 + |C−1|)B>) ≥ m− rank(−C−1 + |C−1|) = m− n

By assumption, we have that dimS2 = r so that, since r > n,

dimS1 + dimS2 ≥ (m− n) + r = m+ (r − n) > m.

Thus, S1 ∩ S2 6= {0}. Now, S1 = ker(B(−C−1 + |C−1|)B>). Hence, for any non-trivial vector v ∈ S1 ∩ S2,(BC−1B> −B|C−1|B>)v = 0 so that we have

〈(A−BC−1B>)v, v〉 = 〈Av, v〉 − 〈B|C−1|B>v, v〉 ≤ 0. (7)

Page 19: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

990991992993994995996997998999100010011002100310041005100610071008100910101011101210131014101510161017101810191020102110221023102410251026102710281029103010311032103310341035103610371038103910401041104210431044

Implicit Learning Dynamics in Stackelberg Games

Note that the inequality in (7) holds because the vector v is in the non-positive eigenspace of A and the second term isclearly non-positive. Thus, A−BC−1B> cannot be positive definite, which gives a contradiction so that r ≤ n.

Claim: κ2λi + µi > 0 is necessary. Let the maps λi(·) denote the eigenvalues of its argument arranged in non-increasingorder. Then, by the Weyl theorem for Hermitian matrices (Horn & Johnson, 2011), we have that

0 < λm(A−BC−1B>) ≤ λi(A) + λm−i+1(−BC−1B>), i ∈ {1, . . . ,m}.

We can now combine this inequality with Lemma 1. Indeed, we have that

0 < λi(A) + ‖B‖2λm−i+1(−C−1) < µm−i+1 + κ2λm−i+1, ∀ i ∈ {m− r + p+ 1, . . . ,m}

which gives the desired result.

Since we have shown both the necessary conditions, this concludes the proof.

Now, let us prove Proposition 9 which gives sufficient conditions for when a stable non-Nash attractor x∗ of x = −ω(x) is adifferential Stackelberg equilibrium. Then, combining this with Proposition 2, we have a sufficient condition under whichstable non-Nash attractors are in fact stable attractors of x = −ωS(x).

Proof of Proposition 9. Let x∗ be a stable non-Nash attractor of x = −ω(x) such that D21f(x∗) and D2

2f(x∗) > 0

are Hermitian. Since D2i f(x∗), i = 1, 2 are both Hermitian, let D2

1f(x∗) = W1MW †1 with W1W†1 = In×n and

M = diag(µ1, . . . , µm), and −D22f(x∗) = W2ΛW †2 with W2W

†2 = Im×m and Λ = diag(λ1, . . . , λn).

By assumption, there exists a diagonal matrix Σ ∈ Rm×n such that D12f(x∗) = W1ΣW †2 where W1 are the orthonormaleigenvectors of D2

1f(x∗) and W2 are orthonormal eigenvectors of −D22f(x∗). Then,

D21f(x∗)−D21f(x∗)>(D2

2f(x∗))−1D21f(x∗) = W1MW †1 +W1ΣW †2 (W2ΛW †2 )−1W2Σ†W †1

= W1(M + ΣΛ−1Σ†)W †1

Hence, to understand the eigenstructure of the Schur complement, we simply need to compare the all negative eigenvalues ofD2

1f(x∗) in increasing order with the most positive eigenvalues of −D22f(x∗) in decreasing order. Indeed, by assumption,

r ≤ n and κ2λi + µi > 0 for each i ∈ {1, . . . , r − p}. Thus,

D21f(x∗)−D21f(x∗)>(D2

2f(x∗))−1D21f(x∗) > 0

since it is a symmetric matrix. Combining this with the fact that−D22f(x∗) > 0, x∗ is a differential Stackelberg equilibrium.

Hence, by Proposition 2 it is an attractor of x = −ωS(x).

It is also worth noting that the fact that the eigenvalues of −J(x∗) are in the open-left-half complex plane is not used inproving this result. We believe that further investigation could lead to a less restrictive sufficient condition. Empirically, byrandomly generating the different block matrices, it is quite difficult to find examples such that J(x∗) has positive eigenvalues,−D2

2f(x∗) > 0, and the Schur complement D21f(x∗)−D21f(x∗)>(D2

2f)−1(x∗)D21f(x∗) is not positive definite. In thescalar case, the proof is straightforward; we suspsect that using the notion of quadratic numerical range (Tretter, 2008)—asuper set of the spectrum of a block operator matrix—along with the fact that the Jacobian of the simultaneous gradient playdynamics, −J , has its spectrum in the open left-half complex plane, we may be able to extend the scalar case to arbitrarydimensions.

We also note that the condition depends on conditions that are difficult to check a priori without knowledge of x∗. Certainclasses of games for which these conditions hold everywhere and not just at the equilibrium can be constructed. For instance,alternative conditions can be given: if the function f which defines the zero-sum game is such that it is concave in x2 andthere exists a K such that

D12f(x) = KD22f(x)

where supx ‖D12f(x)‖ ≤ κ <∞3 and K = W1ΣW †2 with Σ again a (not necessarily positive) diagonal matrix, then theresults of Proposition 9 hold. From a control point of view, one can think about the leader’s update as having a feedbackterm with the follower’s input. On the other hand, the results are useful for the synthesis of games, such as in reward shapingor incentive design, where the goal is to drive agents to particular desirable behavior.

3Functions such that derivative of f is Lipschitz will satisfy this condition.

Page 20: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

1045104610471048104910501051105210531054105510561057105810591060106110621063106410651066106710681069107010711072107310741075107610771078107910801081108210831084108510861087108810891090109110921093109410951096109710981099

Implicit Learning Dynamics in Stackelberg Games

F.1. equivalence in scalar games

Corollary 1. Suppose that xi ∈ R for i = 1, 2 and that player 1 is the leader and player 2 is the follower in a zero-sumsetting defined by f : R2 → R. Then, any non-Nash attractors of x = −ω(x) at which −D2

2f(x) > 0 are differentialStackelberg equilibria.

The proof follows directly from examining the necessary and sufficient conditions for when a 2 × 2 matrix is positivedefinite.

G. Deterministic Convergence ResultsConsider the deterministic Stackelberg update

xk+1,1 = xk,1 − γ1(D1f1(xk)−D>21f2(xk)(D22f2(xk))−1D2f1(xk)) (8)

xk+1,2 = xk,2 − γ2D2f2(xk) (9)

which is equivalent to

xk+1,1 = xk,1 −γ2τ

(D1f1(xk)−D>21f2(xk)(D22f2(xk))−1D2f1(xk)) (10)

xk+1,2 = xk,2 − γ2D2f2(xk) (11)

where τ = γ2/γ1 is the “timescale” separation. Then,

xk+1 = xk − γ2ωSτ (xk)

whereωSτ (xk) =

(τ−1(D1f1(xk)−D>21f2(xk)(D2

2f2(xk))−1D2f1(xk)), D2f2(xk)).

In the following subsections, we provide the proof of Proposition 1 which offers alternative sufficient conditions fordifferential Stackelberg relevant to our dynamics as well as convergence guarantees for the zero-sum and general-sumsettings, respectively.

G.1. Proof of Sufficient Conditions for Differential Stackelberg in General Sum Settings

Proposition 1 provides alternative sufficient conditions for differential Stackelberg in terms of the Jacobian of ωSτ which, inturn, defines the local behavior of the τ -Stackelberg learning dynamics.

Consider now a general-sum, continuous Stackelberg game defined by (f1, f2) with fi ∈ Cq(Rm,R), q ≥ 2. Note that

Df1(x1, r(x1)) = D1f1(x1, r(x1)) +Dr(x1)>D2f1(x1, r(x1))

and ωS(x1, x2) = (Df1(x1, x2), D2f2(x1, x2)) so that the hierarchical play Jacobian of the vector field ωS(x) is given by

JS(x) =

[D1(Df1(x1, x2)) D2(Df1(x1, x2))D21f2(x1, x2) D2

2f2(x1, x2)

](12)

Equilibrium conditions are such that ωS(x1, x2) = 0 and x2 = r(x1) where r(·) is implicitly defined by D2f2(x1, x2) = 0.The following result gives equivalent conditions for a differential Stackelberg equilibrium in terms of JS(x), whichdetermines the local behavior of the dynamical system defined by the learning dynamics.

We use the notation S1(·) to denote the Schur complement of (·) with respect to the bottom block diagonal matrix.Proposition 1. Consider a game (f1, f2) defined by fi ∈ Cq(X,R), i = 1, 2 with q ≥ 2 and player 1 (without loss ofgenerality) taken to be the leader. Let x∗ satisfyD2f2(x∗) = 0 andD2

2f2(x∗) > 0. ThenDf1(x∗) = 0 and S1(JS(x∗)) > 0if and only if x∗ is a DSE. Moreover, in zero-sum games, S1(JS(x)) = S1(J(x)).

Proof. The implicit function theorem implies that there exists neighborhoods U1 of x∗1 and W of D2f2(x∗1, x∗2) and a unique

Cq mapping r : U1 ×W → Rm2 on which D2f2(x1, r(x1)) = 0. The first Schur complement of JS is

S1(JS(x)) = D1(Df1(x1, x2))−D2(Df1(x1, x2))(D22f2(x1, x2))−1D21f2(x1, x2)

Page 21: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

1100110111021103110411051106110711081109111011111112111311141115111611171118111911201121112211231124112511261127112811291130113111321133113411351136113711381139114011411142114311441145114611471148114911501151115211531154

Implicit Learning Dynamics in Stackelberg Games

whereD1(Df1(x1, x2)) = D2

1f1(x1, x2) +D12f1(x1, x2)Dr(x1) +D2f1(x1, x2)D2r(x1)4

andD2(Df1(x1, x2)) = D12f1(x1, x2) +Dr(x1)>D2

2f1(x1, x2).

Now, we also have that the total derivative of Df1(x1, r(x1)) is given by

D(Df1(x1, r(x1))) = D21f1(x1, r(x1)) +D12f1(x1, r(x1))Dr(x1) +D2f1(x1, r(x1))>D2r(x1)

+ (D12f1(x1, r(x1)) +Dr(x1)>D22f1(x1, r(x1))Dr(x1)

Note also that by the implicit function theorem, Dr(x1) = −(D22f2(x1, x2))−1D21f2(x1, x2)|x2=r(x1). Hence, we have

that D(Df1(x1, r(x1))) = S1(JS(x))|x2=r(x1).

G.2. Almost Sure Avoidance of Saddles

Theorem 2 (Almost Sure Avoidance of Saddles). Consider a general sum game defined by fi ∈ Cq(X,R), q ≥ 2 fori = 1, 2 and where, without loss of generality, player 1 is the leader. Suppose that ωSτ is L-Lipschitz with τ > 1 and thatγ2 < 1/L. The τ–Stackelberg learning dynamics converge to saddle points of x = −ωSτ (x) on a set of measure zero.

Proof. To show this, we follow the arguments in Mazumdar & Ratliff (2018) with slight modifications, an argumentwhich also builds on similar results for single player optimization problems (Lee et al., 2016; Panageas & Piliouras,2017). In particular, we show that gSτ is a diffeomorphism, and then apply the center manifold theorem. First, recall thatτ = γ2/γ1 > 1. Let

gSτ (x) = x− γ2(1τ (D1f(x)−D21f

>(x)(D22f(x))−1D2f(x), D2f(x)

)We claim that gS : Rm → Rm is a diffeomorphism. If we can show that gSτ is invertible and a local diffeomorphism, thenthe claim follows.

Consider x 6= y and suppose gSτ (y) = gSτ (x) so that y − x = γ2(ωSτ (y) − ωSτ (x)). The assumptionsupx∈Rm ‖JSτ (x)‖2 ≤ L <∞ implies that ωSτ satisfies the Lipschitz condition on Rm. Hence, ‖ωSτ (y)− ωSτ (x)‖2 ≤L‖y − x‖2. Let Γ = γ2diag(τ−1I, I). Then, ‖x− y‖2 ≤ L‖Γ‖2‖y − x‖2 < ‖y − x‖2 since ‖Γ‖2 ≤ γ2 < 1/L.

Now, observe thatDgSτ = I−γ2JSτ (x). IfDgSτ is invertible, then the implicit function theorem (Lee, 2012, Theorem C.40)implies that gSτ is a local diffeomorphism. Hence, it suffices to show that γ2JSτ (x) does not have an eigenvalue of 1.Indeed, letting ρ(A) be the spectral radius of a matrix A, we know in general that ρ(A) ≤ ‖A‖ for any square matrix A andinduced operator norm ‖ · ‖ so that

ρ(γ2JSτ (x)) ≤ ‖γ2JSτ (x)‖2 ≤ γ2 supx∈Rm

‖JSτ (x)‖2 < γ2L < 1.

Of course, the spectral radius is the maximum absolute value of the eigenvalues, so that the above implies that all eigenvaluesof γ2JSτ (x) have absolute value less than 1. Since gSτ is injective by the preceding argument, its inverse is well-definedand since gSτ is a local diffeomorphism on Rm, it follows that g−1Sτ is smooth on Rm. Thus, gSτ is a diffeomorphism.

Consider all critical points to the game—i.e. Xc = {x ∈ X| ωSτ (x) = 0}. For each u ∈ Xc, let Bu, where u indexes thepoint, be the open ball derived from the center manifold theorem (Shub, 1978, Theorem III.7) and let B = ∪uBu. SinceX ⊆ Rm, Lindelof’s lemma (Kelley, 1955)—every open cover has a countable subcover—gives a countable subcover of B.That is, for a countable set of critical points {ui}∞i=1 with ui ∈ Xc, we have that B = ∪∞i=1Bui .

Starting from some point x0 ∈ X , if gradient-based learning converges to a strict saddle point, then there exists a t0 andindex i such that gtSτ (x0) ∈ Bpi for all t ≥ t0. Again, applying the center manifold theorem (Shub, 1978, Theorem III.7)and using that gSτ (X) ⊂ X—which we note is obviously true if X = Rm—we get that gtSτ (x0) ∈W cs

loc ∩X .

Using the fact that gSτ is invertible, we can iteratively construct the sequence of sets defined by W1(ui) = g−1Sτ (W csloc ∩X)

and Wk+1(ui) = g−1Sτ (Wk(ui) ∩ X). Then we have that x0 ∈ Wt(ui) for all t ≥ t0. The set X0 = ∪∞i=1 ∪∞t=0 Wt(ui)

4Note that D2r(x1) is denotes the appropriately dimensioned tensor so that D2f1(x1, x2)D2r(x1) is m1 ×m1 dimensional. Inparticular, D2f1(x1, x2) is a 1×m2 dimensional row vector and we take D2r(x1) to be a m2 ×m1 ×m1 dimensional tensor so thatD2f1(x1, x2)D2r(x1) means in Einstein summation notation (D2f1(x))i(D

2r(x1))ijk.

Page 22: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

1155115611571158115911601161116211631164116511661167116811691170117111721173117411751176117711781179118011811182118311841185118611871188118911901191119211931194119511961197119811991200120112021203120412051206120712081209

Implicit Learning Dynamics in Stackelberg Games

contains all the initial points in X such that gradient-based learning converges to a strict saddle. Since ui is a strictsaddle, I − γ2DωSτ (ui) has an eigenvalue greater than 1. This implies that the co-dimension of Eu is strictly less than m.(i.e. dim(W cs

loc) < m). Hence, W csloc ∩X has Lebesgue measure zero in Rm.

Using again that gSτ is a diffeomorphism, g−1Sτ ∈ C1 so that it is locally Lipschitz and locally Lipschitz maps are null set

preserving. Hence, Wk(ui) has measure zero for all k by induction so that X0 is a measure zero set since it is a countableunion of measure zero sets.

There is certainly the problem of spending infinite time at saddles.

G.3. Zero-Sum Setting

For zero-sum settings only, when τ →∞, simultaneous gradient play or τ–GDA with the timescaling, i.e.,

xk+1,1 = xk,1 −γ2τD1f(xk) (13)

xk+1,2 = xk,2 + γ2D2f(xk) (14)

converges to a local minimax equilibrium (Jin et al., 2019). Here, we consider a different update all together (i.e., not τ–GDAbut rather τ–Stackelberg) which is defined by the dynamics ωSτ given above.

In the zero-sum (deterministic) setting, the analysis is closely related to the ε–max oracle in (Jin et al., 2019). However, weprovide convergence guarantees using our dynamics as oppose to an exact ε-max oracle; that is, the follower is updating bytaking a step along the gradient direction as opposed to finding an exact ε local minima at each iteration.

Suppose γ2 is such that ρ(I − γ2JSτ ) < 1, then standard theory from numerical analysis for dynamical systems providesguarantees on local asymptotic convergence, and we can provide a finite time convergence guarantee to an ε–differentialStackelberg equilibrium. Indeed, letting

xk+1 = xk − γ2ωSτ (xk) (15)

be the τ–Stackelberg update, then the following well-known result from Ostrowski’s theorem on fixed points gives us exactlysuch convergence guarantees.

Proposition 10 (Ostrowski’s Theorem (Argyros, 1999)). Let x∗ be a fixed point for the discrete dynamical system xk+1 =F (xk). If the spectral radius of the Jacobian satisfies ρ(DF (x∗)) < 1, then F is a contraction at x∗ and hence, x∗ isasymptotically stable.

We note that ρ(DF (x∗)) < 1 implies there exists c > 0 such that ρ(DF (x∗)) ≤ c < 1. Hence, given any ε > 0, there is anorm and a c > 0 such that ‖DF‖ ≤ c+ ε < 1 on a neighborhood of x∗ (Ortega & Rheinboldt, 1970, 2.2.8). Thus, theproposition implies that if ρ(DF (x∗)) = 1− α < 1 for some α, then there exists a ball Bp(x∗) of radius p > 0 such thatfor any x0 ∈ Bp(x∗), and some constant K > 0, ‖xk − x∗‖2 ≤ K(1− α/2)k‖x0 − x∗‖2 using ε = α/4.

For a zero-sum setting defined by cost function f ∈ Cr(X1 ×X2,R) with r ≥ 2, let

S1(J(x)) = D21f(x)−D21f(x)>(D2

2f(x))−1D21f(x)

be the first Schur complement of the Jacobian J of ω(x) = (D1f(x), D2f(x)). Further, let F = (f,−f) : X1 ×X2 → R2

and let DF denote the Jacobian of the vector-valued function F .

Theorem 1 (Zero-Sum Convergence.). Consider a zero-sum game defined by f ∈ Cq(X,R) with q ≥ 2. For a DSE x∗,with α = min{λmin(−D2

2f(x∗)), λmin( 1τ S1(J(x∗)))} and β = ρ(DF (x∗)), the τ–Stackelberg update converges locally

with a rate of O(1− (α/β)2).

Proof. To show convergence, we simply need to show that ρ(I − γ2JSτ ) < 1 and apply Proposition 10. The structureof the Jacobian JSτ is lower-block triangular, with symmetric components along the diagonal—i.e., τ−1S1(J(x)) and−D2

2f(x). Due to this structure, we know that spec(JSτ (x)) = spec(τ−1S1(J(x))) ∪ spec(−D22f(x)). Hence, λmin(I −

γ2JSτ (x)) = 1 − γ2 min{λmin(τ−1S1(J(x))), λmin(−D22f(x))} = 1 − γ2α. Further, with a choice of γ2 = α/(2β2),

spec(I − γ2JSτ ) ∈ [0, 1− α2/(2β2)] since

λmax(τ−1S1(x)) ≤ ‖τ−1D21f(x∗)‖+ τ−1‖D12f(x∗)‖2‖(D2

2f(x∗)−1‖ ≤ τ−1β(1 + β/α) ≤ β(1 + β/α)

Page 23: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

1210121112121213121412151216121712181219122012211222122312241225122612271228122912301231123212331234123512361237123812391240124112421243124412451246124712481249125012511252125312541255125612571258125912601261126212631264

Implicit Learning Dynamics in Stackelberg Games

since τ = γ2/γ1 > 1 and λmax(−D22f(x∗)) ≤ β and ‖D12f(x∗)‖ ≤ β since ρ(DF (x∗)) = β by assumption. Hence,

applying Proposition 10 gives the convergence guarantee for τ–Stackelberg.

Recall that in the general sum case, we provided a local convergence rate on a given agents initialize in a ball of radius qaround the equilibrium. We can provide analogous convergence results for the zero-sum setting. It turns out that

S(x) = 12 (JSτ (x) + J>Sτ (x)) > 0 ⇐⇒ −τ−1D2

2f(x) > 0, D21f(x) > 0. (16)

Hence, an analogous theorem to Theorem 3 in the zero sum case holds for the subset of differential Stackelberg equilibriawhich are differential Nash equilibria.

For the zero-sum game defined by f ∈ Cq(X,R), given a stable differential Stackelberg equilibrium x∗, let Bp(x∗) be thelargest ball of radius p contained in the region of attraction on which S ≡ 1

2 (J>Sτ + JSτ ) is positive definite. Note that dueto (16), −D2

2f(x) > 0 and D21f(x) > 0 on Bp(x∗). Define constants

α = minx∈Bq(x∗) λ2min(S(x)) and β = maxx∈Bq(x∗) λmax(JSτ (x)>JSτ (x)).

Theorem 5. Consider a zero sum game (f,−f) defined by f ∈ Cq(X1 × X2,R) with q ≥ 2 for i = 1, 2 and where,without loss of generality, player 1 is the leader and player 2 is the follower. Suppose that x∗ ∈ X is a stable differentialStackelberg equilibrium and x0 ∈ Bp(x∗). Further, let γ2 =

√α/β where α < β. Given ε > 0, the number of iterations

required by τ–Stackelberg learning to obtain an ε–differential Stackelberg equilibrium is bounded by O(2β log (p/ε) /α).

The equivalence in (16) implies that x∗ is a differential Stackelberg equilibrium for which the conditions of Theorem 5 holdif and only if it is a differential Nash equilibrium.

Corollary 2. Under the conditions of Theorem 5, for any x0 ∈ Bp(x∗), the number of iterations required by τ -Stackelberglearning to obtain an ε-differential Nash equilibrium is O(2β log(q/ε)/α).

Proof of Theorem 5. It suffices to show that for the choice of γ2, ρ(I − γ2JSτ (x)) < 1 on Bp(x∗). Then, an inductiveargument can be made with the inductive hypothesis that xk ∈ Bp(x∗).

Since ωSτ (x∗) = 0, we have that

‖xk+1 − x∗‖2 = ‖xk − x∗ − γ2(ωSτ (xk)− ωSτ (x∗))‖2 ≤ supx∈Bp(x∗) ‖I − γ2JSτ (x)‖2‖xk − x∗‖2.

If supx∈Bp(x∗) ‖I − γΛJ(x)‖2 is less than one, where the norm is the operator 2–norm, then the dynamics are contractingas in Proposition 10. To simplify notation, we drop the explicit dependence on x. Then,

(I − γ2JSτ )>(I − γ2JSτ ) ≤ (1− 2γ2λmin(S) + γ22λmax(J>SτJSτ ))I ≤ (1− α/β)I. (17)

Using the above to bound supx∈Bp(x∗) ‖I − γ2JSτ (x)‖2, we have ‖xk+1 − x∗‖2 ≤ (1 − α/β)1/2‖xk − x∗‖2. Sinceα < β, (1− α/β) < e−α/β so that ‖xk+1 − x∗‖2 ≤ e−Tα/(2β)‖x0 − x∗‖2. This, in turn, implies that xk ∈ Bε(x∗) for allk ≥ T = d2βα log(p/ε)e.

Proposition 5. Consider a zero sum game defined by f ∈ Cq(X,R), q ≥ 2. Suppose that γ2 < 1/L wheremax{spec( 1

τ S1(J(x))) ∪ spec(−D22f(x))} ≤ L. Then, x is a stable critical point of τ–Stackelberg update if and

only if x is a DSE.

The proof of this proposition follows almost immediately from Theorem 1 and Proposition 2.

Theorem 6. Consider a zero sum game defined by f ∈ Cr(X,R), r ≥ 2. Suppose that τ > 1 and γ2 < 1/L wheremax{spec( 1

τ S1(J(x))) ∪ spec(−D22f(x))} ≤ L. The τ–Stackelberg learning dynamics converge to the set of DSE almost

surely.

Proof. Due to the structure of the Jacobian, we know that the at all critical points, the Jacobian of the τ -Stackelbergdynamics satisfies

spec(JSτ (x)) = spec(D21f(x)−D12f(x)>(D2

2f(x))−1D12f(x)) ∪ spec(−D22f(x))

Page 24: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

1265126612671268126912701271127212731274127512761277127812791280128112821283128412851286128712881289129012911292129312941295129612971298129913001301130213031304130513061307130813091310131113121313131413151316131713181319

Implicit Learning Dynamics in Stackelberg Games

Further, both D21f(x)−D12f(x)>(D2

2f(x))−1D12f(x) and −D22f(x) are Hermitian. Hence, all critical points are either

saddle points (i.e., linearly unstable critical points), unstable points (i.e., spec(−JSτ (x)) ⊂ C◦+), or stable differentialStackelberg equilibria. Unstable points are not attractors of the dynamics. Hence, what remains to be shown is that theset of initial conditions for which the Stackelberg learning rule converges to a strict saddle point is of measure zero. Theremainder of the proof follows the proof of Theorem 2.

Of course, as noted earlier escaping saddles can be onerous. However, results on escaping saddle points efficiently apply tothe τ -Stackelberg learning rule since spec(JSτ (x∗)) ⊂ R.

Corollary 3. Suppose that f ∈ Cq(X1 ×X2,R), q ≥ 2. Let S be the set of DSE. Suppose that

ξ < infx∈S{|λmax(D2

1f(x)−D12f(x)>(D22f(x))−1D12f(x)))|, |λmax(−D2

2f(x))|} <∞.

Then, if τ–Stackelberg learning converges to a differential Stackelberg equilibrium for all but a measure zero set of initialconditions, then γ2 < 2/ξ.

The above corollary provides a necessary condition on the stepsize of the follower for almost sure convergence to differentialStackelberg equilibria.

G.4. General Sum Setting

Analogous to the deterministic zero-sum setting, we can provide convergence guarantees for deterministic general sumsettings. However, the convergence guarantee is no longer a global guarantee in that there is potentially other types ofattractors in the general sum setting; this can be seen by the structure of the Jacobian which is no longer lower blocktriangular, nor is it symmetric.

Consider a general sum setting defined by fi ∈ Cq(X,R) with q ≥ 2 for i = 1, 2 and where, without loss of generality,player 1 is the leader and player 2 is the follower.

Unlike the zero-sum case, the structure of the Jacobian JSτ is not lower block triangular and hence, the convergence ratedepends more abstractly on the spectral structure of JSτ as opposed to the second-order sufficient conditions for a localStackelberg equilibrium. It is still an open question as to how the spectrum of JSτ relates to schur(JSτ ).

Given a stable differential Stackelberg equilibrium x∗, let Bp(x∗) be the largest ball of radius p > 0 contained in the regionof attraction on which S ≡ 1

2 (J>Sτ + JSτ ) is positive definite.

Define constants α = minx∈Bp(x∗) λmin

(S(x)>S(x)

)and β = maxx∈Bp(x∗) λmax(JSτ (x)TJSτ (x)).

Theorem 3. Consider a general sum game (f1, f2) with fi ∈ Cq(X,R), q ≥ 2 for i = 1, 2 and where, without loss ofgenerality, player 1 is the leader. Suppose that x∗ ∈ X is a stable DSE and x0 ∈ Bp(x∗). Further, let γ2 =

√α/β where

α < β. Given ε > 0, the number of iterations required by τ–Stackelberg learning to obtain an ε–DSE is bounded byO(2β log (p/ε) /α).

Proof. It suffices to show that for the choice of γ2, ρ(I − γ2JSτ (x)) < 1 on Bp(x∗). Then, an inductive argument can bemade with the inductive hypothesis that xk ∈ Bp(x∗). Since ωSτ (x∗) = 0, we have that

‖xk+1 − x∗‖2 = ‖xk − x∗ − γ2(ωSτ (xk)− ωSτ (x∗))‖2 ≤ supx∈Bp(x∗) ‖I − γ2JSτ (x)‖2‖xk − x∗‖2.

If supx∈Bp(x∗) ‖I − γΛJ(x)‖2 is less than one, where the norm is the operator 2–norm, then the dynamics are contractingas in Proposition 10. For notational convenience, we drop the explicit dependence on x. Then,

(I − γ2JSτ )>(I − γ2JSτ ) ≤ (1− 2γ2λmin(S) + γ22λmax(J>SτJSτ ))I ≤ (1− α/β)I. (18)

Using the above to bound supx∈Bp(x∗) ‖I − γ2JSτ (x)‖2, we have ‖xk+1 − x∗‖2 ≤ (1 − α/β)1/2‖xk − x∗‖2. Sinceα < β, (1− α/β) < e−α/β so that ‖xk+1 − x∗‖2 ≤ e−Tα/(2β)‖x0 − x∗‖2. This, in turn, implies that xk ∈ Bε(x∗) for allk ≥ T = d2βα log(p/ε)e.

Page 25: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

1320132113221323132413251326132713281329133013311332133313341335133613371338133913401341134213431344134513461347134813491350135113521353135413551356135713581359136013611362136313641365136613671368136913701371137213731374

Implicit Learning Dynamics in Stackelberg Games

H. Stochastic Convergence Results and Extended Analysis:In this supplementary section, we provide the formal proofs for the stochastic convergence results.

Before diving into the results, let us describe at a higher level the flavor of the results. Beyond the asymptotic convergenceguarantees above and the extended analysis in Appendix H.3, we can give concentration bounds which provided guaranteeson the probability of getting ’locked in’ to a neighborhood of a differential Stackelberg equilibrium using results in (Borkar& Pattathil, 2018; Thoppe & Borkar, 2019) along with classical dynamical systems tools; however, due to the length of theexposition, we leave this for another paper.

The analysis techniques we employ combine tools from dynamical systems theory with the theory of stochastic approximation.In particular, we leverage the limiting continuous time dynamical systems derived from (2) to characterize concentrationbounds for iterates or samples generated by (2). We note that the hierarchical learning update in (2) with timescale separationγ1,k = o(γ2,k) has a limiting dynamical system that takes the form of a singularly perturbed dynamical system given by

x1(t) = −Df1(x1(t), x2(t))x2(t) = −ν−1D2f2(x1(t), x2(t))

(19)

which, in the limit as ν → 0, approximates (4).

The limiting dynamical system has known convergence properties (asymptotic convergence in a region of attraction fora locally asymptotically stable attractor). Such convergence properties can be translated in some sense to the discretetime system by comparing pseudo-trajectories—in this case, linear interpolations between sample points of the updateprocess—generated by sample points of (2) and the limiting system flow for initializations containing the set of samplepoints of (2).

Let us now we review some mathematical preliminaries from dynamical systems theory.

H.1. Dynamical Systems Theory Primer

Let us first recall some results on stability. Given a sufficiently smooth function f ∈ Cq(X,R), a critical point x∗ of f issaid to be stable if for all t0 ≥ 0 and ε > 0, there exists δ(t0, ε) such that x0 ∈ Bδ(x∗) implies x(t) ∈ Bε(x∗), ∀t ≥ t0.Further, x∗ is said to be asymptotically stable if x∗ is additionally attractive—that is, for all t0 ≥ 0, there exists δ(t0) suchthat x0 ∈ Bδ(x∗) implies limt→∞ ‖x(t) − x∗‖ = 0. A critical point is said to be non-degenerate if the determinant ofthe Jacobian of the dynamics at the critical point is non-zero. For a non-degenerate critical point, the Hartman-Grobmantheorem (Sastry, 1999) enables us to check the eigenvalues of the Jacobian to determine asymptotic stability. In particular, ata non-degenerate critical point, if the eigenvalues of the Jacobian are in the open left-half complex plane, then the criticalpoint is asymptotically stable.

For the dynamics x = −ω(x), let J(x) denote the Jacobian of the vector field ω(x). Similarly, for the dynamics x = −ωS(x),let JS(x) denote the Jacobian of the vector field ωS(x). Then, we say a differential Nash equilibrium of a continuous gamewith corresponding individual gradient vector field ω is stable if spec(−J(x)) ⊂ C◦− where spec(·) denotes the spectrum ofits argument and C◦− denotes the open left-half complex plane. Similarly, we say differential Stackelberg equilibrium isstable if spec(−JS(x)) ⊂ C◦−.

In the stochastic setting, we use chain invariant sets.

Definition 5. Given T > 0, δ > 0, if there exists an increasing sequence of times tj with t0 = 0 and tj+1− tj ≥ T for eachj and solutions ξj(t), t ∈ [tj , tj+1] of ξ = F (ξ) with initialization ξ(0) = ξ0 such that supt∈[tj ,tj+1] ‖ξ

j(t)− z(t)‖ < δfor some bounded, measurable z(·), the we call z a (T, δ)–perturbation.

Lemma 2 (Hirsch Lemma). Given ε > 0, T > 0, there exists δ > 0 such that for all δ ∈ (0, δ), every (T, δ)–perturbationof ξ = F (ξ) converges to an ε–neighborhood of the global attractor set for ξ = F (ξ).

H.2. Learning Stackelberg Solutions for the Leader: A Best Response Analysis

In this supplementary section, we provide convergence results for the leader given that the follower is playing a local bestresponse strategy at each iteration. We consider the stochastic setting in which the leader does not have oracle access to theirgradients, but do have an unbiased estimator. As an example, players could be performing policy gradient reinforcementlearning or alternative gradient-based learning schemes. Let dim(Xi) = mi for each i ∈ {1, 2} and m = m1 +m2.

Page 26: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

1375137613771378137913801381138213831384138513861387138813891390139113921393139413951396139713981399140014011402140314041405140614071408140914101411141214131414141514161417141814191420142114221423142414251426142714281429

Implicit Learning Dynamics in Stackelberg Games

Assumption 1. The following hold:

A1a. The maps Df1 : Rm → Rm1 , D2f2 : Rd → Rm2 are L1, L2 Lipschitz, and ‖Df1‖ ≤M1 <∞.A1b. For each i ∈ I, the learning rates satisfy

∑k γi,k =∞,

∑k γ

2i,k <∞.

A1c. The noise processes {wi,k} are zero mean, martingale difference sequences. That is, given the filtration Fk =σ(xs, w1,s, w2,s, s ≤ k), {wi,k}i∈I are conditionally independent, E[wi,k+1| Fk] = 0 a.s., and E[‖wi,k+1‖| Fk] ≤ci(1 + ‖xk‖) a.s. for some constants ci ≥ 0, i ∈ I.

Suppose that the leader (player 1) operates under the assumption that the follower (player 2) is playing a local optimumin each round. That is, given x1,k, x2,k+1 ∈ arg minx2 f2(x1,k, x2) for which D2f2(x1,k, x2) = 0 is a first-order localoptimality condition. If, for a given (x1, x2) ∈ X1 × X2, D2

2f2(x1, x2) is invertible and D2f2(x1, x2) = 0, then theimplicit function theorem implies that there exists neighborhoods U ⊂ X1 and V ⊂ X2 and a smooth map r : U → V suchthat r(x1) = x2.

Assumption 2. For every x1, x2 = −D2f2(x1, x2) has a globally asymptotically stable equilibrium r(x1) uniformly in x1and r : Rd1 → Rd2 is Lr–Lipschitz.

Consider the leader’s learning rule

x1,k+1 = x1,k − γ1,k(Df1(x1,k, x2,k) + w1,k+1) (20)

where x2,k is defined via the map r2 defined implicitly in a neighborhood of (x1,k, x2,k).

Proposition 11. Suppose that for each x ∈ X , D22f2 is non-degenerate and Assumption 1 holds for i = 1. Then, x1,k

converges almost surely to an (possibly sample path dependent) equilibrium point x∗1 which is a local Stackelberg solutionfor the leader. Moreover, if Assumption 1 holds for i = 2 and Assumption 2 holds, then x2,k → x∗2 = r(x∗1) so that (x∗1, x

∗2)

is a differential Stackelberg equilibrium.

Proof. This proof follows primarily from using known stochastic approximation results. The update rule in (20) is astochastic approximation of x1 = −Df1(x1, x2) and consequently is expected to track this ODE asymptotically. Themain idea behind the analysis is to construct a continuous interpolated trajectory x(t) for t ≥ 0 and show it asymptoticallyalmost surely approaches the solution set to the ODE. Under Assumptions 1–3, results from (Borkar, 2008, §2.1) implythat the sequence generated from (20) converges almost surely to a compact internally chain transitive set of x1 =−Df1(x1, x2). Furthermore, it can be observed that the only internally chain transitive invariant sets of the dynamics aredifferential Stackelberg equilibria since at any stable attractor of the dynamics D2f1(x1, r(x1)) > 0 and from assumptionD2

2f2(x1, r(x1)) > 0. Finally, from (Borkar, 2008, §2.2), we can conclude that the update from (20) almost surelyconverges to a possibly sample path dependent equilibrium point since the only internally chain transitive invariantsets for x1 = −Df1(x1, x2) are equilibria. The final claim that x2,k → r(x∗1) is guaranteed since r is Lipschitz andx1,k → x∗1.

The above result can be stated with a relaxed version of Assumption 2.

Corollary 4. Given a differential Stackelberg equilibrium x∗ = (x∗1, x∗2), let Bp(x∗) = Bp1(x∗1) × Bp2(x∗2) for some

p1, p2 > 0 on which D22f2 is non-degenerate. Suppose that Assumption 1 holds for i = 1 and that x1,0 ∈ Bp1(x∗1). Then,

x1,k converges almost surely to x∗1. Moreover, if Assumption 1 holds for i = 2, r(x1) is a locally asymptotically stableequilibrium uniformly in x1 on the ball Bp2(x∗2), and x2,0 ∈ Bp2(x∗2), then x2,k → x∗2 = r(x∗1).

The proof follows the same arguments as the proof of Proposition 11.

H.3. Learning Stackelberg Equilibria: Two-Timescale Analysis

Now, let us consider the case where the leader again operates under the assumption that the follower is playing (locally)optimally at each round so that the belief is D2f2(x1,k, x2,k) = 0, but the follower is actually performing the updatex2,k+1 = x2,k + g2(x1,k, x2,k) where g2 ≡ −γ2,kE[D2f2]. The learning dynamics in this setting are then

x1,k+1 = x1,k − γ1,k(Df1(xk) + w1,k+1) (21)x2,k+1 = x2,k − γ2,k(D2f2(xk) + w2,k+1) (22)

Page 27: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

1430143114321433143414351436143714381439144014411442144314441445144614471448144914501451145214531454145514561457145814591460146114621463146414651466146714681469147014711472147314741475147614771478147914801481148214831484

Implicit Learning Dynamics in Stackelberg Games

where Df1(x) = D1f1(x) +D2f1(x)Dr(x1). Suppose that γ1,k → 0 faster than γ2,k so that in the limit ν → 0, the aboveapproximates the singularly perturbed system defined by

x1(t) = −Df1(x1(t), x2(t))x2(t) = − 1

νD2f2(x1(t), x2(t))(23)

The learning rates can be seen as stepsizes in a discretization scheme for solving the above dynamics. The condition thatγ1,k = o(γ2,k) induces a timescale separation in which x2 evolves on a faster timescale than x1. That is, the fast transientplayer is the follower and the slow component is the leader since limk→∞ γ1,k/γ2,k = 0 implies that from the perspectiveof the follower, x1 appears quasi-static and from the perspective of the leader, x2 appears to have equilibriated, meaningD2f2(x1, x2) = 0 given x1. From this point of view, the learning dynamics (21)–(22) approximate the dynamics in thepreceding section. Moreover, stable attractors of the dynamics are such that the leader is at a local optima for f1, not justalong its coordinate axis but in both coordinates (x1, x2) constrained to the manifold r(x1); this is to make a distinctionbetween differential Nash equilibria in which agents are at local optima aligned with their individual coordinate axes.

A note on justification for timescale separation. The reason for this timescale separation is that the leader’s update isformulated using the reaction curve of the follower. In the gradient-based learning setting considered, the reaction curve canbe characterized by the set of critical points of f2(x1,k, ·) that have a local positive definite structure in the direction of x2,which is

{x2| D2f2(x1,k, x2) = 0, D22f2(x1,k, x2) > 0}.

This set can be characterized in terms of an implicit map r, defined by the leader’s belief that the follower is playinga best response to its choice at each iteration, which would imply D2f2(x1,k, x2,k) = 0. Moreover, under sufficientregularity conditions, the implicit mapping theorem (Lee, 2012) gives rise to the implicit map r : U → X2 : x1 7→ x2 ona neighborhood U ⊂ X1 of x1,k. We note that when r is defined uniformly in x1 on the domain for which convergenceis being assessed, the update in (2) is well-defined in the sense that the component of the derivative Df1 correspondingto the implicit dependence of the follower’s action on x1 via r is well-defined and locally consistent. In particular, for agiven point x = (x1, x2) such that D2f2(x1, x2) = 0 with D2

2f2(x) an isomorphism, the implicit function theorem impliesthere exists an open set U ⊂ X1 such that there exists a unique continuously differentiable function r : U → X2 such thatr(x1) = x2 and D2f2(x1, r(x1)) = 0 for all x1 ∈ U . Moreover,

Dr(x1) = −(D22f2)−1(x1, r(x1))D21f2(x1, r(x1))

on U . Thus, in the limit of the two-timescale setting, the leader sees the follower as having equilibriated (meaningD2f2 ≡ 0)so that

Df1(x1, x2) = D1f1(x1, x2) +D2f1(x1, x2)Dr(x1) (24)

The map r is an implicit representation of the follower’s reaction curve.

H.3.1. ASYMPTOTIC ALMOST SURE CONVERGENCE

The following two results are fairly classical results in stochastic approximation. They are leveraged here to makingconclusions about convergence to Stackelberg equilibria in hierarchical learning settings.

While we do not need the following assumption for all the results in this section, it is required for asymptotic convergence ofthe two-timescale process in (21)–(22).Assumption 3. The dynamics x1 = −Df1(x1, r(x1)) have a globally asymptotically stable equilibrium.

Under Assumption 1–3, and the assumption that γ1,k = o(γ2,k), classical results imply that the dynamics (21)–(22) convergealmost surely to a compact internally chain transitive set T of (23); see, e.g., (Borkar, 2008, §6.1-2), (Bhatnagar et al., 2012,§3.3). Furthermore, it is straightforward to see that stable differential Nash equilibria are internally chain transitive setssince they are stable attractors of the dynamics ξt = F (ξt) from (23).Remark 1. There are two important points to remark on at this juncture. First, the flow of the dynamics (23) is notnecessarily a gradient flow, meaning that the dynamics may admit non-equilibrium attractors such as periodic orbits. Thedynamics correspond to a gradient vector field if and only if D2(Df1) ≡ D12f2, meaning when the dynamics admit apotential function. Equilibria may also not be isolated unless the Jacobian of ωS , say JS , is non-degenerate at the points.Second, except in the case of zero-sum settings in which (f1, f2) = (f,−f), non-Stackelberg locally asymptotically stableequilibria are attractors. That is, convergence does not imply that the players have settled on a Stackelberg equilibrium, andthis can occur even if the dynamics admit a potential.

Page 28: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

1485148614871488148914901491149214931494149514961497149814991500150115021503150415051506150715081509151015111512151315141515151615171518151915201521152215231524152515261527152815291530153115321533153415351536153715381539

Implicit Learning Dynamics in Stackelberg Games

Let tk =∑k−1l=0 γ1,l be the (continuous) time accumulated after k samples of the slow component x1. Define ξ1,s(t) to be

the flow of x1 = −Df1(x1(t), r(x1(t))) starting at time s from intialization xs.

Proposition 12. Suppose that Assumptions 1 and 2 hold. Then, conditioning on the event {supk∑i ‖xi,k‖2 < ∞}, for

any integer K > 0, limk→∞ sup0≤h≤K ‖x1,k+h − ξ1,tk(tk+h)‖2 = 0 almost surely.

Proof. The proof follows standard arguments in stochastic approximation. We simply provide a sketch here to give someintuition. First, we show that conditioned on the event {supk

∑i ‖x1,k‖2 <∞}, (x1,k, x2,k)→ {(x1, r(x1))| x1 ∈ Rd1}

almost surely. Let ζk =γ1,kγ2,k

(Df1(xk) + w1,k+1). Hence the leader’s sample path is generated by x1,k+1 = x1,k − γ2,kζkwhich tracks x1 = 0 since ζk = o(1) so that it is asymptotically negligible. In particular, (x1,k, x2,k) tracks (x1 = 0, x2 =

−D2f2(x1, x2)). That is, on intervals [tj , tj+1] where tj =∑j−1l=0 γ2,l, the norm difference between interpolated trajectories

of the sample paths and the trajectories of (x1 = 0, x2 = −D2f2(x1, x2)) vanishes a.s. as k → ∞. Since the leader istracking x1 = 0, the follower can be viewed as tracking x2(t) = −D2f2(x1, x2(t)). Then applying Lemma 2 provided inAppendix H, limk→0 ‖x2,k − r(x1,k)‖ → 0 almost surely.

Now, by Assumption 1, Df1 is Lipschitz and bounded (in fact, independent of A1a., since Df1 ∈ Cq, q ≥ 1, it is locallyLipschtiz and, on the event {supk

∑i ‖xi,k‖2 < ∞}, it is bounded). In turn, it induces a continuous globally integrable

vector field, and therefore satisfies the assumptions of Benaım (1999, Prop. 4.1). Moreover, under Assumptions A1b. andA1c., the assumptions of Benaım (1999, Prop. 4.2) are satisfied, which gives the desired result.

Theorem 7. Under Assumption 3 and the assumptions of Proposition 12, (x1,k, x2,k) → (x∗1, r(x∗1)) almost surely

conditioned on the event {supk∑i ‖xi,k‖2 <∞}. That is, the learning dynamics (21)–(22) converge to stable attractors

of (23), the set of which includes the stable DSE.

Proof. Continuing with the conclusion of the proof of Proposition 12, on intervals [tk, tk+1] the norm difference betweeninterpolates of the sample path and the trajectories of x1 = −Df1(x1, r(x1)) vanish asymptotically; applying Lemma 2gives the result.

As with Corollary 4, we can relax Assumption 2 and 3 to local asymptotic stability assumptions.

Proposition 6. Consider a general sum game (f1, f2) with fi ∈ Cq(X,R), q ≥ 2 for i = 1, 2 and where, without loss ofgenerality, player 1 is the leader and γ1,k = o(γ2,k). Given a DSE x∗, let Bp(x∗) = Bp1(x∗1)×Bp2(x∗2) with p1, p2 > 0on which det(D2

2f2(x)) 6= 0. Suppose x0 ∈ Bp(x∗). If the dynamics x2 = −D2f2(x) have a locally asymptotically stableattractor r(x1) uniformly in x1 on Bp2(x∗2) and the dynamics x1 = Df1(x1, r(x1)) have a locally asymptotically stableattractor on Bp1(x∗1), then xk → x∗ almost surely.

The proof follows the same arguments as the proof of Proposition 12.

In the zero-sum setting, we note that the continuous time limiting system admits only saddle points, and stable and unstablecritical points; hence, we conjecture that avoidance of saddles should lead to a global convergence guarantee. In particular,if Stackelberg learning converges, then it converges to the set of DSE.

H.3.2. STOCHASTIC AVOIDANCE OF SADDLES

It is known that stochastic gradient descent in the single player setting with isotropic noise avoids saddles almostsurely (Daneshmand et al., 2018). It is also known that gradient play in non-convex games with stochastic gradientsand isotropic noise avoids saddle points of the game dynamics (Mazumdar & Ratliff, 2018). These results specialize tothe case of Stackelberg learning dynamics with stochastic gradients (i.e., unbiased estimators of the true gradient) andsufficiently rich noise.

The follow results from (Pemantle, 1990) implies saddle avoidance in Stackelberg learning. Consider a general stochasticapproximation framework xt+1 = xt + γt(h(xt)) + εt for h : X → TX with h ∈ C2 and where X ⊂ Rd and where TXdenotes the tangent space of X .

Theorem 8 (Theorem 1 (Pemantle, 1990)). Suppose γt is Ft–measurable and E[wt|Ft] = 0. Let the stochastic process{xt}t≥0 be defined as above for some sequence of random variables {εt} and {γt}. Let p ∈ X with h(p) = 0 and letW be a neighborhood of p. Assume that there are constants η ∈ (1/2, 1] and c1, c2, c3, c4 > 0 for which the followingconditions are satisfied whenever xt ∈ W and t sufficiently large: (i) p is a linear unstable critical point (i.e., a saddle

Page 29: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

1540154115421543154415451546154715481549155015511552155315541555155615571558155915601561156215631564156515661567156815691570157115721573157415751576157715781579158015811582158315841585158615871588158915901591159215931594

Implicit Learning Dynamics in Stackelberg Games

point), (ii) c1/tη ≤ γt ≤ c2/tη , (iii) E[(wt · v)+|Ft] ≥ c3/tη for every unit vector v ∈ TX , and (iv) ‖wt‖2 ≤ c4/tη . ThenP (xt → p) = 0.

The above classical result directly implies avoidance of saddles in Stackelberg learning.

Theorem 4 (Almost Sure Avoidance of Saddles.). Consider a game (f1, f2) with fi ∈ Cq(Rm1 × Rm2 ,R), q ≥ 2 fori = 1, 2 and where without loss of generality, player 1 is the leader. Suppose that for each i = 1, 2, there exists a constantbi > 0 such that E[(wi,t·)+|Fi,t] ≥ bi for every unit vector v ∈ Rmi . Then, Stackelberg learning converges to strict saddlepoints of the game on a set of measure zero.

The proof follows directly from showing that the Stackelberg learning update satisfies Theorem 8, provided the assumptionsof the theorem hold. The assumption that E[(wi,t · v)+|Fi,t] ≥ bi essentially requires the covariance of the noise to befull-rank, and is made to rule out degenerate cases where the noise forces the dynamics to stay on the stable manifold ofstrict saddle points. Indeed, this is exactly the goal of isotropic noise in stochastic gradient descent.

H.3.3. ZERO SUM CONVERGENCE

Leveraging the results in Section 3, the convergence guarantees are stronger since in zero-sum settings all attractors areStackelberg; this contrasts with the Nash equilibrium concept.

Corollary 5. Consider a zero-sum setting (f,−f) on X = X1 ×X2 = Rd. Under the assumptions of Theorem 4 andAssumptions 2–3, conditioning on the event {supk

∑i ‖xi,k‖2 < ∞}, the learning dynamics (21)–(22) asymptotically

converge to the set of stable attractors almost surely, of which the only critical points are differential Stackelberg equilibria.

The proof of this theorem follows the above analysis and invokes Proposition 2.

H.3.4. FINITE-TIME HIGH-PROBABILITY GUARANTEES

As a final note, which was remarked on previously, it is possible to convert the asymptotic results above, both global andlocal convergence guarantees, to high-probability concentration bounds using the recent results in (Borkar & Pattathil, 2018;Thoppe & Borkar, 2019). Generally, the results (and their proofs) in this subsection leverage classical results from stochasticapproximation (Benaım, 1999; Bhatnagar et al., 2012; Borkar, 2008; Kushner & Yin, 2003). We also note that it is possibleto extend the results to the N follower setting. Due to the length of the current exposition, we have left this out.

While asymptotic guarantees of the proceeding section are useful, high-probability finite-time guarantees can be leveragedmore directly in analysis and synthesis, e.g., of mechanisms to coordinate otherwise autonomous agents. In this section,we aim to provide concentration bounds for the purpose of deriving convergence rate and error bounds in support of thisobjective. The results in this section follow the very recent work by Borkar & Pattathil (2018). We highlight key differencesand, in particular, where the analysis may lead to insights relevant for learning in hierarchical decision problems betweennon-cooperative agents.

Consider a locally asymptotically stable differential Stackelberg equilibrium x∗ = (x∗1, r(x∗1)) ∈ X and let Bq0(x∗) be an

q0 > 0 radius ball around x∗ contained in the region of attraction. Stability implies that the Jacobian JS(x∗1, r(x∗1)) is positive

definite and by the converse Lyapunov theorem (Sastry, 1999, Chap. 5) there exists local Lyapunov functions for the dynamicsx1(t) = −Df1(x1(t), r(x1(t))) and for the dynamics x2(t) = −ν−1D2f2(x1, x2(t)), for each fixed x1. In particular,there exists a local Lyapunov function V ∈ C1(Rd1) with lim‖x1‖↑∞ V (x1) =∞, and 〈∇V (x1), Df1(x1, r(x1))〉 < 0 forx1 6= x∗1. For q > 0, let V q = {x ∈ dom(V ) : V (x) ≤ q}. Then, there is also q > q0 > 0 and ε0 > 0 such that for ε < ε0,

{x1 ∈ Rd1 | ‖x1 − x∗1‖ ≤ ε} ⊆ V q0 ⊂ Nε0(V q0) ⊆ V q ⊂ dom(V )

whereNε0(V q0) = {x ∈ Rd1 | ∃x′ ∈ V q0 s.t.‖x′ − x‖ ≤ ε0}.

An analogously defined V exists for the dynamics x2 for each fixed x1.

For now, fix n0 sufficiently large; we specify the values of n0 for which the theory holds before the statement of Theorem 9.Define the event En = {x2(t) ∈ V q ∀t ∈ [tn0

, tn]} where

x2(t) = x2,k + t−tkγ2,k

(x2,k+1 − x2,k)

Page 30: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

1595159615971598159916001601160216031604160516061607160816091610161116121613161416151616161716181619162016211622162316241625162616271628162916301631163216331634163516361637163816391640164116421643164416451646164716481649

Implicit Learning Dynamics in Stackelberg Games

are linear interpolates—i.e., asymptotic pseudo-trajectories—defined for t ∈ (tk, tk+1) with tk+1 = tk + γ2,k and t0 = 0.

The basic idea for constructing a concentration bound is to leverage Alekseev’s formula (Thm. ??) to bound the differencebetween the asymptotic pseudo-trajectories and the flow of the corresponding limiting differential equation on eachcontinuous time interval between each of the successive iterates k and k + 1 by sequences of constants that decayasymptotically. Then, a union bound is used over all time intervals after defined for n ≥ n0 in order to construct aconcentration bound. This is done first for the follower, showing that x2,k tracks the leader’s ’conjecture’ or belief r(x1,k)about the follower’s reaction, and then for the leader.

Follower’s Sequence Tracks the Leader’s Conjectured Sample Path. Let us start by showing that x2,k tracks the leader’s’conjecture’ or belief r(x1,k) about the follower’s reaction.

Following Borkar & Pattathil (2018), we can express the linear interpolates for any n ≥ n0 as

x2(tn+1) = x2(tn0)−

∑nk=n0

γ2,k(D2f2(xk) + w2,k+1)

where γ2,kD2f2(xk) =∫ tk+1

tkD2f2(x1,k, x2(tk)) ds and similarly for the w2,k+1 term. Adding and subtracting∫ tn+1

tn0

D2f2(x1(s), x2(s)) ds, Alekseev’s formula can be applied to get

x2(t) = x2(t) + Φ2(t, s, x1(tn0), x2(tn0

))(x2(tn0)− x2(tn0

)) +∫ ttn0

Φ2(t, s, x1(s), x2(s))ζ2(s) ds

where x1(t) ≡ x1 is constant (since x1 = 0), x2(t) = r(x1), and

ζ2(s) = −D2f2(x1(tk), x2(tk)) +D2f2(x1(s), x2(s)) + w2,k+1. (25)

In addition, for t ≥ s, Φ2(·) satisfies linear system

Φ2(t, s, x0) = J2(x1(t), x2(t))Φ2(t, s, x0),

with Φ2(t, s, x0) = I and x0 = (x1,0, x2,0) and where J2 the Jacobian of −D2f2(x1, ·).

Given that x∗ = (x∗1, r(x∗1)) is a stable differential Stackelberg equilibrium, J2(x∗) is positive definite. Hence, as in (Thoppe

& Borkar, 2019, Lem. 5.3), we can find M , κ2 > 0 such that for t ≥ s, x2,0 ∈ V q, ‖Φ2(t, s, x1,0, x2,0)‖ ≤Me−κ2(t−s);this result follows from standard results on stability of linear systems (see, e.g., Callier & Desoer (1991, §7.2, Thm. 33))along with a bound on ∫ t

s

∥∥D22f2(x1, x2(θ, s, x0))−D2

2f2(x∗)∥∥dθ

for x0 ∈ V q (see, e.g., Thoppe & Borkar (2019, Lem 5.2)).

Now, an interesting point worth making is that this analysis leads to a very nice result for the leader-follower setting.In particular, through the use of the auxiliary variable z, we can show that the follower’s sample path ‘tracks’ theleader’s conjectured sample path. Indeed, consider zk = r(x1,k), that is, where D2f2(x1,k, x2,k) = 0. Then, using aTaylor expansion of the implicitly defined conjecture r, we get zk+1 = zk + Dr(x1,k)(x1,k+1 − x1,k) + δk+1 where‖δk+1‖ ≤ Lr‖x1,k+1 − x1,k‖2 is the error from the remainder terms. Plugging in x1,k+1,

zk+1 = zk + γ2,k(−D2f2(x1,k, zk) + τkDr2(x1,k)(w1,k+1 −Df1(x1,k, x2,k)) + γ−12,kδk+1)

where τk = γ1,k/γ2,k. The terms after −D2f2 are o(1), and hence asymptotically negligible, so that this z sequence tracksdynamics as x2,k. We show that with high probability, they asymptotically contract, leading to the conclusion that thefollower’s dynamics track the leader’s conjecture.

Towards this end, we first bound the normed difference between x2,k and zk. Define constants

Hn0 = (‖x2(tn0 − x2(tn0)‖+ ‖z(tn0)− x2(tn0)‖),

andS2,n =

∑n−1k=n0

( ∫ tk+1

tkΦ2(tn, s, x1(tk), x2(tk))ds)w2,k+1.

Page 31: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

1650165116521653165416551656165716581659166016611662166316641665166616671668166916701671167216731674167516761677167816791680168116821683168416851686168716881689169016911692169316941695169616971698169917001701170217031704

Implicit Learning Dynamics in Stackelberg Games

Lemma 3. For any n ≥ n0, there exists K > 0 such that conditioned on En,

‖x2,n − zn‖ ≤ K(‖S2,n‖+ e−κ2(tn−tn0

)Hn0 + supn0≤k≤n−1 γ2,k + supn0≤k≤n−1 γ2,k‖w2,k+1‖2

+ supn0≤k≤n−1 τk + supn0≤k≤n−1 τk‖w1,k+1‖2)

where τk = γ1,k/γ2,k.

Using this bound, we can provide an asymptotic guarantee that x2,k tracks r(x1,k) and a high-probability guarantee that x2,kgets locked in to a ball around r(x∗1). Fix ε ∈ [0, 1) and let N be such that γ2,n ≤ ε/(8K), τn ≤ ε/(8K) for all n ≥ N .Let n0 ≥ N and with K as in Lemma 3, let T be such that e−κ2(tn−tn0 )Hn0

≤ ε/(8K) for all n ≥ n0 + T .

Theorem 9. Suppose that Assumptions 1, 2, and 3 hold and let γ1,k = o(γ2,k). Given a stable differential Stackelbergequilibrium x∗ = (x∗1, r(x

∗1)), the follower’s sample path generated by (22) with asymptotically track the leader’s conjecture

zk = r(x1,k) and, given ε ∈ [0, 1), will get ‘locked in’ to a ε–neighborhood with high probability conditioned on reachingBq0(x∗) by iteration n0. That is, letting n = n0 + T + 1, for some C1, C2, C3, C4 > 0,

P(‖x2,n − zn‖ ≤ ε,∀n ≥ n|x2,n0 , zn0 ∈ Bq0)

≥ 1−∑∞n=n0

C1e−C2

√ε/γ2,n −

∑∞n=n0

C2e−C2

√ε/τn −

∑∞n=n0

C3e−C4ε

2/βn . (26)

with βn = maxn0≤k≤n−1 e−κ2(

∑n−1i=k+1 γ2,i)γ2,k.

The key technique in proving the above theorem (which is done in detail in Borkar & Pattathil (2018) using results fromThoppe & Borkar (2019)), is taking a union bound of the errors over all the continuous time intervals defined for n ≥ n0.

Proofs of Theorem 9 and Lemma 3. Constructing the asymptotic pseudo-trajectories. Recall that the asymptotic pseudo-trajectories for any n ≥ n0 are given by

x2(tn+1) = x2(tn0)−∑nk=n0

γ2,k(D2f2(xk) + w2,k+1).

Note that ∑nk=n0

γ2,kD2f2(xk) =∑nk=n0

∫ tk+1

tkD2f2(x1,k, x2(tk)) ds

and similarly for the w2,k+1 term, due to the fact that tk+1− tk = γ2,k by construction. Hence, for s ∈ [tk, tk+1), the abovecan be rewritten as

x2(t) = x2(tn0) +

∫ ttn0−D2f2(x1(s), x2(s)) + ζ21(s) + ζ22(s) ds

where ζ21(s) = −D2f2(x1(tk), x2(tk))−D2f2(x1(s), x2(s)) and ζ22(s) = −w2,k+1. Note taht ζ2(s) in (25) is definedsuch that ζ2(s) = ζ21(s) + ζ22(s). Then, by the nonlinear variation of constants formula (Alekseev’s formula), we have

x2(t) = x2(t) + Φ2(t, s, x1(tn0), x2(tn0))(x2(tn0)− x2(tn0)) +∫ ttn0

Φ2(t, s, x1(s), x2(s))(ζ21(s) + ζ22(s)) ds

where x1(t) ≡ x1 is constant (since x1 = 0) and x2(t) = r(x1). Moreover, for t ≥ s, Φ2(·) satisfies linear system

Φ2(t, s, x0) = J2(x1(t), x2(t))Φ2(t, s, x0),

with initial data Φ2(t, s, x0) = I and x0 = (x1,0, x2,0) and where J2 the Jacobian of −D2f2(x1, ·). Given that x∗ =(x∗1, r(x

∗1)) is a stable differential Stackelberg equilibrium, J2(x∗) is positive definite. Hence, as in (Thoppe & Borkar, 2019,

Lem. 5.3), we can find M , κ2 > 0 such that for t ≥ s, x2,0 ∈ V r,

‖Φ2(t, s, x1,0, x2,0)‖ ≤Me−κ2(t−s).

This result follows from standard results on stability of linear systems (see, e.g., Callier & Desoer (1991, §7.2, Thm. 33))along with a bound on

∫ ts‖D2

2f2(x1, x2(θ, s, x0)) − D22f2(x∗)‖dθ for x0 ∈ V q (see, e.g., Thoppe & Borkar (2019,

Lem 5.2)).

Page 32: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

1705170617071708170917101711171217131714171517161717171817191720172117221723172417251726172717281729173017311732173317341735173617371738173917401741174217431744174517461747174817491750175117521753175417551756175717581759

Implicit Learning Dynamics in Stackelberg Games

Recall also that we have the auxiliary sequence representing the leader’s belief (i.e., assumed reponse model) which wederived using a Taylor expansion of the implicitly defined map r:

zk+1 = zk +Dr(x1,k)(x1,k+1 − x1,k) + δk+1 (27)

where δk+1 are the remainder terms which satisfy ‖δk+1‖ ≤ Lr‖x1,k+1 − x1,k‖2 by assumption. Plugging in x1,k+1,

zk+1 = zk + γ2,k(−D2f2(x1,k, zk) + τkDr(x1,k)(w1,k+1 −Df1(x1,k, x2,k)

)+ γ−12,kδk+1).

The terms after−D2f2 are o(1), and hence asymptotically negligible, so that this z sequence tracks dynamics as x2,k. Usingsimilar techniques as above, we can express linear interpolates of the leader’s assumed response model for the follower as

z(t) = z(tn0) +

∫ ttn0−D2f2(x1(s), z(s)) +

∑4j=1 ζ3j(s) ds

where we recall that τk = γ1,k/γ2,k and where the ζ3j’s are defined as follows:

ζ31(s) = −D2f2(x1(tk), z(tk)) +D2f2(x1(s), z(s))

ζ32(s) = τkDr(x1,k)w1,k+1

ζ33(s) = −τkDf1(x1,k, x2,k)Dr(x1,k)

ζ34(s) = 1γ2,k

δk+1

Once again, Alekseev’s formula can be applied where x2(t) = r(x1) and Φ2 is the same as in the application of Alekseev’sto x2,k. Indeed, this gives us

z(tn) = x2(tn) + Φ2(tn, tn0, x1(tn0

), z(tn0))(z(tn0

)− x2(tn0))

+∑n−1k=n0

∫ tk+1

tkΦ2(tn, s, x1(s), z(s))(−D2f2(x1(tk), z(tk)) +D2f2(x1(s), z(s))) ds (a)

+∑n−1k=n0

∫ tk+1

tkΦ2(tn, s, x1(s), z(s))τkDr(x1,k)w1,k+1 ds (b)

−∑n−1k=n0

∫ tk+1

tkΦ2(tn, s, x1(s), z(s))τkDf1(x1,k, x2,k)Dr(x1,k) ds (c)

+∑n−1k=n0

∫ tk+1

tkΦ2(tn, s, x1(s), z(s)) 1

γ2,kδk+1 ds (d)

Applying the linear system stability results, we get that

‖Φ2(tn, tn0 , x1(tn0), z(tn0))(z(tn0)− x2(tn0))‖ ≤ e−κ2(tn−tn0)‖z(tn0)− x2(tn0)‖. (28)

Each of the terms (a)–(d) can be bound as in Lemma III.1–5 in (Borkar & Pattathil, 2018). The bounds are fairlystraightforward using (28).

Tracking bounds and Proof of Lemma 3. Now that we have each of these asymptotic pseudo-trajectories, we can showthat with high probability, x2,k and zk asymptotically contract to one another, leading to the conclusion that the follower’sdynamics track the leader’s belief about the follower’s reaction.

Moreover, we can bound the difference between x2,k, using x2(t2,k) = x2,k, and the continuous flow x2(t) on each interval[t2,k, ti,k+1) where t2,k = tk. The normed-difference bound can then be leveraged to obtain concentration bounds bytaking a union bound across all continuous time intervals defined after sufficiently large n0 and conditioned on the eventsEn = {x2(t) ∈ V q ∀t ∈ [tn0

, tn]}.

Towards this end, define Hn0= (‖x2(tn0

− x2(tn0)‖+ ‖z(tn0

)− x2(tn0)‖), and

S2,n =∑n−1k=n0

(∫ tk+1

tkΦ2(tn, s, x1(tk), x2(tk))ds

)w2,k+1.

Applying Lemma 5.8 (Thoppe & Borkar, 2019), conditioned on En, we get there exists some constant K > 0 such that

‖x2(tn)− x2(tn)‖ ≤ ‖Φ2(tn, tn0, x1, x2(tn0

))(x2(tn0)− x2(tn0

))‖+K(‖S2,n‖

+ supn0≤k≤n−1 γ2,k + supn0≤k≤n−1 γ2,k‖w2,k+1‖2).

Page 33: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

1760176117621763176417651766176717681769177017711772177317741775177617771778177917801781178217831784178517861787178817891790179117921793179417951796179717981799180018011802180318041805180618071808180918101811181218131814

Implicit Learning Dynamics in Stackelberg Games

Using the bound on the linear system Φ2(·), this exactly leads to the bound

‖x2(tn)− x2(tn)‖ ≤ K(e−κ2(tn−tn0

)‖x2(tn0)− x2(tn0)‖

+ ‖S2,n‖+ supn0≤k≤n−1 γ2,k + supn0≤k≤n−1 γ2,k‖w2,k+1‖2)

Thus, leveraging Lemma III.1–5 (Thoppe & Borkar, 2019), we obtain the result of Lemma 3.

Proving Concentration Bounds in Theorem 9. Now, to obtain the concentration bounds, fix ε ∈ [0, 1) and let N besuch that γ2,n ≤ ε/(8K), τn ≤ ε/(8K) for all n ≥ N . Let n0 ≥ N and with K as in Lemma 3, let T be such thate−κ2(tn−tn0 )Hn0

≤ ε/(8K) for all n ≥ n0 + T .

Using Lemma 3 and Lemma 3.1 (Thoppe & Borkar, 2019),

P(‖x2,n − zn‖ ≤ ε,∀n ≥ n|x2,n0, zn0

∈ Bq0)

≥ 1− P(⋃∞n=n0

A1,n ∪⋃∞n=n0

A2,n ∪⋃∞n=n0

A3,n| x2,n0, zn0

∈ Bq0)

where

A1,n ={En, ‖S2,n‖ > ε

8K

}, A2,n =

{En, γ2,k‖w2,n+1‖2 > ε

8K

}, A3,n =

{En, τn‖w1,n+1‖2 > ε

8K

}.

Taking a union bound gives

P(‖x2,n − zn‖ ≤ ε, ∀n ≥ n|x2,n0 , zn0 ∈ Bq0) ≥1−∑∞n=n0

P(A1,n| x2,n0 , zn0 ∈ Bq0)

+∑∞n=n0

P(A2,n| x2,n0, zn0

∈ Bq0)

+∑∞n=n0

P(A3,n)| x2,n0, zn0

∈ Bq0).

Theorem 6.2 (Thoppe & Borkar, 2019), gives bounds∑∞n=n0

P(A2,n| x2,n0, zn0

∈ Bq0) ≤ K1

∑∞n=n0

exp(−K

2√ε√γ2,k

), (29)

and ∑∞n=n0

P(A3,n)| x2,n0, zn0

∈ Bq0) ≤ K1

∑∞n=n0

exp(−K

2√ε√τk

). (30)

By Theorem 6.3 (Thoppe & Borkar, 2019), we also have that∑∞n=n0

P(A1,n| x2,n0, zn0

∈ Bq0) ≤ K2

∑∞n=n0

exp(−K3ε

2

βn

)(31)

withβn = max

n0≤k≤n−1e−κ2(

∑n−1i=k+1 γ2,i)γ2,k

for some K1,K2,K3 > 0. This proves the result of Theorem 9 with C1 = K1, C2 = K2, C3 = K2, C4 = K3.

The above theorem can be restated to give a guarantee on getting locked-in to an ε-neighborhood of a stable differenitalStackelberg equilibria x∗ if the learning processes are initialized in Bq0(x∗).

Corollary 6. Fix ε ∈ [0, 1) and suppose that γ2,n ≤ ε/(8K) for all n ≥ 0. With K as in Lemma 3, let T be suchthat e−κ2(tn−t0)H0 ≤ ε/(8K) for all n ≥ T . Under the assumptions of Theorem 9, x2,k will will get ‘locked in’ to aε–neighborhood with high probability conditioned on x0 ∈ Bq0(x∗) where the high-probability bound is given in (26) withn0 = 0.

Leader’s Sample Path Gets Locked in to a Neighborhood of the Stackelberg Equilibrium. Given that the follower’saction x2,k tracks r(x1,k), we can also show that x1,k gets locked into an ε–neighborhood of x∗1 after a finite time with highprobability. First, a similar bound as in Lemma 3 can be constructed for x1,k.

Page 34: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

1815181618171818181918201821182218231824182518261827182818291830183118321833183418351836183718381839184018411842184318441845184618471848184918501851185218531854185518561857185818591860186118621863186418651866186718681869

Implicit Learning Dynamics in Stackelberg Games

Define the event En = {x1(t) ∈ V q ∀t ∈ [tn0, tn]} where for each t,

x1(t) = x1,k + t−tkγ1,k

(x1,k+1 − x1,k)

is a linear interpolates between the samples {x1,k}, tk+1 = tk + γ1,k, and t0 = 0. Then as above, Alekseev’s formula canagain be applied to get

x1(t) = x1(t, tn0, y(tn0

)) + Φ1(t, tn0, x1(tn0

))(x1(tn0)− x1(tn0

)) +∫ ttn0

Φ1(t, s, x1(s))ζ1(s) ds

where x1(t) ≡ x∗1,

ζ1(s) = Df1(x1,k, r(x1,k))−Df1(x1(s), r(x1(s))) +Df1(xk)−Df1(x1,k, r(x1,k)) + w1,k+1,

and Φ1 is the solution to a linear system with dynamics J1(x∗1, r(x∗1)), the Jacobian of −Df1(·, r(·)), and with initial data

Φ1(s, s, x1,0) = I . This linear system, as above, has bound ‖Φ1(t, s, x1,0)‖ ≤M1eκ1(t−1) for some M1, κ1 > 0. Define

S1,n =∑n−1k=n0

∫ tk+1

tkΦ1(tn, s, x1(tk))ds · w1,k+1.

Lemma 4. For any n ≥ n0, there exists K > 0 such that conditioned on En,

‖x1(tn)− x1(tn)‖ ≤K(‖S1,n‖+ supn0≤k≤n−1 ‖S2,k‖+ supn0≤k≤n−1 γ2,k + supn0≤k≤n−1 τk

+ supn0≤k≤n−1 γ2,k‖w2,k+1‖2 + supn0≤k≤n−1 τk‖w1,k+1‖2 + supn0≤k≤n−1 τkHn0

+ eκ1(tn−tn0)‖x1(tn0)− x1(tn0)‖

).

Using this lemma, we can get the desired guarantees on x1,k. Indeed, as above, fix ε ∈ (0, 1] and let N be such thatγ2,n ≤ ε/(8K) and τn ≤ ε/(8K), ∀ n ≥ N . Then, for any n0 ≥ N and K as in Lemma 3, let T be such that

e−κ2(tn−tn0)Hn0 ≤ ε/(8K), ∀ n ≥ n0 + T.

Moreover, with K as in Lemma 4, let

e−κ1(tn−tn0 )(‖x1(tn0)− x1(tn0

)‖ ≤ ε/(8K), ∀ n ≥ n0 + T.

Theorem 10. Suppose that Assumptions 1–3 hold and that γ1,k = o(γ2,k). Given a stable differential Stackelbergequilibrium x∗ and ε ∈ [0, 1), xk will get ‘locked in’ to a ε-neighborhood of x∗ with high probability conditioned reachingBq0(x∗) by iteration n0. That is, letting n = n0 + T + 1, for some constants Cj > 0, j ∈ {1, . . . , 6},

P(‖x1,n − x1(tn)‖ ≤ ε,∀n ≥ n|xn0, xn0

∈ Bq0)

≥ 1 +∑∞n=n0

C1e−C2√ε/√γ2,n −

∑∞n=n0

C1e−C2√ε/√τn

−∑∞n=n0

C3e−C4ε

2/βn −∑∞n=n0

C5e−C6ε

2/ηn (32)

with ηn = maxn0≤k≤n−1(e−κ1(

∑n−1i=k+1 γ1,i)γ1,k

).

The proof of this theorem is exactly analogous to that of Theorem 9.

Proof Sketch of Theorem 10 and Lemma 4. Using the linear interpolates x1(t) as above, Alekseev’s formula can again beapplied to get

x1(t) = x1(t, tn0, y(tn0

)) + Φ1(t, tn0, x1(tn0

))(x1(tn0)− x1(tn0

))

+∫ ttn0

Φ1(t, s, x1(s))(ζ11(s) + ζ12(s) + ζ13(s)) ds

where x1(t) ≡ x∗1 (again, since x1 = 0) and the following hold:

ζ11(s) = Df1(x1,k, r2(x1,k))−Df1(x1(s), r2(x1(s)))

ζ12(s) = Df1(xk)−Df1(x1,k, r(x1,k))

ζ13(s) = w1,k+1

Page 35: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

1870187118721873187418751876187718781879188018811882188318841885188618871888188918901891189218931894189518961897189818991900190119021903190419051906190719081909191019111912191319141915191619171918191919201921192219231924

Implicit Learning Dynamics in Stackelberg Games

Moreover, Φ1 is the solution to a linear system with dynamics J1(x∗1, r(x∗1)), the Jacobian of −Df1(·, r(·)), and with initial

data Φ1(s, s, x1,0) = I . This linear system, as above, has bound

‖Φ1(t, s, x1,0)‖ ≤M1eκ1(t−1)

for some M1, κ1 > 0.

Moreover, we can bound the difference between x1,k, using x1(t1,k) = x1,k, and the continuous flow x1(t) on each interval[t1,k, t1,k+1) where t1,k = tk. The normed-difference bound can then be leveraged to obtain concentration bounds bytaking a union bound across all continuous time intervals defined after sufficiently large n0 and conditioned on the eventsEn = {x2(t) ∈ V q ∀t ∈ [tn0

, tn]} and En = {x1(t) ∈ V q ∀t ∈ [tn0, tn]}. Indeed, following a similar argument as in

the proof of Theorem 9 and Lemma 3, we obtain the bound in Lemma 4. From here, the argument follows exactly theconcentration bound argument in the remainder of the proof of Theorem 9.

An analogous corollary to Corollary 6 can be stated for x1,k with n0 = 0; we do not present it to save space.

I. Stability and Genericity in Zero-Sum Continuous Stackelberg GamesIt is important to reiterate that the conditions that define or characterize a differential Stackelberg equilibrium are sufficientconditions for a local Stackelberg equilibrium; indeed, if a point x∗ is a differential Stackelberg equilibrium, then it is alocal Stackelberg equilibrium.

Necessary conditions for a local Stackelberg solution for the leader follow from necessary conditions in nonlinear optimiza-tion.

Proposition 13. Suppose that x∗ = (x∗1, x∗2) is a local Stackelberg equilibrium such that the follower (player 2) is at a

strict local minimum—i.e., D2f2(x∗1, x∗2) = 0 and D2

2f2(x∗1, x∗2) > 0. Then, Df1(x∗1, x

∗2) = 0 and D2f1(x∗1, x

∗2) ≥ 0

Proof. The proof is straightforward. Indeed, suppose that given x∗1, x∗2 is a strict local minimum for the follower. By theimplicit function theorem (Lee, 2012), there exists a Cq map r : x1 7→ x2 is defined on a neighborhood of x∗1 such thatr(x∗1) = x∗2 and Dr ≡ −(D2

2f2)−1 ◦D21f2. Hence, necessary conditions for the leader reduce to necessary conditions onthe problem minx1

f1(x1, r(x1)).

Proposition 14. In zero-sum q-differentiable continuous games, all differential Stackelberg equilibria are non-degenerate,and hence hyperbolic critical points of the vector field ωS(x).

Proof. Indeed, the Jacobian of the vector field ωS(x) is lower block triangular as shown in (12). Hence, det(JS(x)) 6= 0 ifand only if det(D1(Df(x))) 6= 0 and det(−D2

2f(x)) 6= 0. Since differential Stackelberg are such that D1(Df(x)) > 0and−D2

2f(x) > 0, the fact that all differential Stackelberg are non-degenerate follows immediately. Further, the lower blocktriangular structure implies that spec(JS(x)) = spec(S1(J(x))) ∪ spec(−D2

2f(x)). Hence, all differential Stackelbergequilibria are hyperbolic.

As with the case of simultaneous play and the corresponding equilibrium concept, the remainder of this section is dedicatedto showing that for a generic zero-sum q-differentiable game, all local Stackelberg equilibria of the game are differentialStackelberg equilibria, and further they are structurally stable. For a zero-sum q-differentiable game G = (f,−f), if welet DSE(G) be the differential Stackelberg equilibria, NDSE(G) the non-degenerate differential Stackelberg equilibria, andLSE(G) the local Stackelberg equilibria, then we know that NDSE(G) =DSE(G) ⊆ LSE(G). What we show is that for genericf ∈ Cq(X,R), the game G = (f,−f) is such that LSE(G) =DSE(G). In particular, we show that the set of zero-sumq-differentiable games admitting any local Stackelberg equilibria which are not non-degenerate differential Stackelberg is ofmeasure zero in Cq(Rm,R).

I.1. Mathematical Preliminaries

In this appendix, we provide some additional mathematical preliminaries; the interested reader should see standard referencesfor a more detailed introduction (Abraham et al., 1988; Lee, 2012).

Page 36: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

1925192619271928192919301931193219331934193519361937193819391940194119421943194419451946194719481949195019511952195319541955195619571958195919601961196219631964196519661967196819691970197119721973197419751976197719781979

Implicit Learning Dynamics in Stackelberg Games

A smooth manifold is a topological manifold with a smooth atlas. Euclidean space, as considered in this paper, is asmooth manifold. For a vector space E, we define the vector space of continuous (p + s)–multilinear maps T ps (E) =Lp+s(E∗, . . . , E∗, E, . . . , E;R) with s copies of E and p copes of E∗ and where E∗ denotes the dual. Elements of T ps (E)are tensors on E, and T ps (X) denotes the vector bundle of such tensors (Abraham et al., 1988, Definition 5.2.9).

Consider smooth manifolds X and Y of dimension nx and ny respectively. An k–jet from X to Y is an equivalence class[x, f, U ]k of triples (x, f, U) where U ⊂ X is an open set, x ∈ U , and f : U → Y is a Ck map. The equivalence relationsatisfies [x, f, U ]k = [y, g, V ]k if x = y and in some pair of charts adapted to f at x, f and g have the same derivativesup to order k. We use the notation [x, f, U ]k = jkf(x) to denote the k–jet of f at x. The set of all k–jets from X to Y isdenoted by Jk(X,Y ). The jet bundle Jk(X,Y ) is a smooth manifold (see (Hirsch, 1976, Chapter 2) for the construction).For each Ck map f : X → Y we define a map jkf : X → Jk(X,Y ) by x 7→ jkf(x) and refer to it as the k–jet extension.Definition 6. Let X , Y be smooth manifolds and f : X → Y be a smooth mapping. Let Z be a smooth submanifold ofY and u a point in X . Then f intersects Z transversally at u (denoted f t Z at u) if either f(u) /∈ Z or f(u) ∈ Z andTf(u)Y = Tf(u)Z + (f∗)u(TuX).

For 1 ≤ k < s ≤ ∞ consider the jet map jk : Cs(X,Y )→ Cs−k(X, Jk(X,Y )) and let Z ⊂ Jk(X,Y ) be a submanifold.Define ⋂

| s(X,Y ; jk, Z) = {h ∈ Cs(X,Y )| jkh t Z}. (33)

A subset of a topological space X is residual if it contains the intersection of countably many open–dense sets. We say aproperty is generic if the set of all points of X which possess this property is residual (Broer & Takens, 2010).Theorem 11. (Jet Transversality Theorem, Chap. 2 (Hirsch, 1976)). Let X , Y be C∞ manifolds without boundary, and letZ ⊂ Jk(X,Y ) be a C∞ submanifold. Suppose that 1 ≤ k < s ≤ ∞. Then,

⋂| s(X,Y ; jk, Z) is residual and thus dense in

Cs(X,Y ) endowed with the strong topology, and open if Z is closed.Proposition 15. (Chap. II.4, Proposition 4.2 (Golubitsky & Guillemin, 1973)). Let X,Y be smooth manifolds and Z ⊂ Ya submanifold. Suppose that dimX < codimZ. Let f : X → Y be smooth and suppose that f t Z. Then, f(X) ∩ Z = ∅.

I.2. Genericity

We show that local Stackelberg equilibria of zero-sum games are generically non-degenerate differential Stackelbergequilibria. Towards this end, we utilize the well-known fact that non-degeneracy of critical points is a generic property ofsufficiently smooth functions.Lemma 5 ((Broer & Takens, 2010, Chapter 1)). For Cq(Rm,R) functions with q ≥ 2 it is a generic property that all thecritical points are non-degenerate.

The above lemma implies that for a generic function f ∈ Cq(Rm,R), the Hessian

H(x) =

D21f(x) · · · D1mf(x)

.... . .

...Dm1f(x) · · · D2

mf(x)

is non-degenerate at critical points—that is, det(H(x)) 6= 0. For a generic function f , we know that det(D2

i f(x)) 6= 0(cf. Lemma 7). For such a function,

S1(J(x)) = D21f(x)−D>21f(x)(D2

2f(x))−1D21f(x).

Furthermore, we denote the Stackelberg game Jacobian by

JS(x) =

[D1(Df(x)) 0−D>12f(x) −D2

2f(x)

].

where JS(x) is obtained5 by taking the Jacobian of the vector field (Df(x),−D2f(x)). At critical points—i.e., x such that(Df(x),−D2f(x)) = 0—we note that

JS(x) =

[S1(J(x)) 0−D>12f(x) −D2

2f(x)

].

5Note that the structure is lower block triangular since D2(Df(x)) = D21f(x) + Dr(x1)>D22f(x) = D21f(x) −

D21f(x)(D22f(x))−1D2

2f(x) = 0.

Page 37: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

1980198119821983198419851986198719881989199019911992199319941995199619971998199920002001200220032004200520062007200820092010201120122013201420152016201720182019202020212022202320242025202620272028202920302031203220332034

Implicit Learning Dynamics in Stackelberg Games

Lemma 6. Consider f ∈ Cq(Rm,R), q ≥ 2 and the corresponding zero-sum game G = (f,−f). For any critical pointx ∈ Rm (i.e., x ∈ {x ∈ Rm| ωS(x) = 0}), det(H(x)) 6= 0 ⇐⇒ det(J(x)) 6= 0 and, if det(−D2

2f(x)) 6= 0, thendet(H(x)) 6= 0 ⇐⇒ det(JS(x)) 6= 0.

Proof. Consider a fixed x = (x1, x2) ∈ Rm1 × Rm2 . Note that H(x) is equal to J(x) with the last m2 rows scaled eachby −1. Hence, J(x) = PH(x) where P = blockdiag(Im1 ,−Im2) with each Imi the mi ×mi identity matrix, so thatdet(H(x)) = (−1)m2 det(J(x)), which in turn proves teh first equivalence. For the second equivalence, suppose thatdet(−D2

2f(x)) 6= 0 so that JS(x) is well-defined. Then, using the Schur decomposition of J(x) it is easily seen thatdet(J(x)) = det(S1(J(x))) det(−D2

2f(x)) = det(JS(x)) where the last equality holds since JS(x) is a lower blocktriangular matrix with S1(x) and −D2

2f(x) on the diagonal at critical points. Hence, the result.

This equivalence between the non-degeneracy of the Hessian and the game Jacobian J (and the relationship to the determinantof JS via the Schur decomposition) allows us to lift the generic property of non-degeneracy of critical points of functions tocritical points of the Stackelberg learning dynamics.

The Jet Transversality Theorem and Proposition 15 can be used to show a subset of a jet bundle having a particular set ofdesired properties is generic. Indeed, consider the jet bundle Jk(X,Y ) and recall that it is a manifold that contains jetsjkf : X → Jk(X,Y ) as its elements where f ∈ Ck(X,Y ). Let Z ⊂ Jk(X,Y ) be the submanifold of the jet bundle thatdoes not possess the desired properties. If dimX < codim Z, then for a generic function f ∈ Ck(X,Y ) the image of thek–jet extension is disjoint from Z implying that there is an open–dense set of functions having the desired properties.

Without loss of generality, we let player 1 be the leader.Lemma 7. Consider f ∈ Cq(Rm1+m2 ,R), q ≥ 2 such thatD2

i f ∈ Rmi×mi . It is a generic property that det(D2i f(x)) 6= 0

for any i = 1, 2.

Proof. Let us start with f ∈ Cq with q ≥ 3. First, critical points of a function f are such that Dif(x) = 0, i = 1, 2.Furthermore, the J2(Rm,R) bundle associated to f is diffeomorphic to Rm×R×Rm×Rm(m+1)/2 and the 2-jet extensionof f at any point x ∈ Rm is given by (x, f(x), Df(x), D2f(x)).

Now, let us denote by S(k) the space of k × k symmetric matrices, and consider the subset of J2(Rm,R) defined by

Di = Rm × R× {0} × Zi(mi)× Rm1×m2 × S(m−mi)

where Z(mi) = {A ∈ S(mi)| det(A) = 0}. The set Z(mi) is algebraic; hence, we can use the Whitney stratificationtheorem (Gibson et al., 1976, Ch. 1, Thm. 2.7) to get that each Z(mi) is the union of submanifolds of co-dimension atleast 1. Hence, it is the union of sub-manifolds of codimension at least one and, in turn, Di is the union of sub-manifoldsof codimension at least m + 1. Thus, it follows from the Jet Transversality theorem 11 (by way of Proposition 15 sincem+ 1 > m) that for a generic function f , the image of the 2-jet extension j2f is disjoint from Di. Hence, for such an f ,for each x that is a critical point, the Hessian of f is such that det(D2

i f(x)) 6= 0.

Furthermore, if we consider the subset D ⊂ J2(Rm,R) defined by

D = Rm × R× {0} × Z(mi)× Rm1×m2 × Z(m−mi),

then both Z(mi) and Z(m −mi) are algebraic, and so they each are of co-dimension at least one. In turn, D is the theunion of sub-manifolds of codimension at least m+ 2. Applying the Jet Transversality Theorem 11 again, we get that forsuch an f , for each x that is a critical point, the Hessian of f is such that det(D2

i f(x)) 6= 0 for i ∈ {1, 2}.

The extension to the q ≥ 2 setting follows directly from the fact that non-degeneracy is an open condition in the C2 topology,and any function can be C2 approximated by a C3 function (see, e.g., (?, Thm. 2.4)), which can then be approximatedby a function without critical points such that det(D2

i f(x)) = 0 (by the above argument), which we will call coordinatedegenerate. This, in turn, implies that functions without critical points that are coordinate-degenerate are dense in the C2

space of functions.

While the theorems we leverage from differential geometry and dynamical systems theory are similar, the architecture ofthe proof of the following theorem deviates quite a bit from (Mazumdar & Ratliff, 2019; Ratliff et al., 2014) due to theheirarchical structure of the game.

Page 38: Implicit Learning Dynamics in Stackelberg Games ...faculty.washington.edu/ratliffl/projects/papers/2019longstack.pdf · and hyperparameter optimization (Maclaurin et al.,2015) 1Anonymous

2035203620372038203920402041204220432044204520462047204820492050205120522053205420552056205720582059206020612062206320642065206620672068206920702071207220732074207520762077207820792080208120822083208420852086208720882089

Implicit Learning Dynamics in Stackelberg Games

Theorem 12. For two-player, q-differentiable zero-sum continuous games with q ≥ 2, differential Stackelberg equilibriaare generic amongst local Stackelberg equilibria. That is, given a generic f ∈ Cq(Rm,R), all local Stackelberg equilibriaof the game (f,−f) are differential Stackelberg equilibria.

Proof. Let J2(Rm,R) denote the second-order jet bundle containing 2–jets j2f such that f : Rm → R. Then, J2(Rm,R)

is locally diffeomorphic to Rm × R × Rm × Rm(m+1)

2 and the 2–jet extension of f at any point x ∈ Rm is given by(x, f(x), Df(x), D2f(x)).

By Lemma 7, we know thatD2 = Rm × R× {0} × Z(m2)× Rm1+m2 × S(m1)

has co-dimension at least m+ 1 in J2(Rm,R) so that there exists an open dense set of functions F2 ⊂ Cq(Rm,R) suchthat det(D2

2f(x)) 6= 0 at critical points (i.e., where (D1f(x), D2f(x)) = (0, 0)).

Now, we also know that there is an open dense set of functions F1 in Cq(Rm,R) such that det(H(x)) 6= 0 at criticalpoints. The intersection of open dense sets is open dense. Let F = F1 ∩ F2. Now, for any f ∈ F , we have that at criticalpoints det(H(x)) 6= 0 and det(D2

2f(x)) 6= 0. Hence, by Lemma 6, det(JS(x)) 6= 0 for all f ∈ F , and in particular,det(S1(x)) 6= 0.

For all functions f ∈ F , the critical points of ωS(x) = (Df(x),−D2f(x)) coincide with the critical points of the functionf . Indeed,

(D1f(x), D2f(x)) = (0, 0) ⇐⇒ (Df(x),−D2f(x)) = 0

since for all f ∈ F , det(D22f(x)) 6= 0 and D2f(x) = 0 so that the Cq implicit map at a critical point D2f(x) = 0 is

well-defined.

Thus, we have constructed an open dense set F ⊂ Cq(Rm,R) such that for all f ∈ F , if x ∈ Rm is a local Stackelbergequilibrium for (f,−f), then x is a non-degenerate differential Stackelberg equilibrium. Indeed, suppose f ∈ F andx ∈ Rm is a local Stackelberg equilibrium. Then, a necessary condition is that −D2f(x) = 0 and −D2

2f(x) ≥ 0. However,since f ∈ F , we have that det(−D2

2f(x)) = (−1)m2 det(D22f(x)) 6= 0 so that, in fact, −D2

2f(x) > 0. Hence, a localStackelberg equilibrium such that −D2f(x) = 0 and −D2

2f(x) > 0 necessarily satisfies ωS(x) = 0 and S1(x) ≥ 0.However, again since f ∈ F , and det(S1(x)) 6= 0 so that, in fact, S1(x) > 0. Furthermore, due to the lower block triangularstructure of JS(x), det(−D2

2f(x)) = (−1)m2 det(D22f(x)) 6= 0 and det(S1(x)) 6= 0 also imply that det(JS(x)) 6= 0,

which completes the proof.

Corollary 7. For two-player, q-differentiable zero-sum continuous games, local Stackelberg equilibria are genericallynon-degenerate, hyperbolic critical points of the vector field ωS(x).

This corollary follows direction from Theorem 12 and Proposition 13. Hyperbolicity implies that there are no eigenvalues ofJS(x) with zero real part.

I.3. Structural Stability

Structural stability follows naturally from genericity. We further show that (non-degenerate) differential Stackelberg arestructurally stable, meaning that they persist under smooth perturbations within the class of zero-sum games.Theorem 13. For zero-sum games, differential Stackelberg equilibria are structurally stable: given f ∈ Cr(Rm1×Rm2 ,R),ζ ∈ Cr(Rm1 × Rm2 ,R), and a differential Stackelberg equilibrium (x1, x2) ∈ Rm1 × Rm2 , there exists neighborhoodsU ⊂ R of zero and V ⊂ Rm1 × Rm2 such that for all t ∈ U there exists a unique differential Stackelberg equilibrium(x1, x2) ∈ V for the zero-sum game (f + tζ,−f − tζ).

Proof. Let Rm = Rm1 × Rm2 . Define the smoothly perturbed cost function f : Rm × R → R by f(x, y, t) =f(x, y)+tζ(x, y), and ωS : Rm×R→ T ∗(Rm) by ωS(x, y, t) = (Df(x, y)),−D2f(x, y)), for all t ∈ R and (x, y) ∈ Rm.

Since (x1, x2) is a differential Stackelberg equilibrium, DωS(x, y, 0) is necessarily non-degenerate. Invoking the implicitfunction theorem (Lee, 2012), there exists neighborhoods V ⊂ R of zero andW ⊂ Rm and a smooth function σ ∈ Cr(V,W )such that for all t ∈ V and (x1, x2) ∈W , ωS(x1, x2, s) = 0 ⇐⇒ (x1, x2) = σ(t). Since ωS is continuously differentiable,there exists a neighborhood U ⊂ W of zero such that DωS(σ(t), t) is invertible for all t ∈ U . Thus, for all t ∈ U , σ(t)must be the unique local Stackelberg equilibrium of (f + tζ|W ,−f − tζ|W ).