+ All Categories
Home > Documents > Efficient Iterative Linear-Quadratic Approximations …dfk/pdfs/ILQGames.pdfrepeatedly refine an...

Efficient Iterative Linear-Quadratic Approximations …dfk/pdfs/ILQGames.pdfrepeatedly refine an...

Date post: 16-Jan-2020
Category:
Upload: others
View: 2 times
Download: 0 times
Share this document with a friend
8
Efficient Iterative Linear-Quadratic Approximations for Nonlinear Multi-Player General-Sum Differential Games David Fridovich-Keil, Ellis Ratner, Lasse Peters, Anca D. Dragan, and Claire J. Tomlin Abstract— Many problems in robotics involve multiple de- cision making agents. To operate efficiently in such settings, a robot must reason about the impact of its decisions on the behavior of other agents. Differential games offer an expressive theoretical framework for formulating these types of multi-agent problems. Unfortunately, most numerical so- lution techniques scale poorly with state dimension and are rarely used in real-time applications. For this reason, it is common to predict the future decisions of other agents and solve the resulting decoupled, i.e., single-agent, optimal control problem. This decoupling neglects the underlying interactive nature of the problem; however, efficient solution techniques do exist for broad classes of optimal control problems. We take inspiration from one such technique, the iterative linear- quadratic regulator (ILQR), which solves repeated approxima- tions with linear dynamics and quadratic costs. Similarly, our proposed algorithm solves repeated linear-quadratic games. We experimentally benchmark our algorithm in several examples with a variety of initial conditions and show that the resulting strategies exhibit complex interactive behavior. Our results indicate that our algorithm converges reliably and runs in real- time. In a three-player, 14-state simulated intersection problem, our algorithm initially converges in < 0.25 s. Receding hori- zon invocations converge in < 50 ms in a hardware collision- avoidance test. I. I NTRODUCTION Many problems in robotics require an understanding of how multiple intelligent agents interact. For example, in the intersection depicted in Fig. 1, two cars and a pedestrian wish to reach their respective goals without colliding or leaving their lanes. Successfully navigating the intersection requires either explicit, or perhaps implicit, coordination amongst the agents. Often, these interactions are decoupled, with each autonomous agent predicting the behavior of others and then planning an appropriate response. This decoupling necessi- tates strong predictive assumptions on how agents’ decisions impact one another. Differential game theory provides a principled formalism for expressing these types of multi- agent decision making problems without requiring a priori predictive assumptions. Unfortunately, most classes of differential games have no analytic solution, and many numerical techniques suffer from Department of EECS, UC Berkeley, {dfk, eratner, lasse.peters, anca, tomlin}@eecs.berkeley.edu. This research is supported by an NSF CAREER award, the Air Force Office of Scientific Research (AFOSR), NSF’s CPS FORCES and VeHICaL projects, the UC-Philippine-California Advanced Research Institute, the ONR MURI Embedded Humans, a DARPA Assured Autonomy grant, and the SRC CONIX Center. D. Fridovich-Keil is supported by an NSF Graduate Research Fellowship. E. Ratner is supported by a NASA Space Technology Research Fellowship. Fig. 1: Demonstration of the proposed algorithm for a three-player general- sum game modeling an intersection. Two cars (red and green triangles) navigate the intersection while a pedestrian (blue triangle) traverses a crosswalk. Observe how both cars swerve slightly to avoid one another and provide extra clearance to the pedestrian. the so-called “curse of dimensionality” [1]. Numerical dy- namic programming solutions for general nonlinear systems have been studied extensively, though primarily in cases with a priori known objectives and constraints which permit offline computation, such as automated aerial refueling [2]. Approaches such as [3, 4] which separate offline game analy- sis from online operation are promising. Still, scenarios with more than two players remain extremely challenging, and the practical restriction of solving games offline prevents them from being widely used in many applications of interest, such as autonomous driving. To simplify matters, decision making problems for multi- ple agents are often decoupled (see, e.g., [5–7]). For example, the red car in Fig. 1 may wish to simplify its decision prob- lem by predicting the future motion of the other agents and plan reactively. This simplification reduces the differential game to an optimal control problem, for which there often exist efficient solution techniques. However, the decisions of the other agents will depend upon what the red car chooses to do. By ignoring this dependence, the red car is incapable of discovering strategies which exploit the reactions of others, and moreover, trusting in a nominal prediction—e.g., that the pedestrian will get of the way—may lead to unsafe behavior. A differential game formulation of this problem, by contrast, explicitly accounts for the mutual dependence of all agents’ decisions.
Transcript
Page 1: Efficient Iterative Linear-Quadratic Approximations …dfk/pdfs/ILQGames.pdfrepeatedly refine an initial control strategy by efficiently solving linear-quadratic (LQ) approximations

Efficient Iterative Linear-Quadratic Approximationsfor Nonlinear Multi-Player General-Sum Differential Games

David Fridovich-Keil, Ellis Ratner, Lasse Peters, Anca D. Dragan, and Claire J. Tomlin

Abstract— Many problems in robotics involve multiple de-cision making agents. To operate efficiently in such settings,a robot must reason about the impact of its decisions onthe behavior of other agents. Differential games offer anexpressive theoretical framework for formulating these typesof multi-agent problems. Unfortunately, most numerical so-lution techniques scale poorly with state dimension and arerarely used in real-time applications. For this reason, it iscommon to predict the future decisions of other agents andsolve the resulting decoupled, i.e., single-agent, optimal controlproblem. This decoupling neglects the underlying interactivenature of the problem; however, efficient solution techniquesdo exist for broad classes of optimal control problems. Wetake inspiration from one such technique, the iterative linear-quadratic regulator (ILQR), which solves repeated approxima-tions with linear dynamics and quadratic costs. Similarly, ourproposed algorithm solves repeated linear-quadratic games. Weexperimentally benchmark our algorithm in several exampleswith a variety of initial conditions and show that the resultingstrategies exhibit complex interactive behavior. Our resultsindicate that our algorithm converges reliably and runs in real-time. In a three-player, 14-state simulated intersection problem,our algorithm initially converges in < 0.25 s. Receding hori-zon invocations converge in < 50ms in a hardware collision-avoidance test.

I. INTRODUCTION

Many problems in robotics require an understanding ofhow multiple intelligent agents interact. For example, in theintersection depicted in Fig. 1, two cars and a pedestrian wishto reach their respective goals without colliding or leavingtheir lanes. Successfully navigating the intersection requireseither explicit, or perhaps implicit, coordination amongst theagents. Often, these interactions are decoupled, with eachautonomous agent predicting the behavior of others and thenplanning an appropriate response. This decoupling necessi-tates strong predictive assumptions on how agents’ decisionsimpact one another. Differential game theory provides aprincipled formalism for expressing these types of multi-agent decision making problems without requiring a prioripredictive assumptions.

Unfortunately, most classes of differential games have noanalytic solution, and many numerical techniques suffer from

Department of EECS, UC Berkeley, {dfk, eratner,lasse.peters, anca, tomlin}@eecs.berkeley.edu.

This research is supported by an NSF CAREER award, the Air ForceOffice of Scientific Research (AFOSR), NSF’s CPS FORCES and VeHICaLprojects, the UC-Philippine-California Advanced Research Institute, theONR MURI Embedded Humans, a DARPA Assured Autonomy grant, andthe SRC CONIX Center. D. Fridovich-Keil is supported by an NSF GraduateResearch Fellowship. E. Ratner is supported by a NASA Space TechnologyResearch Fellowship.

Fig. 1: Demonstration of the proposed algorithm for a three-player general-sum game modeling an intersection. Two cars (red and green triangles)navigate the intersection while a pedestrian (blue triangle) traverses acrosswalk. Observe how both cars swerve slightly to avoid one anotherand provide extra clearance to the pedestrian.

the so-called “curse of dimensionality” [1]. Numerical dy-namic programming solutions for general nonlinear systemshave been studied extensively, though primarily in caseswith a priori known objectives and constraints which permitoffline computation, such as automated aerial refueling [2].Approaches such as [3, 4] which separate offline game analy-sis from online operation are promising. Still, scenarios withmore than two players remain extremely challenging, and thepractical restriction of solving games offline prevents themfrom being widely used in many applications of interest, suchas autonomous driving.

To simplify matters, decision making problems for multi-ple agents are often decoupled (see, e.g., [5–7]). For example,the red car in Fig. 1 may wish to simplify its decision prob-lem by predicting the future motion of the other agents andplan reactively. This simplification reduces the differentialgame to an optimal control problem, for which there oftenexist efficient solution techniques. However, the decisions ofthe other agents will depend upon what the red car chooses todo. By ignoring this dependence, the red car is incapable ofdiscovering strategies which exploit the reactions of others,and moreover, trusting in a nominal prediction—e.g., that thepedestrian will get of the way—may lead to unsafe behavior.A differential game formulation of this problem, by contrast,explicitly accounts for the mutual dependence of all agents’decisions.

Page 2: Efficient Iterative Linear-Quadratic Approximations …dfk/pdfs/ILQGames.pdfrepeatedly refine an initial control strategy by efficiently solving linear-quadratic (LQ) approximations

We propose a novel local algorithm for recovering inter-active strategies in a broad class of differential games. Thesestrategies qualitatively resemble local Nash equilibria, thoughthere are subtle differences. By solving the underlying gamewe account for the fundamental interactive nature of theproblem, and by seeking a local solution we avoid the curseof dimensionality which arises when searching for globalNash equilibria. Our algorithm builds upon the iterativelinear-quadratic regulator (ILQR) [8], a local method usedin smooth optimal control problems [9–11]. ILQR repeat-edly refines an initial control strategy by efficiently solvingapproximations with linear dynamics and quadratic costs.Like linear-quadratic (LQ) optimal control problems, LQgames also afford an efficient closed form solution [12]. Ouralgorithm exploits this analytic solution to solve successiveLQ approximations, and thereby finds a local solution to theoriginal game in real-time. For example, our algorithm ini-tially solves the three-player 14-state intersection scenario ofFig. 1 in < 0.25 s, and receding horizon problems convergein < 50 ms in a hardware collision-avoidance test.

The rest of the paper proceeds as follows. Section IIsummarizes related work in both differential game theoryand iterative LQ methods. Section III provides mathematicalpreliminaries. Section IV presents our iterative LQ solutiontechnique, describes its fixed points, and discusses runtimecomplexity. Section V concludes with several examples,including a Monte Carlo study of sensitivity to initial con-ditions and a proof-of-concept hardware test.

II. BACKGROUND & RELATED WORK

Differential games have been widely studied since theintroduction of pursuit-evasion games in [13]. Here, wesurvey both zero-sum and general-sum games and discussapproximate solution techniques. We also summarize itera-tive linear-quadratic methods used both for optimal controland games and discuss their relationship to this work.

A. Zero-sum games

In zero-sum games, two (groups of) players choose con-trol strategies to optimize equal and opposite objectives.Two-player zero-sum games are often formulated througha Hamilton-Jacobi-Isaacs (HJI) PDE, e.g. [14–18]. Morecomplicated games, such as active target defense and multi-player capture-avoid games are also addressed in a zero-sumframework in [19, 20] and [21], respectively.

B. General-sum games

Initially formulated in [22, 23], general-sum differen-tial games generalize zero-sum games to model situa-tions in which players have competing—but not necessar-ily opposite—objectives. Like zero-sum games, general-sumgames are characterized by Hamilton-Jacobi equations [22]in which all players’ Hamiltonians are coupled with oneother. Both zero-sum and general-sum games, and especiallygames with many players, are generally difficult to solvenumerically. However, efficient methods do exist for solvinggames with linear dynamics and quadratic costs, e.g. [12, 24].

Dockner et al. [25] also characterize classes of games whichadmit tractable open loop, rather than feedback, solutions.

C. Approximation techniques

While general-sum games may be analyzed by solvingcoupled Hamilton-Jacobi equations [22], doing so requiresboth exponential time and computational memory. A numberof more tractable approximate solution techniques have beenproposed for zero-sum games, many of which require linearsystem dynamics, e.g. [26–29], or decomposable dynamics[30]. Approximate dynamic programming techniques such as[31] are not restricted to linear dynamics or zero-sum set-tings. Still, scalability to online, real-time operation remainsa challenge.

Iterative best response algorithms form another class of ap-proximate methods for solving general-sum games. Here, ineach iteration every player solves (or approximately solves)the optimal control problem that results from holding otherplayers’ strategies fixed. This reduction to a sequence ofoptimal control problems is attractive; however, it can alsobe computationally inefficient. Still recent work demonstratesthe effectiveness of iterative best response in lane changes[4] and multi-vehicle racing [32].

Another similarly-motivated class of approximations in-volves changing the information structure of the game. Forexample, Chen et al. [33] solve a multi-player reach-avoidgame by pre-specifying an ordering amongst the playersand allowing earlier players to communicate their intendedstrategies to later players. Zhou et al. [34] and Liu et al.[35] operate in a similar setting, but solve for open-loopconservative strategies.

D. Iterative linear-quadratic (LQ) methods

Iterative LQ approximation methods are increasingly com-mon in the robotics and control communities. Our workbuilds directly upon the iterative linear-quadratic regula-tor (ILQR) algorithm [8, 36]. ILQR is closely related todifferential dynamic programming [37, 38], and is widelyused to find local solutions to smooth nonlinear optimalcontrol problems. ILQR has been demonstrated in a varietyof applications including driving [11], humanoid locomotion[9], and grasping [10]. There are also many extensions toILQR, including trajectory smoothing [39] and constrainthandling via barrier functions [11].

At each iteration, ILQR simulates the full nonlinear systemtrajectory, computes a discrete-time linear dynamics approx-imation and quadratic cost approximation, and solves a LQRsubproblem to generate the next control strategy iterate.While structurally similar to ILQR, our approach solves aLQ game at each iteration instead of a LQR problem. Thiscore idea is related to the sequential linear-quadratic methodof [40, 41], which is restricted to the two-player zero-sumcontext. In this paper, we show that LQ approximations canbe applied in N -player, general-sum games. In addition, weexperimentally characterize the quality of solutions in severalcase studies and demonstrate real-time operation.

Page 3: Efficient Iterative Linear-Quadratic Approximations …dfk/pdfs/ILQGames.pdfrepeatedly refine an initial control strategy by efficiently solving linear-quadratic (LQ) approximations

III. PROBLEM FORMULATION

We consider a N -player finite horizon general-sum differ-ential game characterized by nonlinear system dynamics

x = f(t, x, u1:N ) , (1)

where x ∈ Rn is the state of the system, and ui ∈ Rmi , i ∈[N ] ≡ {1, . . . , N} is the control input of player i, andu1:N ≡ (u1, u2, . . . , uN ). Each player has a cost function Jidefined as an integral of running costs gi. Ji is understood todepend implicitly upon the state trajectory x(·), which itselfdepends upon initial state x(0) and control signals u1:N (·):

Ji(u1:N (·)

),∫ T

0

gi(t, x(t), u1:N (t)

)dt,∀i ∈ [N ] . (2)

We shall presume that f is continuous in t and continu-ously differentiable in {x, ui} uniformly in t. We shall alsorequire gi to be twice differentiable in {x, ui},∀t.

Ideally, we would like to find time-varying state feedbackcontrol strategies γ∗i ∈ Γi for each player i which constitutea global Nash equilibrium for the game defined by (1) and(2). Here, the strategy space Γi for player i is the set ofmeasurable functions γi : [0, T ] × Rn → Rmi mappingtime and state to player i’s control input. Note that, inthis formulation, player i only observes the state of thesystem at each time and is unaware of other players’ controlstrategies. With a slight abuse of notation Ji(γ1; . . . ; γN ) ≡Ji(γ1(·, x(·)), . . . , γN (·, x(·))

), the global Nash equilibrium

is defined as the set of strategies {γi} for which the followinginequalities hold (see, e.g., [12, Chapter 6]):

J∗i , Ji(γ∗1 ; . . . ; γ∗i−1; γ∗i ; γ∗i+1; . . . γ∗N )

≤ Ji(γ∗1 ; . . . ; γ∗i−1; γi; γ∗i+1; . . . γ∗N ),∀i ∈ [N ] .

(3)

In (3), the inequalities must hold for all γi ∈ Γi,∀i ∈[N ]. Informally, a set of feedback strategies (γ∗1 , . . . , γ

∗N )

is a global Nash equilibrium if no player has a unilateralincentive to deviate from their current strategy.

Since finding a global Nash equilibrium is generally com-putationally intractable, recent work in adversarial learning[42] and motion planning [32, 43] consider local Nashequilibria instead. Further, [32, 43] simplify the informationstructure of the game and consider open loop, rather thanfeedback, strategies. Local Nash equilibria are characterizedsimilarly to (3), except that the inequalities may only holdin a local neighborhood within the strategy space [44,Definition 1]. In this paper, we shall seek a related type ofequilibrium, which we describe more precisely in Section IV-B. Intuitively, we seek strategies which satisfy the globalNash conditions (3) for the limit of a sequence of localapproximations to the game. Although it is difficult to for-mally characterize the relationship between this equilibriumconcept and a local Nash equilibrium, our experimentalresults indicate that it does yield highly interactive strategiesin a variety of differential games.

Algorithm 1: Iterative LQ GamesInput: initial state x(0), control strategies {γ0

i }i∈[N ],time horizon T , running costs {gi}i∈[N ]

Output: converged control strategies {γ∗i }i∈[N ]

1 for iteration k = 1, 2, . . . do2 ξk ≡ {x(t), u1:N (t)}t∈[0,t] ←3 getTrajectory

(x(0), {γk−1

i });

4 {A(t), Bi(t)} ← linearizeDynamics(ξk);

5 {li(t), Qi(t), rij(t), Rij(t)} ←quadraticizeCost

(ξk);

6 {γki } ← solveLQGame(

7 {A(t), Bi(t), li(t), Qi(t), rij(t), Rij(t)});

8 {γki } ← stepToward({γk−1

i , γki });

9 if converged then10 return {γki }

IV. ITERATIVE LINEAR-QUADRATIC GAMES

We approach the N -player general-sum game with dynam-ics (1) and costs (2) from the perspective of classical LQgames. It is well known that Nash equilibrium strategies forfinite-horizon LQ games satisfy coupled Riccati differentialequations. These coupled Riccati equations may be derivedby substituting linear dynamics and quadratic running costsinto the generalized HJ equations [23] and analyzing thefirst order necessary conditions of optimality for each player[12, Chapter 6]. These coupled differential equations maybe solved approximately in discrete-time using dynamicprogramming [12]. We will leverage the existence and com-putational efficiency of this discrete-time LQ solution tosolve successive approximations to the original nonlinearnonquadratic game.

A. Iterative LQ game algorithm

Our iterative LQ game approach proceeds in stages, assummarized in Algorithm 1. We begin with an initial statex(0) and initial feedback control strategies {γ0

i } for eachplayer i, and integrate the system forward (line 3 of Al-gorithm 1) to obtain the current trajectory iterate ξk ≡{x(t), u1:N (t)}t∈[0,T ]. Next (line 4) we obtain a Jacobianlinearization of the dynamics f about trajectory ξk. At eachtime t ∈ [0, T ] and for arbitrary states x(t) and controls ui(t)we define deviations from this trajectory δx(t) = x(t)− x(t)and δui(t) = ui(t) − ui(t). Thus equipped, we compute acontinuous-time linear system approximation about ξk:

˙δx(t) ≈ A(t)δx(t) +∑i∈[N ]

Bi(t)δui(t), (4)

where A(t) is the Jacobian Dxf(t, x(t), u1:N (t)

)and Bi(t)

is likewise Duif(t, x(t), u1:N (t)

).

We also obtain a quadratic approximation to the running

Page 4: Efficient Iterative Linear-Quadratic Approximations …dfk/pdfs/ILQGames.pdfrepeatedly refine an initial control strategy by efficiently solving linear-quadratic (LQ) approximations

cost gi for each player i (see line 5 of Algorithm 1)

gi(t, x(t), u1:N (t)

)≈

gi(t, x(t), u1:N (t)

)+

1

2δx(t)T (Qi(t)δx(t) + 2li(t)) +

1

2

∑j∈[N ]

δuj(t)T (Rij(t)δuj(t) + 2rij(t)) , (5)

where vector li(t) is the gradient Dxgi, rij is Dujgi, and

matrices Qi and Rij are Hessians D2xxgi and D2

uj ujgi,

respectively. We neglect mixed partials D2uj uk

gi and D2xuj

gias they rarely appear in cost structures of practical interest,although they could be incorporated if needed.

Thus, we have constructed a finite-horizon continuous-time LQ game, which may be solved via coupled Riccatidifferential equations [12, 45]. This results in a new setof candidate feedback strategies {γki } which constitute afeedback (global) Nash equilibrium of the LQ game [12]. Infact, these feedback strategies are affine maps of the form:

γki(t, x(t)

)= ui(t)− P k

i (t)δx(t)− αki (t) , (6)

with gains P ki (t) ∈ Rmi×n and affine terms αk

i (t) ∈ Rmi .However, we find that choosing γki = γki often causes

Algorithm 1 to diverge because the trajectory resulting from{γi} is far enough from the current trajectory iterate ξk thatthe dynamics linearizations (Algorithm 1, line 4) and costquadraticizations (line 5) no longer hold. As in ILQR [46],to improve convergence, we take only a small step in the“direction” of γki .1 More precisely, for some choice of stepsize η ∈ (0, 1], we set

γki(t, x(t)

)= ui(t)− P k

i (t)δx(t)− ηαki (t) , (7)

which corresponds to line 8 in Algorithm 1. Note that att = 0, δx(0) = 0 and γki

(0, x(0)

)= ui(0)− ηαk

i (0). Thus,taking η = 0, we have γki

(t, x(t)

)= ui(t) (which may

be verified recursively). That is, when η = 0 we recoverthe open-loop controls from the previous iterate, and hencex(t) = x(t). Taking η = 1, we recover the LQ solution in(6). Similar logic implies the following lemma.

Lemma 1: Suppose that trajectory ξ∗ is a fixed point ofAlgorithm 1, with η 6= 0. Then, the converged affine terms{α∗i (t)} must all be identically zero for all time.

In ILQR, it is important to perform a line-search overstep size η to ensure a sufficient decrease in cost at everyiteration, and thereby improve convergence (e.g., [46]). In thecontext of a noncooperative game, however, line-searchingto decrease “cost” does not make sense, as costs {Ji} mayconflict. For this reason, like other local methods in games(e.g., [32]), our approach is not guaranteed to converge fromarbitrary initializations. In practice, however, we find thatour algorithm typically converges for a fixed, small step size(e.g. η = 0.01). Heuristically decaying step size with eachiteration k or line-searching until ‖ξk − ξk−1‖ is smallerthan a threshold are also promising alternatives. Further

1We also note that, in practice, it is often helpful to “regularize” theproblem by adding scaled identity matrices εI to Qi and/or Rij .

investigation of line-search methods in games is a rich topicof future research.

Note: Although we have presented our algorithm incontinuous-time, in practice, we solve the coupled Riccatiequations analytically in discrete-time via dynamic program-ming. Please refer to [12, Corollary 6.1] for a full derivation.To discretize time at resolution ∆t, we employ Runge-Kuttaintegration of nonlinear dynamics (1) with a zero-order holdfor control input over each time interval ∆t. That is, wenumerically compute:

x(t+ ∆t) = x(t)+

∫ t+∆t

t

f(s, x(s), u1:N (s))ds

where ui(s) = γk−1i

(t, x(t)

),∀i ∈ [N ], and

x(0) = x(0). (8)

B. Characterizing fixed points

Suppose Algorithm 1 converges to a fixed point(γ∗1 , . . . , γ

∗N ). These strategies are the global Nash equi-

librium of a local LQ approximation of the original gameabout the limiting operating point ξ∗. While it is tempting topresume that such fixed points are also local Nash equilibriaof the original game, this is not always true because con-verged strategies are only optimal for a LQ approximationof the game at every time rather than the original game.This approximation neglects higher order coupling effectsbetween each player’s running cost gi and other players’inputs uj , j 6= i. These coupling effects arise in the gamesetting but not in the optimal control setting, where ILQRconverges to local minima.

We defer a more detailed, formal characterization of therelationship between these fixed points and local Nash equi-libria for future work. The remainder of this paper providesa computational and empirical evaluation of Algorithm 1. Inpractice, we observe that fixed points exhibit the nontrivial,coordinated interactions which arise in multi-agent roboticsapplications and originally motivated our study of differentialgames in Section I.

C. Computational complexity and runtime

The per-iteration computational complexity of our ap-proach is comparable to that of ILQR, and scales modestlywith the number of players, N . Specifically, at each iteration,we first linearize system dynamics about ξk. Presuming thatthe state dimension n is larger than the control dimension mi

for each player, linearization requires computing O(n2) par-tial derivatives at each time step (which also holds for ILQR).We also quadraticize costs, which requires O(Nn2) partialderivatives at each time step (compared to O(n2) for ILQR).Finally, solving the coupled Riccati equations of the resultingLQ game at each time step has complexity O(N3n3), whichmay be verified by inspecting [12, Corollary 6.1] (for ILQR,this complexity is O(n3)).

Total algorithmic complexity depends upon the numberof iterations, which we currently have no theory to bound.However, empirical results are extremely promising. Forthe three-player 14-state game described in Section V-B,

Page 5: Efficient Iterative Linear-Quadratic Approximations …dfk/pdfs/ILQGames.pdfrepeatedly refine an initial control strategy by efficiently solving linear-quadratic (LQ) approximations

each iteration takes < 8 ms and the entire game can besolved from a zero initialization (P 0

i (·) = 0, α0i (·) = 0)

in < 0.25 s. Moreover, receding horizon invocations in ahardware collision-avoidance test can be solved in < 50 ms(Section V-C). All computation times are reported for single-threaded operation on a 2017 MacBook Pro with a 2.8GHz Intel Core i7 CPU. For reference, the iterative bestresponse scheme of [43] reports solving a receding horizontwo-player zero-sum racing game at 2 Hz, and the methodof [41] reportedly takes several minutes to converge for adifferent two-player zero-sum example. The dynamics andcosts in both cases differ from those in Section V (or are notclearly reported); nonetheless, the runtime of our approachcompares favorably.

V. EXAMPLES

In this section, we demonstrate our algorithm experimen-tally in three-player noncooperative settings, both in softwaresimulation and hardware.2

A. Monte Carlo study

We begin by presenting a Monte Carlo study of theconvergence properties of Algorithm 1. As we shall see,the solution to which Algorithm 1 converges depends uponthe initial strategy of each player, γ0

i . For clarity, we studythis sensitivity in a game with simplified cost structure sothat differences in solution are more easily attributable tocoupling between players.

Concretely, we consider a three-player “hallway naviga-tion” game with time horizon 10 s and discretization 0.1 s.Here, three people wish to interchange positions in a narrowhallway while maintaining at least 1 m clearance betweenone another. We model each player i’s motion as:

px,i = vi cos(θi) , θi = ωi ,

py,i = vi sin(θi) , vi = ai ,(9)

where pi := (px,i, py,i) denotes player i’s position, θiheading angle, vi speed, and input ui := (ωi, ai) yawrate and longitudinal acceleration. Concatenating all players’states into a global state vector x := (px,i, py,i, θi, vi)

3i=1, the

game has 12 state dimensions and six input dimensions.We encode this problem with running costs gi (2) ex-

pressed as weighted sums of the following:

wall: 1{|py,i| > dhall}(|py,i| − dhall)2 (10)

proximity: 1{‖pi − pj‖ < dprox}(dprox − ‖pi − pj‖)2 (11)

goal: 1{t > T − tgoal}‖pi − pgoal,i‖2 (12)

input: uTi Riiui (13)

Here, 1{·} is the indicator function, i.e., it takes the value 1if the given condition holds, and 0 otherwise. dhall and dproxdenote threshold distances from hallway center and betweenplayers, which we set to 0.75 and 1 m, respectively. The goalcost is active only for the last tgoal seconds, and the goalposition is given by pgoal,i for each player i. Control inputs

2Video summary available at https://youtu.be/KPEPk-QrkQ8.

Fig. 2: Monte Carlo results for a three-player hallway navigation game.(A1, B1, C1) Converged trajectories clustered by total Euclidean distance;each cluster corresponds to a qualitatively distinct mode of interaction. (A2,B2, C2) Costs for each player at each solver iteration. The shaded regioncorresponds to one standard deviation. (D) Several converged trajectories didnot match a cluster (A-C). (E) Trajectories resulting from 500 random initialstrategies. (F) Histogram of iterations until state trajectory has converged.

are penalized quadratically, with Rii a diagonal matrix. Thehallway is too narrow for all players to cross simultaneouslywithout incurring a large proximity cost; hence, this proxim-ity cost induces strong coupling between players’ strategies.

Fig. 2 displays the results of our Monte Carlo study.We seed Algorithm 1 with 500 random sinusoidal open-loop initial strategies, which correspond to the trajectoriesshown in Fig. 2(E). From each of these initializations, werun Algorithm 1 for 100 iterations and cluster the resultingtrajectories by Euclidean distance. As shown in Fig. 2(A1,B1, C1), these clusters correspond to plausible modes ofinteraction; in each case, one or more players incur slightlyhigher cost to make room for the others to pass. Beside eachof these clusters in Fig. 2(A2, B2, C2), we also report themean and standard deviation of each player’s cost at eachsolver iteration. As shown in Fig. 2(F), state trajectoriesconverge within an `∞ tolerance of 0.01 in well under 100iterations.

In these 500 random samples, only 6 did not converge and

Page 6: Efficient Iterative Linear-Quadratic Approximations …dfk/pdfs/ILQGames.pdfrepeatedly refine an initial control strategy by efficiently solving linear-quadratic (LQ) approximations

0.0 ∑ t ∑ 0.8 0.8 ∑ t ∑ 1.5 1.5 ∑ t ∑ 5.0

1.5 ≤ t ≤ 5.00.0 ∑ t ∑ 0.8 0.8 ∑ t ∑ 1.5 1.5 ∑ t ∑ 5.0

0.8 ≤ t ≤ 1.50.0 ∑ t ∑ 0.8 0.8 ∑ t ∑ 1.5 1.5 ∑ t ∑ 5.0

Goals

LanesCar

Pedestrian

Crosswalk

0.0 ≤ t ≤ 0.8

Car

Fig. 3: Three-player intersection game. (Left) Green car seeks the lanecenter and then swerves slightly to avoid the pedestrian. (Center) Red caraccelerates in front of the green car and slows slightly to allow the pedestrianto pass. (Right) Red car swerves left to give pedestrian a wide berth.

had to be resampled, and 5 converged to trajectories whichwere outliers from the clusters depicted in Fig. 2(A-C). Theseoutliers are shown in Fig. 2(D). We observe that, in these 5cases, the players come within 0.5 m of one another.

B. Three-player intersection

Next, we consider a more complicated game intended tomodel traffic at an intersection. As shown in Fig. 3, weconsider an intersection with two cars and one pedestrian,all of which must cross paths to reach desired goal locations.We use a time horizon of 5 s with discretization of 0.1 s, andAlgorithm 1 terminates in under 0.25 s.

We model the pedestrian’s dynamics as in (9) and eachcars i’s dynamics as follows:

px,i = vi cos(θi) , θi = vi tan(φi)/Li, φi = ψi

py,i = vi sin(θi) , vi = ai ,(14)

where the state variables are as before (9) except for frontwheel angle φi. Li is the inter-axle distance, and inputui := (ψi, ai) is the front wheel angular rate and longitudinalacceleration, respectively. Together, the state of this game is14-dimensional.

The running cost for each player i are specified asweighted sums of (11)–(13), and the following:

lane center: d`(pi)2 (15)

lane boundary: 1{d`(pi) > dlane}(dlane − d`(pi))2 (16)

nominal speed: (vi − vref,i)2 (17)

speed bounds: 1{vi > vi}(vi − vi)2

+ 1{vi < vi}(vi − vi)2 (18)

Here, dlane denotes the lane half-width, and d`(pi) :=minp`∈` ‖p` − pi‖ measures player i’s distance to lanecenterline `. Speed vi is penalized quadratically away froma fixed reference vref,i also outside limits vi and vi.

Fig. 3 shows a time-lapse of the converged solution iden-tified by Algorithm 1. These strategies exhibit non-trivial co-ordination among the players as they compete to reach theirgoals efficiently while sharing responsibility for collision-avoidance. Such competitive behavior would be difficult forany single agent to recover from a decoupled, optimal controlformulation. Observe how, between 0 ≤ t ≤ 0.8 s (left), thegreen car initially seeks the lane center to minimize its cost,

but then turns slightly to avoid the pedestrian (blue). Between0.8 ≤ t ≤ 1.5 s (center), the red car turns right to pass infront of the green car, and then slows and begins to turn leftto give the pedestrian time to cross. Finally (right), the redcar turns left to give the pedestrian a wide berth.

C. Receding horizon motion planning

Differential games are appropriate in a variety of ap-plications including multi-agent modeling and coordinatedplanning. Here we present a proof-of-concept for their usein single-agent planning in a dynamic environment. In thissetting, a single robot operates amongst multiple other agentswhose true objectives are unknown. The robot models theseobjectives and formulates the interaction as a differentialgame. Then, crucially, the robot re-solves the differentialgame along a receding time horizon to account for possibledeviations between the other agents’ decisions and thosewhich result from the game solution.

We implement Algorithm 1 in C++3 within the RobotOperating System (ROS) framework, and evaluate it in areal-time hardware test onboard a TurtleBot 2 ground robot,in a motion capture room with two human participants.The TurtleBot wishes to cross the room while maintaining> 1 m clearance to the humans, and it models the humanslikewise. We model the TurtleBot dynamics as (9) andhumans likewise but with constant speed vi, i.e.:

px,i = vi cos(θi) , py,i = vi sin(θi) , θi = ωi . (19)

We use a similar cost structure as in Section V-B, andinitialize Algorithm 1 with all agents’ strategies identicallyzero (i.e., P 0

i (·), α0i (·) ≡ 0). We re-solve the game in a 10 s

receding horizon with time discretization of 0.1 s, and warm-start each successive receding horizon invocation with theprevious solution. Replanning every 0.25 s, Algorithm 1 re-liably converges in under 50 ms. We gather state informationfor all agents using a motion capture system. Fig. 4 showsa time-lapse of a typical interaction.

Initially, in frame (a) Algorithm 1 identifies a set ofstrategies which steer each agent to their respective goalswhile maintaining a large separation. Of course, the humanparticipants do not actually follow these precise trajectories;hence later receding horizon invocations converge to slightlydifferent strategies. In fact, between frames (c) and (d) thered participant makes an unanticipated sharp right-hand turn,which forces the (blue) robot to stay to the right of itsprevious plan and then turn left in order to maintain sufficientseparation between itself and both humans. We note thatour assumed cost structure models all agents as wishingto avoid collision. Thus, the resulting strategies may beless conservative than those that would arise from a non-game-theoretic motion planning approach. As our primaryobjective is to demonstrate the real-time performance ofAlgorithm 1, we leave a more complete study of agent intentmodeling and its impact on local game solutions for futurework.

3Code available at: github.com/HJReachability/ilqgames

Page 7: Efficient Iterative Linear-Quadratic Approximations …dfk/pdfs/ILQGames.pdfrepeatedly refine an initial control strategy by efficiently solving linear-quadratic (LQ) approximations

5RERW�JRDO +XPDQ�JRDOV

�D� �H��G��E� �F�Fig. 4: Time-lapse of a hardware demonstration of Algorithm 1. We model the interaction of a ground robot (blue triangle) and two humans (purple andred triangles) using a differential game in which each agent wishes to reach a goal location while maintaining sufficient distance from other agents. Ouralgorithm solves receding horizon instantiations of this game in real-time, and successfully plans and executes interactive collision-avoiding maneuvers.Planned (and predicted) trajectories are shown in blue (robot), purple, and red (humans).

VI. DISCUSSION

We have presented a novel algorithm for finding localsolutions in multi-player general-sum differential games. Ourapproach is closely related to the iterative linear-quadraticregulator (ILQR) [8], and offers a straightforward way foroptimal control practitioners to directly account for multi-agent interactions via differential games. We performed aMonte Carlo study which demonstrated the reliability ofour algorithm and its ability to identify complex interactivestrategies for multiple agents. These solutions display thecompetitive behavior associated with local Nash equilibria,although there are subtle differences. We also showcased ourmethod in a three-player 14-dimensional traffic example, andtested it in real-time operation in a hardware robot navigationscenario, following a receding time horizon.

There are several other approaches to identifying localsolutions in differential games, such as iterative best response[32]. We have shown the computational efficiency of ourapproach. However, quantitatively comparing the solutionsidentified by different algorithms is challenging due to differ-ences in equilibrium concept, information structure (feedbackvs. open loop), and implementation details. Furthermore, inarbitrary general-sum games, different players may preferdifferent equilibria. Studying the qualitative differences inthese equilibria is an important direction of future research.

Although our experiments show that our algorithm con-verges reliably, we have no a priori theoretical guarantee ofconvergence from arbitrary initializations. Future work willseek a theoretical explanation of this empirical property. Wealso intend to investigate inequality-constrained differentialgames; here, we believe that interior point methods maypresent a promising direction. Finally, it will be critical todevelop a theory for online estimation of other players’objectives, and for understanding the sensitivity of localsolutions to misspecified objectives and sub-optimal play.

ACKNOWLEDGMENTS

The authors would like to thank Andrew Packard for hishelpful insights on LQ games, as well as Forrest Laine, Chih-Yuan Chiu, Somil Bansal, Jaime Fisac, Tyler Westenbroek,and Eric Mazumdar for helpful discussions.

REFERENCES

[1] R. Bellman. Dynamic programming. Tech. rep. RANDCORP SANTA MONICA CA, 1956.

[2] J. Ding, J. Sprinkle, S. S. Sastry, and C. J. Tomlin. “Reach-ability calculations for automated aerial refueling”. 47thConference on Decision and Control (CDC). IEEE. 2008.

[3] S. L. Herbert*, M. Chen*, S. Han, S. Bansal, J. F. Fisac,and C. J. Tomlin. “FaSTrack: a Modular Framework for Fastand Guaranteed Safe Motion Planning”. 56th Conference onDecision and Control (CDC) (2017).

[4] J. F. Fisac, E. Bronstein, E. Stefansson, D. Sadigh, S. S.Sastry, and A. D. Dragan. “Hierarchical game-theoretic plan-ning for autonomous vehicles”. International Conference onRobotics and Automation (ICRA). IEEE. 2019.

[5] B. D. Ziebart, N. Ratliff, G. Gallagher, C. Mertz, K. Peterson,J. A. Bagnell, M. Hebert, A. K. Dey, and S. Srinivasa.“Planning-based prediction for pedestrians”. InternationalConference on Intelligent Robots and Systems (IROS). IEEE.2009.

[6] H. Bai, S. Cai, N. Ye, D. Hsu, and W. S. Lee. “Intention-aware online POMDP planning for autonomous driving in acrowd”. International Conference on Robotics and Automa-tion (ICRA). IEEE. 2015.

[7] E. Schmerling, K. Leung, W. Vollprecht, and M. Pavone.“Multimodal probabilistic model-based planning for human-robot interaction”. 2018 IEEE International Conference onRobotics and Automation (ICRA). IEEE. 2018.

[8] W. Li and E. Todorov. “Iterative linear quadratic regu-lator design for nonlinear biological movement systems.”ICINCO. 2004.

[9] J. Koenemann, A. Del Prete, Y. Tassa, E. Todorov, O.Stasse, M. Bennewitz, and N. Mansard. “Whole-body model-predictive control applied to the HRP-2 humanoid”. Interna-tional Conference on Intelligent Robots and Systems (IROS).IEEE. 2015.

Page 8: Efficient Iterative Linear-Quadratic Approximations …dfk/pdfs/ILQGames.pdfrepeatedly refine an initial control strategy by efficiently solving linear-quadratic (LQ) approximations

[10] N. Kitaev, I. Mordatch, S. Patil, and P. Abbeel. “Physics-based trajectory optimization for grasping in cluttered en-vironments”. International Conference on Robotics and Au-tomation (ICRA). IEEE. 2015.

[11] J. Chen, W. Zhan, and M. Tomizuka. “Constrained iterativeLQR for on-road autonomous driving motion planning”. In-ternational Conference on Intelligent Transportation Systems(ITSC). IEEE. 2017.

[12] T. Basar and G. J. Olsder. Dynamic noncooperative gametheory. SIAM, 1999.

[13] R. Isaacs. Games of pursuit. Tech. rep. Rand Corporation,1951.

[14] R. Isaacs. Differential games: a mathematical theory withapplications to warfare and pursuit, control and optimiza-tion. Courier Corporation, 1999.

[15] L. C. Evans and P. E. Souganidis. Differential Games andRepresentation Formulas for Solutions of Hamilton-Jacobi-Isaacs Equations. Tech. rep. Wisconsin University Mathe-matics Research Center, 1983.

[16] P.-L. Lions and P. E. Souganidis. “Differential games, opti-mal control and directional derivatives of viscosity solutionsof Bellman’s and Isaacs’ equations”. Journal on Control andOptimization 23.4 (1985).

[17] I. M. Mitchell, A. M. Bayen, and C. J. Tomlin. “A time-dependent Hamilton-Jacobi formulation of reachable setsfor continuous dynamic games”. Transactions on AutomaticControl 50.7 (2005).

[18] K. Margellos and J. Lygeros. “Hamilton–Jacobi formulationfor reach–avoid differential games”. Transactions on Auto-matic Control 56.8 (2011).

[19] E. Garcia, D. W. Casbeer, and M. Pachter. “Design and anal-ysis of state-feedback optimal strategies for the differentialgame of active defense”. IEEE Transactions on AutomaticControl 64.2 (2019).

[20] J. F. Fisac and S. S. Sastry. “The pursuit-evasion-defensedifferential game in dynamic constrained environments”.54th Conference on Decision and Control (CDC). IEEE.2015.

[21] Z. Zhou, W. Zhang, J. Ding, H. Huang, D. M. Stipanovic, andC. J. Tomlin. “Cooperative pursuit with Voronoi partitions”.Automatica 72 (2016).

[22] A. W. Starr and Y.-C. Ho. “Nonzero-sum differential games”.Journal of Optimization Theory and Applications 3.3 (1969).

[23] A. Starr and Y.-C. Ho. “Further properties of nonzero-sumdifferential games”. Journal of Optimization Theory andApplications 3.4 (1969).

[24] T. Li and Z. Gajic. “Lyapunov iterations for solving coupledalgebraic Riccati equations of Nash differential games andalgebraic Riccati equations of zero-sum games”. New trendsin dynamic games and applications. Springer, 1995.

[25] E. Dockner, G. Feichtinger, and S. Jørgensen. “Tractableclasses of nonzero-sum open-loop Nash differential games:theory and examples”. Journal of Optimization Theory andApplications 45.2 (1985).

[26] A. B. Kurzhanski and P. Varaiya. “Ellipsoidal techniquesfor reachability analysis: internal approximation”. Systems& Control Letters 41.3 (2000).

[27] A. B. Kurzhanski and P. Varaiya. “On ellipsoidal techniquesfor reachability analysis. part ii: Internal approximations box-valued constraints”. Optimization Methods and Software 17.2(2002).

[28] M. R. Greenstreet and I. Mitchell. “Reachability analysisusing polygonal projections”. International Workshop onHybrid Systems: Computation and Control. Springer. 1999.

[29] J. N. Maidens, S. Kaynama, I. M. Mitchell, M. M. Oishi, andG. A. Dumont. “Lagrangian methods for approximating theviability kernel in high-dimensional systems”. Automatica49.7 (2013).

[30] M. Chen, S. L. Herbert, M. S. Vashishtha, S. Bansal, andC. J. Tomlin. “Decomposition of reachable sets and tubesfor a class of nonlinear systems”. Transactions on AutomaticControl 63.11 (2018).

[31] D. P. Bertsekas and J. N. Tsitsiklis. Neuro-dynamic program-ming. Athena Scientific Belmont, MA, 1996.

[32] Z. Wang, R. Spica, and M. Schwager. “Game TheoreticMotion Planning for Multi-robot Racing”. Distributed Au-tonomous Robotic Systems. Springer, 2019.

[33] M. Chen, J. F. Fisac, S. Sastry, and C. J. Tomlin.“Safe sequential path planning of multi-vehicle systems viadouble-obstacle Hamilton-Jacobi-Isaacs variational inequal-ity”. 2015 European Control Conference (ECC). IEEE. 2015.

[34] Z. Zhou, R. Takei, H. Huang, and C. J. Tomlin. “A general,open-loop formulation for reach-avoid games”. 51st Confer-ence on Decision and Control (CDC). IEEE. 2012.

[35] S.-Y. Liu, Z. Zhou, C. Tomlin, and J. K. Hedrick. “Evasionof a team of dubins vehicles from a hidden pursuer”. Inter-national Conference on Robotics and Automation (ICRA).IEEE. 2014.

[36] E. Todorov and W. Li. “A generalized iterative LQG methodfor locally-optimal feedback control of constrained nonlinearstochastic systems”. American Control Conference (ACC).IEEE. 2005.

[37] D. Mayne. “A second-order gradient method for determiningoptimal trajectories of non-linear discrete-time systems”.International Journal of Control 3.1 (1966).

[38] D. H. Jacobson and D. Q. Mayne. “Differential dynamicprogramming” (1970).

[39] J. van den Berg. “Iterated LQR smoothing for locally-optimalfeedback control of systems with non-linear dynamics andnon-quadratic cost”. American Control Conference (ACC).IEEE. 2014.

[40] H. Mukai, A. Tanikawa, I. Tunay, I. Katz, H. Schattler,P. Rinaldi, I. Ozcan, G. Wang, L. Yang, and Y. Sawada.“Sequential linear quadratic method for differential games”.Proc. 2nd DARPA-JFACC Symposium on Advances in En-terprise Control. Citeseer. 2000.

[41] A. Tanikawa, H. Mukai, and M. Xu. “Local Convergence ofthe Sequential Quadratic Method for Differential Games”.Transactions of the Institute of Systems, Control and Infor-mation Engineers 25.12 (2012).

[42] E. V. Mazumdar, M. I. Jordan, and S. S. Sastry. “On findinglocal nash equilibria (and only local nash equilibria) in zero-sum games”. arXiv preprint arXiv:1901.00838 (2019).

[43] M. Wang, Z. Wang, J. Talbot, J. C. Gerdes, and M. Schwager.“Game Theoretic Planning for Self-Driving Cars in Compet-itive Scenarios”. Robotics: Science & Systems. 2019.

[44] L. J. Ratliff, S. A. Burden, and S. S. Sastry. “On the char-acterization of local Nash equilibria in continuous games”.Transactions on Automatic Control 61.8 (2016).

[45] M. Green and D. J. Limebeer. Linear robust control. CourierCorporation, 2012.

[46] Y. Tassa, N. Mansard, and E. Todorov. “Control-limited dif-ferential dynamic programming”. International Conferenceon Robotics and Automation (ICRA). IEEE. 2014.


Recommended