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

Continuous time approximation of Nash equilibria

Romeo Awi111Department of Mathematics, Hampton University. , Ryan Hynd222Department of Mathematics, University of Pennsylvania.  , and Henok Mawi333Department of Mathematics, Howard University.
Abstract

We consider the problem of approximating Nash equilibria of NN functions f1,,fNf_{1},\dots,f_{N} of NN variables. In particular, we deduce conditions under which systems of the form

u˙j(t)=xjfj(u(t))\dot{u}_{j}(t)=-\nabla_{x_{j}}f_{j}(u(t))

(j=1,,N)(j=1,\dots,N) are well posed and in which the large time limits of their solutions u(t)=(u1(t),,uN(t))u(t)=(u_{1}(t),\dots,u_{N}(t)) are Nash equilibria for f1,,fNf_{1},\dots,f_{N}. To this end, we will invoke the theory of maximal monotone operators. We will also identify an application of these ideas in game theory and show how to approximate equilibria in some game theoretic problems in function spaces.

1 Introduction

Let us first recall the notion of a Nash equilibrium. Consider a collection of NN sets X1,,XNX_{1},\dots,X_{N} and define

X:=X1××XN.X:=X_{1}\times\dots\times X_{N}.

For a given x=(x1,,xN)Xx=(x_{1},\dots,x_{N})\in X, j{1,,N}j\in\{1,\dots,N\}, and yjXjy_{j}\in X_{j}, we will use the notation (yj,xj)(y_{j},x_{-j}) for the point in XX in which yjy_{j} replaces xjx_{j} in the coordinates of xx. That is,

(yj,xj):=(x1,,xj1,yj,xj+1,,xN).(y_{j},x_{-j}):=(x_{1},\dots,x_{j-1},y_{j},x_{j+1},\dots,x_{N}).

A collection of functions f1,,fN:Xf_{1},\dots,f_{N}:X\rightarrow\mathbb{R} has a Nash equilibrium at xXx\in X provided

fj(x)fj(yj,xj)f_{j}(x)\leq f_{j}(y_{j},x_{-j})

for all yjXjy_{j}\in X_{j} and j=1,,Nj=1,\dots,N.

Nash recognized that an equilibrium is a fixed point of the set valued mapping

XxargminyX{j=1Nfj(yj,xj)}.X\ni x\mapsto\operatorname*{arg\,min}_{y\in X}\left\{\sum^{N}_{j=1}f_{j}(y_{j},x_{-j})\right\}.

In particular, he applied Kakutani’s fixed point theorem [19] to show that if X1,,XNX_{1},\dots,X_{N} are nonempty, convex, compact subsets of Euclidean space and each f1,,fNf_{1},\dots,f_{N} is multilinear, then f1,,fNf_{1},\dots,f_{N} has a Nash equilibrium at some xXx\in X [24, 25]. Nash was interested in multilinear f1,,fNf_{1},\dots,f_{N} as they correspond to the expected cost of players assuming mixed strategies in noncooperative games. More generally, his existence theorem holds if each fj:Xf_{j}:X\rightarrow\mathbb{R} is continuous and

Xyj=1Nfj(yj,xj) is convex for each xXX\ni y\mapsto\sum^{N}_{j=1}f_{j}(y_{j},x_{-j})\text{ is convex for each $x\in X$. } (1.1)

As the existence of a Nash equilibrium is due to a nonconstructive fixed point theorem, it seems unlikely that there are good ways to approximate these points. Indeed, it has been established that the approximation of Nash equilibria is computationally challenging [9, 10, 21]. Nevertheless, we contend that there is a nontrivial class of functions f1,,fNf_{1},\dots,f_{N} for which the approximation of Nash equilibria is at least theoretically feasible. We will explain below that a sufficient condition for the approximation of Nash equilibria is that f1,,fNf_{1},\dots,f_{N} satisfy

j=1N(xjfj(x)xjfj(y))(xjyj)0\sum^{N}_{j=1}(\nabla_{x_{j}}f_{j}(x)-\nabla_{x_{j}}f_{j}(y))\cdot(x_{j}-y_{j})\geq 0 (1.2)

for each x,yXx,y\in X. Here we are considering each XjX_{j} as a subset of a Euclidean space and use ‘\cdot’ to denote the dot products on any of these spaces.

Our approach starts with the observation that if (1.1) holds and X1,,XNX_{1},\dots,X_{N} are convex, then xx is a Nash equilibrium if and only if

j=1Nxjfj(x)(yjxj)0\sum^{N}_{j=1}\nabla_{x_{j}}f_{j}(x)\cdot(y_{j}-x_{j})\geq 0 (1.3)

for each yXy\in X. As a result, it is natural to consider the differential inequalities

j=1N(u˙j(t)+xjfj(u(t)))(yjuj(t))0\sum^{N}_{j=1}(\dot{u}_{j}(t)+\nabla_{x_{j}}f_{j}(u(t)))\cdot(y_{j}-u_{j}(t))\geq 0 (1.4)

for t0t\geq 0 and yXy\in X. Here

u:[0,)X;t(u1(t),,uN(t))u:[0,\infty)\rightarrow X;t\mapsto(u_{1}(t),\dots,u_{N}(t))

is the unknown. We will below argue that for appropriately chosen initial condition u(0)Xu(0)\in X, the Cesàro mean of uu

t1t0tu(s)𝑑st\mapsto\frac{1}{t}\int^{t}_{0}u(s)ds (1.5)

will converge to a Nash equilibrium of f1,,fNf_{1},\dots,f_{N} as tt\rightarrow\infty provided that (1.2) holds.

1.1 A few concrete examples

An elementary example which illustrates this approach may be observed when X1=X2=[1,1]X_{1}=X_{2}=[-1,1], X=[1,1]2X=[-1,1]^{2},

f1(x1,x2)=x1x2,andf2(x1,x2)=x1x2.f_{1}(x_{1},x_{2})=x_{1}x_{2},\;\;\text{and}\;\;f_{2}(x_{1},x_{2})=-x_{1}x_{2}.

It is not hard to check that the origin (0,0)(0,0) is the unique Nash equilibrium of f1,f2f_{1},f_{2} and that f1,f2f_{1},f_{2} satisfy (1.2) (with equality holding) for each x,y[1,1]2x,y\in[-1,1]^{2}. We now seek an absolutely continuous path

u:[0,)[1,1]2u:[0,\infty)\rightarrow[-1,1]^{2}

such that

(u˙1(t)+u2(t))(y1u1(t))+(u˙2(t)u1(t))(y2u2(t))0(\dot{u}_{1}(t)+u_{2}(t))(y_{1}-u_{1}(t))+(\dot{u}_{2}(t)-u_{1}(t))(y_{2}-u_{2}(t))\geq 0 (1.6)

holds for almost every t0t\geq 0 and each (y1,y2)[1,1](y_{1},y_{2})\in[-1,1].

It turns out that if

u1(0)2+u2(0)21,u_{1}(0)^{2}+u_{2}(0)^{2}\leq 1,

then the unique solution of these differential inequalities is

{u1(t)=u1(0)cos(t)u2(0)sin(t)u2(t)=u2(0)cos(t)+u1(0)sin(t).\begin{cases}\displaystyle u_{1}(t)=u_{1}(0)\cos(t)-u_{2}(0)\sin(t)\\ \\ \displaystyle u_{2}(t)=u_{2}(0)\cos(t)+u_{1}(0)\sin(t).\end{cases}

This solution simply parametrizes the circle of radius u1(0)2+u2(0)2\sqrt{u_{1}(0)^{2}+u_{2}(0)^{2}}. In particular,

limt1t0tu(s)𝑑s=(0,0).\lim_{t\rightarrow\infty}\frac{1}{t}\int^{t}_{0}u(s)ds=(0,0).

For general initial conditions (u1(0),u2(0))[1,1]2(u_{1}(0),u_{2}(0))\in[-1,1]^{2}, it can be shown that u1(s)2+u2(s)21u_{1}(s)^{2}+u_{2}(s)^{2}\leq 1 in a finite time s0s\geq 0. Refer to Figure 1 for a schematic. It then follows that the same limit above holds for this solution.

Refer to caption
Figure 1: Plot of a solution uu of (1.6) where the initial position of uu is indicated in red. This solution starts out on a circle centered at the origin traversing it counterclockwise until it hits the boundary line segment (x1=1,1x21)(x_{1}=1,-1\leq x_{2}\leq 1). Then it proceeds upwards along this boundary line segment until it arrives at some time ss at the orange marker which is located at (1,0)(1,0). For times tst\geq s, the position u(t)u(t) as shown in purple remains on the unit circle traversing it counterclockwise.

The method we present can be extended in certain cases when X1,,XNX_{1},\dots,X_{N} are not compact. For example, suppose X1=X2=X_{1}=X_{2}=\mathbb{R},

f1(x1,x2)=12x12x1x2x1,andf2(x1,x2)=12x22+x1x22x2.f_{1}(x_{1},x_{2})=\frac{1}{2}x_{1}^{2}-x_{1}x_{2}-x_{1},\;\;\text{and}\;\;f_{2}(x_{1},x_{2})=\frac{1}{2}x_{2}^{2}+x_{1}x_{2}-2x_{2}.

It is easy to check that f1,f2f_{1},f_{2} satisfy (1.2). Furthermore,

{x1f1(x1,x2)=x1x21=0x2f2(x1,x2)=x2+x12=0\begin{cases}\partial_{x_{1}}f_{1}(x_{1},x_{2})=x_{1}-x_{2}-1=0\\ \\ \partial_{x_{2}}f_{2}(x_{1},x_{2})=x_{2}+x_{1}-2=0\end{cases}

provided

x1=32andx2=12.x_{1}=\frac{3}{2}\quad\text{and}\quad x_{2}=\frac{1}{2}.

As a result, this is the unique Nash equilibrium of f1,f2f_{1},f_{2}.

We note that the solution of

{u˙1(t)=(u1(t)u2(t)1)u˙2(t)=(u2(t)+u1(t)2)\begin{cases}\dot{u}_{1}(t)=-(u_{1}(t)-u_{2}(t)-1)\\ \\ \dot{u}_{2}(t)=-(u_{2}(t)+u_{1}(t)-2)\end{cases}

is given by

{u1(t)=32+(u1(0)32)etcos(t)+(u2(0)12)etsin(t)u2(t)=12+(u2(0)12)etcos(t)+(32u1(0))etsin(t).\begin{cases}\displaystyle u_{1}(t)=\frac{3}{2}+\left(u_{1}(0)-\frac{3}{2}\right)e^{-t}\cos(t)+\left(u_{2}(0)-\frac{1}{2}\right)e^{-t}\sin(t)\\ \\ \displaystyle u_{2}(t)=\frac{1}{2}+\left(u_{2}(0)-\frac{1}{2}\right)e^{-t}\cos(t)+\left(\frac{3}{2}-u_{1}(0)\right)e^{-t}\sin(t).\end{cases}

It is now clear that

limt(u1(t),u2(t))=(32,12).\lim_{t\rightarrow\infty}(u_{1}(t),u_{2}(t))=\left(\frac{3}{2},\frac{1}{2}\right).

In particular, we do not need to employ the Cesàro mean of uu in order to approximate the Nash equilibrium of f1,f2f_{1},f_{2}.

1.2 A general setting

It just so happens that our method does not rely on the spaces X1,,XNX_{1},\dots,X_{N} being finite dimensional. Consequently, we will consider the following version of our approximation problem. Let VV be a reflexive Banach space over \mathbb{R} with norm \|\cdot\| and continuous dual VV^{*}. We further suppose

X1,,XNVare closed and convex with nonempty interiorsX_{1},\dots,X_{N}\subset V\;\text{are closed and convex with nonempty interiors} (1.7)

and consider NN functions

f1,,fN:Xf_{1},\dots,f_{N}:X\rightarrow\mathbb{R} (1.8)

which satisfy

{f1,,fNare weakly lowersemicontinuous,Xyj=1Nfj(yj,xj)is convex for each xX,Xxj=1Nfj(yj,xj)is weakly continuous for each yX, andj=1N(xjfj(x)xjfj(y),xjyj)0for each x,yX.\begin{cases}f_{1},\dots,f_{N}\;\text{are weakly lowersemicontinuous},\\ \\ \displaystyle X\ni y\mapsto\sum^{N}_{j=1}f_{j}(y_{j},x_{-j})\;\text{is convex for each $\displaystyle x\in X$,}\\ \\ X\ni x\mapsto\displaystyle\sum^{N}_{j=1}f_{j}(y_{j},x_{-j})\;\text{is weakly continuous for each $y\in X$, and}\\ \\ \displaystyle\sum^{N}_{j=1}(\partial_{x_{j}}f_{j}(x)-\partial_{x_{j}}f_{j}(y),x_{j}-y_{j})\geq 0\;\;\text{for each $x,y\in X$}.\end{cases} (1.9)

In the last condition listed above, we mean

j=1N(pjqj,xjyj)0\sum^{N}_{j=1}(p_{j}-q_{j},x_{j}-y_{j})\geq 0

for each

pjxjfj(x):={ζV:fj(z,xj)fj(x)+(ζ,zxj),zXj}\displaystyle p_{j}\in\partial_{x_{j}}f_{j}(x):=\left\{\zeta\in V^{*}:f_{j}(z,x_{-j})\geq f_{j}(x)+(\zeta,z-x_{j}),\;z\in X_{j}\right\}

and qjxjfj(y)q_{j}\in\partial_{x_{j}}f_{j}(y) for j=1,,Nj=1,\dots,N.

When VV is finite dimensional, we naturally identify VV and VV^{*} with Euclidean space of the same dimension. Alternatively, when VV is infinite dimensional, we will suppose there is a real Hilbert space HH for which

VHVV\subset H\subset V^{*}

and VHV\subset H is continuously embedded. It is with respect to this space HH in which we consider the following initial value problem: for a given u0Xu^{0}\in X, find an absolutely continuous u:[0,)HNu:[0,\infty)\rightarrow H^{N} such that

{u(0)=u0,andj=1N(u˙j(t)+xjfj(u(t)),yjuj(t))0for a.e. t0 and each yX.\begin{cases}u(0)=u^{0},\;\text{and}\\ \\ \displaystyle\sum^{N}_{j=1}(\dot{u}_{j}(t)+\partial_{x_{j}}f_{j}(u(t)),y_{j}-u_{j}(t))\geq 0\;\text{for a.e. $t\geq 0$ and each $y\in X$.}\end{cases} (1.10)

Our central result involving this initial value problem is as follows. In this statement, we will make use of the set

𝒟={xX:j=1N(xjfj(x)+nXj(xj))H}.{\cal D}=\left\{x\in X:\bigcap^{N}_{j=1}\left(\partial_{x_{j}}f_{j}(x)+n_{X_{j}}(x_{j})\right)\cap H\neq\emptyset\right\}. (1.11)

Here nXj:V2Vn_{X_{j}}:V\rightarrow 2^{V^{*}} is the normal cone of XjX_{j} defined as

nXj(z):={ζV:(ζ,yz)0all yXj}n_{X_{j}}(z):=\{\zeta\in V^{*}:(\zeta,y-z)\leq 0\;\text{all $y\in X_{j}$}\} (1.12)

for zVz\in V (j=1,,N)(j=1,\dots,N).

Theorem 1.1.

Assume f1,,fNf_{1},\dots,f_{N} satisfy (1.9).

  1. (i)(i)

    For any u0𝒟u^{0}\in{\cal D} there is a unique Lipschitz continuous u:[0,)HNu:[0,\infty)\rightarrow H^{N} satisfying u(t)𝒟u(t)\in{\cal D} for each t0t\geq 0 and the initial value problem (1.10).

  2. (ii)(ii)

    If f1,,fNf_{1},\dots,f_{N} has a Nash equilibrium, then

    1t0tu(s)𝑑s\frac{1}{t}\int^{t}_{0}u(s)ds (1.13)

    converges weakly in HH to a Nash equilibrium of f1,,fNf_{1},\dots,f_{N} as tt\rightarrow\infty.

We note that if X1==XN=VX_{1}=\cdots=X_{N}=V, then nX1nXN{0}n_{X_{1}}\equiv\cdots\equiv n_{X_{N}}\equiv\{0\}. In this case

𝒟={xX:j=1N(xjfj(x)H)},{\cal D}=\left\{x\in X:\bigcap^{N}_{j=1}\left(\partial_{x_{j}}f_{j}(x)\cap H\right)\neq\emptyset\right\}, (1.14)

and the initial value problem reduces to

{u(0)=u0,andu˙j(t)+xjfj(u(t))0for a.e. t0 and each j=1,,N.\begin{cases}u(0)=u^{0},\;\text{and}\\ \\ \dot{u}_{j}(t)+\partial_{x_{j}}f_{j}(u(t))\ni 0\;\text{for a.e. $t\geq 0$ and each $j=1,\dots,N$}.\end{cases} (1.15)

We will also explain that if in addition (x1f1,,xNfN)(\partial_{x_{1}}f_{1},\dots,\partial_{x_{N}}f_{N}) is single-valued, everywhere-defined, monotone, and hemicontinuous we can obtain the same result as above without assuming (1.9). This follows directly from a theorem of Browder and Minty.

We will phrase our initial value problem (1.10) as the evolution generated by a monotone operator on HNH^{N} with domain 𝒟{\cal D}. According to pioneering work of Brézis [5], the crucial task will be to verify the maximality of this operator. Regarding the large time behavior of solutions, the convergence to equilibria of maximal monotone operators by the Cesàro means of semigroups they generate originates in the work of Brézis and Baillon [3]. We also refer the reader to Chapter 3 of [2] which gives a detailed discussion of this phenomenon.

Of course, the asymptotic statements we made above are predicated on the existence of a Nash equilibrium. This is only guaranteed to be the case if XX is weakly compact or if there is some θ>0\theta>0 such that

j=1N(xjfj(x)xjfj(y),xjyj)θxy2\displaystyle\sum^{N}_{j=1}(\partial_{x_{j}}f_{j}(x)-\partial_{x_{j}}f_{j}(y),x_{j}-y_{j})\geq\theta\|x-y\|^{2} (1.16)

for each x,yXx,y\in X. When such coercivity holds, f1,,fNf_{1},\dots,f_{N} has a unique Nash equilibrium and solutions of the initial value problem (1.10) converge exponentially fast in HNH^{N} to this equilibrium point.

We will prove Theorem 1.1 in the following section. Then we will apply this result to flows in Euclidean spaces of the form (1.4) in section 3. Finally, we will consider examples in Lebesgue and Sobolev spaces in section 4. A prototypical collection of functionals we will study is

fj(v)=Ω12|vj|2+Fj(v)dxf_{j}(v)=\int_{\Omega}\frac{1}{2}|\nabla v_{j}|^{2}+F_{j}(v)dx (1.17)

for j=1,,Nj=1,\dots,N, where v=(v1,,vN)H01(Ω)Nv=(v_{1},\dots,v_{N})\in H_{0}^{1}(\Omega)^{N} and Ωd\Omega\subset\mathbb{R}^{d} is a bounded domain. For this particular example, we will argue that if F1,,FN:NF_{1},\dots,F_{N}:\mathbb{R}^{N}\rightarrow\mathbb{R} satisfy suitable growth and monotonicity conditions, then f1,,fNf_{1},\dots,f_{N} has a unique Nash equilibrium which can be approximated by solutions u1,,uN:Ω×[0,)u_{1},\dots,u_{N}:\Omega\times[0,\infty)\rightarrow\mathbb{R} of the parabolic initial/boundary value problem

{tuj=ΔujzjFj(u)inΩ×(0,)uj=0onΩ×(0,)uj=uj0onΩ×{0}\begin{cases}\partial_{t}u_{j}=\Delta u_{j}-\partial_{z_{j}}F_{j}(u)\;\;&\text{in}\;\Omega\times(0,\infty)\\ \;\;\;u_{j}=0\;\;&\text{on}\;\partial\Omega\times(0,\infty)\\ \;\;\;u_{j}=u^{0}_{j}\;\;&\text{on}\;\Omega\times\{0\}\end{cases} (1.18)

for appropriately chosen initial conditions u10,,uN0:Ωu^{0}_{1},\dots,u^{0}_{N}:\Omega\rightarrow\mathbb{R}.

We also remark that in finite dimensions, similar results were known to Flåm [15]. In particular, Flåm seems to be the first author to identify the important role of the monotonicity condition (1.2) in approximating Nash equilibrium. However, Rosen appears to be the first to consider using equations such as (1.4) to approximate equilibrium points [26]. In addition, we would like to point out that the finite dimensional variational inequality (1.3) and its relation to Nash equilibrium is studied in depth in the monograph by Facchinei and Pang [13]. In infinite dimensions, there have also been several recent works which discuss discrete time approximations for variational inequalities that correspond to Nash equilibrium [7, 6, 1, 11, 18].

Acknowledgements: This material is based upon work supported by the National Science Foundation under Grants DMS-1440140, DMS-1554130, and HRD-1700236, the National Security Agency under Grant No. H98230-20-1-0015, and the Sloan Foundation under Grant No. G-2020-12602 while the authors participated in a program hosted by the Mathematical Sciences Research Institute in Berkeley, California, during the summer of 2020.

2 Approximation theorem

In this section, we will briefly recall the notion of a maximal monotone operator on a Hilbert space and state a few key results for these operators. Then we will apply these results to prove Theorem 1.1 and a few related corollaries.

2.1 Maximal monotone operators on a Hilbert space

Let HH be Hilbert space with inner product ,\langle\cdot,\cdot\rangle and norm |||\cdot|. We will denote 2H2^{H} as the power set or collection of all subsets of HH. We recall that B:H2HB:H\rightarrow 2^{H} is monotone if

pq,xy0\langle p-q,x-y\rangle\geq 0

for all x,yHx,y\in H, pBxp\in Bx, and qByq\in By. Moreover, BB is maximally monotone if the only monotone C:H2HC:H\rightarrow 2^{H} such that BxCxBx\subset Cx for all xHx\in H is C=BC=B. Minty’s well known maximality criterion is as follows [22].

Theorem (Minty’s Lemma).

A monotone operator B:H2HB:H\rightarrow 2^{H} is maximal if and only if for each yHy\in H, there is a unique xHx\in H such that

x+Bxy.x+Bx\ni y.

Let us now recall a fundamental theorem for the initial value problem

{u˙(t)+Bu(t)0,a.e.t0u(0)=u0.\begin{cases}\dot{u}(t)+Bu(t)\ni 0,\;\text{a.e.}\;t\geq 0\\ u(0)=u^{0}.\end{cases} (2.1)

As discussed in Chapter II of [5] and Chapter 3 of [2], the initial value problem is well-posed provided that BB is maximally monotone and u0u^{0} is an element of the domain of BB

D(B)={xH:Bx}.D(B)=\{x\in H:Bx\neq\emptyset\}.
Theorem (Brézis’s Theorem).

Suppose B:H2HB:H\rightarrow 2^{H} is maximally monotone and u0D(B)u^{0}\in D(B). Then there is a unique Lipschitz continuous

u:[0,)Hu:[0,\infty)\rightarrow H

solution of (2.1) such that u(t)D(B)u(t)\in D(B) for all t0t\geq 0. Moreover,

|u˙(t)|min{|p|:pBu0}|\dot{u}(t)|\leq\min\{|p|:p\in Bu^{0}\} (2.2)

for almost every t0t\geq 0.

Remark 2.1.

The full statement of Brézis’s Theorem (Theorem 3.1 of [2]) is more extensive than what is written above.

Remark 2.2.

We note that the right hand side of (2.2) is finite as the images of maximal monotone operators are closed and convex subsets of HH.

We also note that the operator BB generates a contraction. Let uu be a path as described in Brézis’ Theorem, and suppose v:[0,)Hv:[0,\infty)\rightarrow H is any other Lipschitz continuous path with v(t)D(B)v(t)\in D(B) for t0t\geq 0 and

v˙(t)+Bv(t)0,a.e.t0.\dot{v}(t)+Bv(t)\ni 0,\;\text{a.e.}\;t\geq 0.

Then

|u(t)v(t)||u(0)v(0)|,t0.|u(t)-v(t)|\leq|u(0)-v(0)|,\quad t\geq 0. (2.3)

If, in addition, there is some λ>0\lambda>0 such that

pq,xyλ|xy|2\langle p-q,x-y\rangle\geq\lambda|x-y|^{2} (2.4)

for all x,yHx,y\in H, pBxp\in Bx, and qByq\in By, then (2.3) can be improved to

|u(t)v(t)|eλt|u(0)v(0)|,t0.|u(t)-v(t)|\leq e^{-\lambda t}|u(0)-v(0)|,\quad t\geq 0. (2.5)

Remarkably, it is also possible to use solutions of the initial value problem (as described in Brézis’s Theorem) to approximate equilibria of BB. These are points yHy\in H such that

By0.By\ni 0.

The following theorem was proved in [3].

Theorem (The Baillon-Brézis Theorem).

Suppose BB has an equilibrium point and uu is a solution of the initial value problem (2.1). Then

1t0tu(s)𝑑s\frac{1}{t}\int^{t}_{0}u(s)ds

converges weakly in HH to some equilibrium point of BB.

Remark 2.3.

If BB satisfies (2.4), it is plain to see that By0By\ni 0 can have at most one solution. In this case,

|u(t)y|eλt|u(0)y|,t0|u(t)-y|\leq e^{-\lambda t}|u(0)-y|,\quad t\geq 0

for any solution of the initial value problem (2.1). This follows from (2.5) as v(t)=yv(t)=y is a solution of the initial value problem with v(0)=yv(0)=y.

2.2 Proof of Theorem 1.1

Let us now suppose the spaces VV, X1,,XNX_{1},\dots,X_{N}, and HH are as described in the subsection 1.2 of the introduction. We will further assume f1,,fN:Xf_{1},\dots,f_{N}:X\rightarrow\mathbb{R} are given functions which satisfy (1.9). An elementary but important observation we will use is that a given yHy\in H induces a linear form in VV^{*} by the formula

(y,x):=y,x,(xV).(y,x):=\langle y,x\rangle,\quad(x\in V).

That this linear form is continuous is due to the continuity of the embedding VHV\subset H. For convenience, we will use \|\cdot\| to denote both the norm on VV and the associated norm on VNV^{N}. Namely, we’ll write

x=(j=1Nxj2)1/2\|x\|=\left(\sum^{N}_{j=1}\|x_{j}\|^{2}\right)^{1/2}

for x=(x1,,xN)VNx=(x_{1},\dots,x_{N})\in V^{N}. We will also use this convention for the inner product ,\langle\cdot,\cdot\rangle and norm |||\cdot| on HH and on HNH^{N} and the induced norm \|\cdot\|_{*} on VV^{*} and on (V)N(V^{*})^{N}.

Let us specify A:HN2HNA:H^{N}\rightarrow 2^{H^{N}} via

Ax={(x1f1(x)+nX1(x1))H××(xNfN(x)+nXN(xN))H,x𝒟,x𝒟Ax=\begin{cases}\left(\partial_{x_{1}}f_{1}(x)+n_{X_{1}}(x_{1})\right)\cap H\times\dots\times\left(\partial_{x_{N}}f_{N}(x)+n_{X_{N}}(x_{N})\right)\cap H,&x\in{\cal D}\\ \emptyset,&x\not\in{\cal D}\end{cases} (2.6)

for xHNx\in H^{N}. Here 𝒟{\cal D} is defined in (1.11).

Lemma 2.4.

The domain of AA is 𝒟{\cal D}, and AA is monotone. Moreover, an absolutely continuous u:[0,)HNu:[0,\infty)\rightarrow H^{N} solves

{u˙(t)+Au(t)0,a.e.t0u(0)=u0\begin{cases}\dot{u}(t)+Au(t)\ni 0,\;\text{a.e.}\;t\geq 0\\ u(0)=u^{0}\end{cases} (2.7)

if and only it solves (1.10).

Proof.

By definition, the domain of AA is 𝒟{\cal D}. Suppose x,y𝒟x,y\in{\cal D},

ζj(xjfj(x)+nXj(xj))H,andξj(xjfj(y)+nXj(yj))H\zeta_{j}\in\left(\partial_{x_{j}}f_{j}(x)+n_{X_{j}}(x_{j})\right)\cap H,\;\text{and}\;\xi_{j}\in\left(\partial_{x_{j}}f_{j}(y)+n_{X_{j}}(y_{j})\right)\cap H

for j=1,,Nj=1,\dots,N. In view of the last condition listed in (1.9) and the fact that each nXjn_{X_{j}} is monotone,

j=1Nζjξj,xjyj\displaystyle\sum^{N}_{j=1}\langle\zeta_{j}-\xi_{j},x_{j}-y_{j}\rangle =j=1N(ζjξj,xjyj)0.\displaystyle=\sum^{N}_{j=1}(\zeta_{j}-\xi_{j},x_{j}-y_{j})\geq 0. (2.8)

Thus, AA is monotone.

Note that if uu solves (2.7), then

u˙j(t)(xjfj(u(t))+nXj(uj(t)))Hxjfj(u(t))+nXj(uj(t))-\dot{u}_{j}(t)\in(\partial_{x_{j}}f_{j}(u(t))+n_{X_{j}}(u_{j}(t)))\cap H\subset\partial_{x_{j}}f_{j}(u(t))+n_{X_{j}}(u_{j}(t))

for almost every t0t\geq 0 and each j=1,,Nj=1,\dots,N. It now follows from the definition of the normal cone that uu is a solution of (1.10). Conversely, if uu solves (1.10) then

u˙j(t)xjfj(u(t))+nXj(uj(t))-\dot{u}_{j}(t)\in\partial_{x_{j}}f_{j}(u(t))+n_{X_{j}}(u_{j}(t))

for almost every t0t\geq 0 and each j=1,,Nj=1,\dots,N. As uu is absolutely continuous, u˙(t)H-\dot{u}(t)\in H for almost every t0t\geq 0, and therefore, u˙(t)Au(t)-\dot{u}(t)\in Au(t) for almost every t0t\geq 0. ∎

In view of Brézis’ Theorem and the Baillon-Brézis Theorem, we can conclude Theorem 1.1 once we verify that AA is maximal. To this end, we will verify the hypotheses of Minty’s Lemma.

Proof of Theorem 1.1.

For a given yHNy\in H^{N}, it suffices to show there is xXx\in X such that

j=1N(xj+xjfj(x)yj,zjxj)0\displaystyle\sum^{N}_{j=1}(x_{j}+\partial_{x_{j}}f_{j}(x)-y_{j},z_{j}-x_{j})\geq 0 (2.9)

for each zXz\in X. In this case, yjxjHy_{j}-x_{j}\in H and

yjxjxjfj(x)+nXj(xj)y_{j}-x_{j}\in\partial_{x_{j}}f_{j}(x)+n_{X_{j}}(x_{j})

for j=1,,Nj=1,\dots,N so that x𝒟x\in{\cal D} and yxAxy-x\in Ax.

In order to solve (2.9), we will employ the auxiliary functions defined by

gj(x):=12xj2(yj,xj)+fj(x)g_{j}(x):=\displaystyle\frac{1}{2}\|x_{j}\|^{2}-(y_{j},x_{j})+f_{j}(x)

for xXx\in X and j=1,,Nj=1,\dots,N. In view of (1.9),

{g1,,gNare weakly lowersemicontinuous,Xyj=1Ngj(yj,xj)is convex for each xX,Xxj=1Ngj(yj,xj)is weakly continuous for each yX, andj=1N(xjgj(x)xjgj(y),xjyj)xy2for each x,yX.\begin{cases}g_{1},\dots,g_{N}\;\text{are weakly lowersemicontinuous},\\ \\ X\ni y\mapsto\displaystyle\sum^{N}_{j=1}g_{j}(y_{j},x_{-j})\;\text{is convex for each $\displaystyle x\in X$,}\\ \\ X\ni x\mapsto\displaystyle\sum^{N}_{j=1}g_{j}(y_{j},x_{-j})\;\text{is weakly continuous for each $y\in X$, and}\\ \\ \displaystyle\sum^{N}_{j=1}(\partial_{x_{j}}g_{j}(x)-\partial_{x_{j}}g_{j}(y),x_{j}-y_{j})\geq\|x-y\|^{2}\;\;\text{for each $x,y\in X$}.\end{cases} (2.10)

Note that xXx\in X is a Nash equilibrium of g1,,gNg_{1},\dots,g_{N} if and only if xx solves (2.9). Since we are assuming that each XjX_{j} has nonempty interior, this follows from the Pshenichnii–Rockafellar conditions for the minimum of a proper, lowersemicontinuous, convex function on a closed subset of a Banach space (Theorem 4.3.6 of [4]). Consequently, we now aim to show that g1,,gNg_{1},\dots,g_{N} has a Nash equilibrium.

For each r>0r>0, set

Xr:={xX:xr}.X^{r}:=\{x\in X:\|x\|\leq r\}.

It is clear that XrX^{r} is convex and weakly closed. And as each X1,,XNX_{1},\dots,X_{N} is nonempty, XrX^{r} is nonempty for all rr greater or equal to some fixed s>0s>0. Let us consider the map Φr:Xr2Xr\Phi^{r}:X^{r}\rightarrow 2^{X^{r}} specified as

Φr(x):=argminyXr{j=1Ngj(yj,xj)}\Phi^{r}(x):=\operatorname*{arg\,min}_{y\in X^{r}}\left\{\sum^{N}_{j=1}g_{j}(y_{j},x_{-j})\right\}

(xXr)(x\in X^{r}) for r>sr>s.

The first and second properties of g1,,gNg_{1},\dots,g_{N} listed in (2.10) imply that Φr(x)\Phi^{r}(x)\neq\emptyset and convex for any xXrx\in X^{r} and r>sr>s; and the first and third properties imply that the graph of Φr\Phi^{r} is weakly closed. Indeed let us suppose (xk)k,(yk)kX(x^{k})_{k\in\mathbb{N}},(y^{k})_{k\in\mathbb{N}}\subset X, xkxx^{k}\rightharpoonup x, ykyy^{k}\rightharpoonup y, and

j=1Ngj(yjk,xjk)j=1Ngj(zj,xjk)\sum^{N}_{j=1}g_{j}(y^{k}_{j},x^{k}_{-j})\leq\sum^{N}_{j=1}g_{j}(z_{j},x^{k}_{-j})

for each zXz\in X with zr\|z\|\leq r and all kk\in\mathbb{N}. We can then send kk\rightarrow\infty to get

j=1Ngj(yj,xj)j=1Ngj(zj,xj).\sum^{N}_{j=1}g_{j}(y_{j},x_{-j})\leq\sum^{N}_{j=1}g_{j}(z_{j},x_{-j}).

The Kakutani-Glicksberg-Fan theorem [19, 16, 14] then implies there is xrΦr(xr)x^{r}\in\Phi^{r}(x^{r}). In particular,

j=1Ngj(xr)j=1Ngj(zj,xjr)\sum^{N}_{j=1}g_{j}(x^{r})\leq\sum^{N}_{j=1}g_{j}(z_{j},x^{r}_{-j}) (2.11)

for each zXz\in X with zr\|z\|\leq r. It follows that there are pjrxjgj(xr)p^{r}_{j}\in\partial_{x_{j}}g_{j}(x^{r}) for each j=1,,Nj=1,\dots,N such that

j=1N(pjr,zjxjr)0\sum^{N}_{j=1}(p^{r}_{j},z_{j}-x^{r}_{j})\geq 0

for all zXz\in X with zr\|z\|\leq r. Since u0𝒟u^{0}\in{\cal D}, there are

wjxjgj(u0)w_{j}\in\partial_{x_{j}}g_{j}(u^{0})

for j=1,,Nj=1,\dots,N.

Note that the fourth listed property of g1,,gNg_{1},\dots,g_{N} in (2.10) gives

xru02\displaystyle\|x^{r}-u^{0}\|^{2} j=1N(pjrwj,xjruj0)\displaystyle\leq\sum^{N}_{j=1}(p^{r}_{j}-w_{j},x^{r}_{j}-u^{0}_{j})
j=1N(wj,xjruj0)\displaystyle\leq\sum^{N}_{j=1}(-w_{j},x^{r}_{j}-u^{0}_{j})
wxru0\displaystyle\leq\|w\|_{*}\|x^{r}-u^{0}\|

for all rr sufficiently large. As a result, there is a sequence (rk)k(r_{k})_{k\in\mathbb{N}} which increases to infinity such that (xrk)kX(x^{r_{k}})_{k\in\mathbb{N}}\subset X converges weakly to some xXx\in X. Upon sending r=rkr=r_{k}\rightarrow\infty in (2.11), we find that xx is the desired Nash equilibrium of g1,,gNg_{1},\dots,g_{N}. ∎

Corollary 2.5.

If XX is weakly compact, then f1,,fNf_{1},\dots,f_{N} has a Nash equilibrium.

Proof.

If XX is weakly compact, we can repeat the argument given in the proof above used to show g1,,gN:Xrg_{1},\dots,g_{N}:X^{r}\rightarrow\mathbb{R} has a Nash equilibrium. All we would need to do is to replace XrX^{r} with XX and g1,,gNg_{1},\dots,g_{N} with f1,,fNf_{1},\dots,f_{N}. ∎

Corollary 2.6.

Suppose there is θ>0\theta>0 such that

j=1N(xjfj(x)xjfj(y),xjyj)θxy2\displaystyle\sum^{N}_{j=1}(\partial_{x_{j}}f_{j}(x)-\partial_{x_{j}}f_{j}(y),x_{j}-y_{j})\geq\theta\|x-y\|^{2} (2.12)

for each x,yXx,y\in X. Then f1,,fNf_{1},\dots,f_{N} has a unique Nash equilibrium zXz\in X. Moreover, if u:[0,)HNu:[0,\infty)\rightarrow H^{N} is a solution of the initial value problem as described in Theorem 1.1, there is λ>0\lambda>0 such that

|u(t)z|eλt|u(0)z||u(t)-z|\leq e^{-\lambda t}|u(0)-z| (2.13)

for each t0t\geq 0.

Proof.

We can repeat the argument given in the proof of Theorem 1.1 that shows g1,,gNg_{1},\dots,g_{N} has a Nash equilibrium to conclude that f1,,fNf_{1},\dots,f_{N} has a Nash equilibrium.

If xx is a Nash equilibrium of f1,,fNf_{1},\dots,f_{N}, then

j=1N(xjfj(x),zjxj)0\sum^{N}_{j=1}(\partial_{x_{j}}f_{j}(x),z_{j}-x_{j})\geq 0

for each zXz\in X. Likewise,

j=1N(xjfj(y),wjyj)0\sum^{N}_{j=1}(\partial_{x_{j}}f_{j}(y),w_{j}-y_{j})\geq 0

for each wXw\in X if yy is another Nash equilibrium. Choosing z=yz=y and w=xw=x and adding these inequalities give

θxy2j=1N(xjfj(x)xjfj(y),xjyj)0.\theta\|x-y\|^{2}\leq\sum^{N}_{j=1}(\partial_{x_{j}}f_{j}(x)-\partial_{x_{j}}f_{j}(y),x_{j}-y_{j})\leq 0.

As VNHNV^{N}\subset H^{N} is continuously embedded, there is a constant C>0C>0 such that

|x|Cx,xVN.|x|\leq C\|x\|,\quad x\in V^{N}.

Setting

λ:=θC2,\lambda:=\frac{\theta}{C^{2}},

gives

AxAy,xyθxy2λ|xy|2\langle Ax-Ay,x-y\rangle\geq\theta\|x-y\|^{2}\geq\lambda|x-y|^{2}

for x,yD(A)x,y\in D(A). The inequality (2.13) now follows from Remark 2.3. ∎

We now consider the special case mentioned in the introduction. In the statement below, we will suppose that VV and HH are as above. However, we will not assume f1,,fNf_{1},\dots,f_{N} satisfy (1.9). We also note that the proof essentially follows from an observation made by Brézis in Remark 2.3.7 of [5].

Theorem 2.7.

Suppose f1,,fN:VNf_{1},\dots,f_{N}:V^{N}\rightarrow\mathbb{R} satisfy

{xjfj(x)={xjfj(x)}for each xVN,xjfj(x+ty)xjfj(x)as t0+ for each x,yVN, andj=1N(xjfj(x)xjfj(y),xjyj)0for each x,yVN,\begin{cases}\partial_{x_{j}}f_{j}(x)=\{\nabla_{x_{j}}f_{j}(x)\}\;\text{for each $x\in V^{N}$,}\\ \\ \nabla_{x_{j}}f_{j}(x+ty)\rightharpoonup\nabla_{x_{j}}f_{j}(x)\;\text{as $t\rightarrow 0^{+}$ for each $x,y\in V^{N}$, and}\\ \\ \displaystyle\sum^{N}_{j=1}(\nabla_{x_{j}}f_{j}(x)-\nabla_{x_{j}}f_{j}(y),x_{j}-y_{j})\geq 0\;\;\text{for each $x,y\in V^{N}$},\end{cases} (2.14)

and define

𝒟={xVN:x1f1(x),,xNfN(x)H}.{\cal D}=\{x\in V^{N}:\nabla_{x_{1}}f_{1}(x),\dots,\nabla_{x_{N}}f_{N}(x)\in H\}.

Then for each u0𝒟u^{0}\in{\cal D}, there is a unique Lipschitz continuous u:[0,)HNu:[0,\infty)\rightarrow H^{N} such that u(t)𝒟u(t)\in{\cal D} for each t0t\geq 0 and

{u(0)=u0,andu˙j(t)+xjfj(u(t))=0for a.e. t0 and each j=1,,N.\begin{cases}u(0)=u^{0},\;\text{and}\\ \\ \dot{u}_{j}(t)+\nabla_{x_{j}}f_{j}(u(t))=0\;\text{for a.e. $t\geq 0$ and each $j=1,\dots,N$}.\end{cases} (2.15)
Proof.

It suffices to show that A:HN2HNA:H^{N}\rightarrow 2^{H^{N}} defined via

Ax:={{(x1f1(x),,xNfN(x))},x𝒟,x𝒟Ax:=\begin{cases}\{(\nabla_{x_{1}}f_{1}(x),\dots,\nabla_{x_{N}}f_{N}(x))\},&x\in{\cal D}\\ \emptyset,&x\not\in{\cal D}\end{cases} (2.16)

(xHNx\in H^{N}) is maximal. To this end, we first consider the operator B:VN(V)NB:V^{N}\rightarrow(V^{*})^{N} defined via

Bx:=(x1+x1f1(x),,xN+xNfN(x))Bx:=(x_{1}+\nabla_{x_{1}}f_{1}(x),\dots,x_{N}+\nabla_{x_{N}}f_{N}(x))

for xVNx\in V^{N}. Notice that

B(x+ty)BxB(x+ty)\rightharpoonup Bx

and

(BxBy,xy)xy2(Bx-By,x-y)\geq\|x-y\|^{2}

for all x,yVNx,y\in V^{N}.

By a theorem due independently to Browder [8] and Minty [23], BB is surjective; see also Corollary 1.8 of [20]. In particular, if yHNy\in H^{N}, there is xVNx\in V^{N} such that

Bx=y.Bx=y.

That is,

xj+xjfj(x)=yjx_{j}+\nabla_{x_{j}}f_{j}(x)=y_{j}

for j=1,,Nj=1,\dots,N. Then xjfj(x)=yjxjH\nabla_{x_{j}}f_{j}(x)=y_{j}-x_{j}\in H for j=1,,Nj=1,\dots,N, which implies x𝒟x\in{\cal D} and

x+Axy.x+Ax\ni y.

Remark 2.8.

If f1,,fNf_{1},\dots,f_{N} additionally satisfy

j=1N(xjfj(x)xjfj(y),xjyj)θxy2\sum^{N}_{j=1}(\nabla_{x_{j}}f_{j}(x)-\nabla_{x_{j}}f_{j}(y),x_{j}-y_{j})\geq\theta\|x-y\|^{2}

(x,yVNx,y\in V^{N}), then f1,,fNf_{1},\dots,f_{N} has a unique Nash equilibrium and (2.13) holds for any u(0)𝒟u(0)\in{\cal D}.

3 Finite dimensional flows

In this section, we will study a few implications of Theorem 1.1 in the particular case

V=H=dV=H=\mathbb{R}^{d}

equipped with the standard Euclidean dot product and norm. As there is just one topology to consider, the statements we’ll make will be simpler than in the infinite dimensional setting. We will also identify a potential application to noncooperative games.

3.1 Compact domains

Let X=X1××XNX=X_{1}\times\cdots\times X_{N}, where X1,,XNdX_{1},\dots,X_{N}\subset\mathbb{R}^{d} are convex, compact subsets with nonempty interior. Further assume

{f1,,fN:Xare continuous,Xyj=1Nfj(yj,xj)is convex for each xX, andj=1N(xjfj(x)xjfj(y))(xjyj)0for each x,yX.\begin{cases}f_{1},\dots,f_{N}:X\rightarrow\mathbb{R}\;\text{are continuous},\\ \\ X\ni y\mapsto\displaystyle\sum^{N}_{j=1}f_{j}(y_{j},x_{-j})\;\text{is convex for each $\displaystyle x\in X$, and}\\ \\ \displaystyle\sum^{N}_{j=1}(\partial_{x_{j}}f_{j}(x)-\partial_{x_{j}}f_{j}(y))\cdot(x_{j}-y_{j})\geq 0\;\;\text{for each $x,y\in X$.}\end{cases} (3.1)

As before we set

𝒟={xX:j=1N(xjfj(x)+nXj(x))},{\cal D}=\left\{x\in X:\bigcap^{N}_{j=1}(\partial_{x_{j}}f_{j}(x)+n_{X_{j}}(x))\neq\emptyset\right\},

where

nXj(z)={ζd:ζ(yz)0for all yXj}n_{X_{j}}(z)=\{\zeta\in\mathbb{R}^{d}:\zeta\cdot(y-z)\leq 0\;\text{for all $y\in X_{j}$}\}

(j=1,,N(j=1,\dots,N). Theorem 1.1 implies the following.

Proposition 3.1.

For each u0𝒟u^{0}\in{\cal D}, there is a unique Lipschitz continuous u:[0,)Ndu:[0,\infty)\rightarrow\mathbb{R}^{Nd} with u(t)𝒟u(t)\in{\cal D} for t0t\geq 0 and which satisfies

{u(0)=u0,andj=1N(u˙j(t)+xjfj(u(t)))(yjuj(t))0for a.e. t0 and each yX.\begin{cases}u(0)=u^{0},\;\text{and}\\ \\ \displaystyle\sum^{N}_{j=1}(\dot{u}_{j}(t)+\partial_{x_{j}}f_{j}(u(t)))\cdot(y_{j}-u_{j}(t))\geq 0\;\text{for a.e. $t\geq 0$ and each $y\in X$}.\end{cases} (3.2)

Furthermore, the limit

limt1t0tu(s)𝑑s\lim_{t\rightarrow\infty}\frac{1}{t}\int^{t}_{0}u(s)ds

exists and equals a Nash equilibrium of f1,,fNf_{1},\dots,f_{N}.

Remark 3.2.

In view of Corollary 2.6, if

j=1N(xjfj(x)xjfj(y))(xjyj)θ|xy|2\displaystyle\sum^{N}_{j=1}(\partial_{x_{j}}f_{j}(x)-\partial_{x_{j}}f_{j}(y))\cdot(x_{j}-y_{j})\geq\theta|x-y|^{2}

for x,y,Xx,y,\in X, f1,,fNf_{1},\dots,f_{N} has a unique Nash equilibrium and the solution uu in the proposition above converges exponentially fast to this equilibrium point.

3.2 An application to game theory

Again suppose X1,,XNdX_{1},\dots,X_{N}\subset\mathbb{R}^{d} are convex, compact subsets with nonempty interior. Another interesting case to consider is when f1,,fN:Xf_{1},\dots,f_{N}:X\rightarrow\mathbb{R} are each NN-linear. That is,

Xjyjfj(yj,xj)is linearX_{j}\ni y_{j}\mapsto f_{j}(y_{j},x_{-j})\;\text{is linear} (3.3)

for each j=1,,Nj=1,\dots,N and xXx\in X. These are precisely the types of functions Nash considered in his celebrated work on noncooperative games [24, 25]. Note that (3.3) implies

yjxjfj(x)=fj(yj,xj)y_{j}\cdot\nabla_{x_{j}}f_{j}(x)=f_{j}(y_{j},x_{-j}) (3.4)

for yjXjy_{j}\in X_{j}, xNdx\in\mathbb{R}^{Nd}, and j=1,,Nj=1,\dots,N. Consequently,

j=1N(xjfj(x)xjfj(y))(xjyj)=j=1N(fj(x)fj(yj,xj)+fj(y)fj(xj,yj))0\displaystyle\sum^{N}_{j=1}\left(\nabla_{x_{j}}f_{j}(x)-\nabla_{x_{j}}f_{j}(y)\right)\cdot(x_{j}-y_{j})=\sum^{N}_{j=1}(f_{j}(x)-f_{j}(y_{j},x_{-j})+f_{j}(y)-f_{j}(x_{j},y_{-j}))\geq 0 (3.5)

if and only if

j=1N(fj(x)+fj(y))j=1N(fj(yj,xj)+fj(xj,yj)).\sum^{N}_{j=1}(f_{j}(x)+f_{j}(y))\geq\sum^{N}_{j=1}(f_{j}(y_{j},x_{-j})+f_{j}(x_{j},y_{-j})). (3.6)

With these assumptions, Theorem 1.1 implies the following statement.

Proposition 3.3.

Suppose f1,,fNf_{1},\dots,f_{N} are NN-linear and satisfy (3.6). Then for each u0Xu^{0}\in X, there is a unique Lipschitz u:[0,)Xu:[0,\infty)\rightarrow X such that

{u(0)=u0j=1N(u˙j(t)+xjfj(u(t)))(yjuj(t))0,for a.e. t0 and all yX.\begin{cases}u(0)=u^{0}\\ \\ \displaystyle\sum^{N}_{j=1}(\dot{u}_{j}(t)+\nabla_{x_{j}}f_{j}(u(t)))\cdot(y_{j}-u_{j}(t))\geq 0,\;\text{for a.e. $t\geq 0$ and all $y\in X$.}\end{cases} (3.7)

Moreover,

1t0tu(s)𝑑s\frac{1}{t}\int^{t}_{0}u(s)ds

converges as tt\rightarrow\infty to a Nash equilibrium of f1,,fNf_{1},\dots,f_{N}.

Remark 3.4.

Using (3.4), we can rewrite the inequality in (3.7) as

j=1N(u˙j(t)(yjuj(t))+fj(yj,uj(t))fj(u(t)))0.\sum^{N}_{j=1}(\dot{u}_{j}(t)\cdot(y_{j}-u_{j}(t))+f_{j}(y_{j},u_{-j}(t))-f_{j}(u(t)))\geq 0.

3.3 The whole space

Finally, we consider the case where f1,,fNf_{1},\dots,f_{N} are defined on the whole space Nd\mathbb{R}^{Nd}. We will assume

{f1,,fN are continuous,Ndyj=1Nfj(yj,xj) is convex for each xNd, andj=1N(xjfj(x)xjfj(y))(xjyj)0 for each x,yNd.\begin{cases}f_{1},\dots,f_{N}\text{ are continuous,}\\ \\ \displaystyle\mathbb{R}^{Nd}\ni y\mapsto\sum^{N}_{j=1}f_{j}(y_{j},x_{-j})\text{ is convex for each $x\in\mathbb{R}^{Nd}$, and}\\ \\ \displaystyle\sum^{N}_{j=1}(\partial_{x_{j}}f_{j}(x)-\partial_{x_{j}}f_{j}(y))\cdot(x_{j}-y_{j})\geq 0\text{ for each $x,y\in\mathbb{R}^{Nd}$}.\end{cases} (3.8)

Let us also set

𝒟:={xNd:j=1Nxjfj(x)}.{\cal D}:=\left\{x\in\mathbb{R}^{Nd}:\bigcap^{N}_{j=1}\partial_{x_{j}}f_{j}(x)\neq\emptyset\right\}.

Theorem 1.1 gives us the following proposition.

Proposition 3.5.

For each u0𝒟u^{0}\in{\cal D}, there is a unique Lipschitz continuous u:[0,)Ndu:[0,\infty)\rightarrow\mathbb{R}^{Nd} such that u(t)𝒟u(t)\in{\cal D} for each t0t\geq 0 and which satisfies

{u(0)=u0u˙j(t)+xjfj(u(t))0,for a.e. t0 and j=1,,N.\begin{cases}u(0)=u^{0}\\ \\ \;\dot{u}_{j}(t)+\partial_{x_{j}}f_{j}(u(t))\ni 0,\;\text{for a.e. $t\geq 0$ and $j=1,\dots,N$.}\end{cases} (3.9)

If f1,,fNf_{1},\dots,f_{N} has a Nash equilibrium, then

limt1t0tu(s)𝑑s\lim_{t\rightarrow\infty}\frac{1}{t}\int^{t}_{0}u(s)ds

exists and is also Nash equilibrium.

Remark 3.6.

If

j=1N(xjfj(x)xjfj(y))(xjyj)θ|xy|2\sum^{N}_{j=1}(\partial_{x_{j}}f_{j}(x)-\partial_{x_{j}}f_{j}(y))\cdot(x_{j}-y_{j})\geq\theta|x-y|^{2}

for all x,yNdx,y\in\mathbb{R}^{Nd} and some θ>0\theta>0, then f1,,fNf_{1},\dots,f_{N} has a unique Nash equilibrium at some zNdz\in\mathbb{R}^{Nd} and

|u(t)z|eθt|u0z|,t0.|u(t)-z|\leq e^{-\theta t}|u^{0}-z|,\quad t\geq 0.
Remark 3.7.

A basic example we had in mind when considering Proposition 3.5 was

fj(x)=k,=1N12Ak,jxkxf_{j}(x)=\sum^{N}_{k,\ell=1}\frac{1}{2}A^{j}_{k,\ell}x_{k}\cdot x_{\ell}

for x=(x1,,xN)Ndx=(x_{1},\dots,x_{N})\in\mathbb{R}^{Nd} and j=1,,Nj=1,\dots,N. Here Ak,jA^{j}_{k,\ell} are d×dd\times d symmetric matrices which additionally satisfy

Ak,j=A,kjA^{j}_{k,\ell}=A^{j}_{\ell,k}

(j,k,=1,,N)(j,k,\ell=1,\dots,N). The required monotonicity condition on f1,,fNf_{1},\dots,f_{N} is satisfied provided

j,k=1NAk,jjyjyk0\sum^{N}_{j,k=1}A^{j}_{k,j}y_{j}\cdot y_{k}\geq 0 (3.10)

for each yNdy\in\mathbb{R}^{Nd}.

4 Examples in function spaces

In this final section, we will apply the ideas we have developed to consider functionals defined on Lebesgue and Sobolev spaces. As we have done above, we will consider both the existence of Nash equilibria and their approximation by continuous time flows. Throughout, we will assume Ωd\Omega\subset\mathbb{R}^{d} is a bounded domain with smooth boundary. We will also change notation by using vv for the variables in which our functionals are defined. This allows us to reserve xx for points in Ω\Omega. Finally, we will use |||\cdot| for any norm on a finite dimensional space.

4.1 A few remarks on Sobolev spaces

Before discussing examples, let us recall a few facts about Sobolev spaces and establish some notation. A good reference for this material is Chapter 5 and 6 of [12]. First we note

H01(Ω)L2(Ω)H1(Ω)H^{1}_{0}(\Omega)\subset L^{2}(\Omega)\subset H^{-1}(\Omega)

where each inclusion is compact. Here L2(Ω)L^{2}(\Omega) is the space of Lebesgue measurable functions on Ω\Omega which are square integrable equipped with the standard inner product. The space H01(Ω)H^{1}_{0}(\Omega) is the closure of all smooth functions having compact support in Ω\Omega in the norm

u(Ω|Du|2𝑑x)1/2.u\mapsto\left(\int_{\Omega}|Du|^{2}dx\right)^{1/2}.

The topological dual of H01(Ω)H^{1}_{0}(\Omega) is denoted H1(Ω)H^{-1}(\Omega). Recall that the Dirichlet Laplacian Δ:H01(Ω)H1(Ω)-\Delta:H^{1}_{0}(\Omega)\rightarrow H^{-1}(\Omega) is defined via

(Δu,v)=ΩDuDv𝑑x(-\Delta u,v)=\int_{\Omega}Du\cdot Dvdx

for u,vH01(Ω)u,v\in H^{1}_{0}(\Omega). Here we are using (,)(\cdot,\cdot) as the natural pairing between H1(Ω)H^{-1}(\Omega) and H01(Ω)H^{1}_{0}(\Omega). Moreover, Δ-\Delta is an isometry; in particular, this map is invertible. This allows us to define the following inner product on H1(Ω)H^{-1}(\Omega)

(f,g)(f,(Δ)1g).(f,g)\mapsto(f,(-\Delta)^{-1}g).

We also note that if fL2(Ω)f\in L^{2}(\Omega), the inner product between ff and gg simplifies to

(f,(Δ)1g)=Ωf(Δ)1g𝑑x.(f,(-\Delta)^{-1}g)=\int_{\Omega}f(-\Delta)^{-1}gdx. (4.1)

In addition, we will consider the space H2(Ω)H01(Ω)H^{2}(\Omega)\cap H^{1}_{0}(\Omega) consisting of H01(Ω)H^{1}_{0}(\Omega) functions for which ΔuL2(Ω)-\Delta u\in L^{2}(\Omega). This is a Hilbert space endowed with the inner product

(u,v)ΩΔuΔv𝑑x.(u,v)\mapsto\int_{\Omega}\Delta u\;\Delta vdx.

4.2 H=L2(Ω)H=L^{2}(\Omega)

For vH01(Ω)Nv\in H^{1}_{0}(\Omega)^{N}, define

fj(v)=ΩHj(vj)+Fj(v)hjvjdxf_{j}(v)=\int_{\Omega}H_{j}(\nabla v_{j})+F_{j}(v)-h_{j}v_{j}\;dx (4.2)

for j=1,,Nj=1,\dots,N. Here each hjL2(Ω)h_{j}\in L^{2}(\Omega), and Hj:dH_{j}:\mathbb{R}^{d}\rightarrow\mathbb{R} is assumed to be smooth and satisfy

θ|pq|2(Hj(p)Hj(q))(pq)1θ|pq|2\theta|p-q|^{2}\leq(\nabla H_{j}(p)-\nabla H_{j}(q))\cdot(p-q)\leq\frac{1}{\theta}|p-q|^{2} (4.3)

for all p,qdp,q\in\mathbb{R}^{d} and some θ(0,1]\theta\in(0,1]. We’ll also suppose F1,,FN:NF_{1},\dots,F_{N}:\mathbb{R}^{N}\rightarrow\mathbb{R} are smooth,

j=1N(zjFj(z)zjFj(w))(zjwj)0\sum^{N}_{j=1}(\partial_{z_{j}}F_{j}(z)-\partial_{z_{j}}F_{j}(w))(z_{j}-w_{j})\geq 0 (4.4)

for z,wNz,w\in\mathbb{R}^{N}, and there is CC such that

|Fj(z)|C(1+|z|2)and|zjFj(z)|C(1+|z|)|F_{j}(z)|\leq C(1+|z|^{2})\;\text{and}\quad|\partial_{z_{j}}F_{j}(z)|\leq C(1+|z|) (4.5)

for zNz\in\mathbb{R}^{N} and j=1,,Nj=1,\dots,N.

Using these assumptions, it is straightforward to verify that f1,,fNf_{1},\dots,f_{N} fulfills (1.9) with X1==XN=V=H01(Ω).X_{1}=\cdots=X_{N}=V=H^{1}_{0}(\Omega). The following lemma also follows by routine computations.

Lemma 4.1.

For each v,wH01(Ω)Nv,w\in H^{1}_{0}(\Omega)^{N} and j=1,,Nj=1,\dots,N,

vjfj(v)={vjfj(v)}\partial_{v_{j}}f_{j}(v)=\{\nabla_{v_{j}}f_{j}(v)\}

with

(vjfj(v),wj)=ΩHj(vj)wj+zjFj(v)wjhjwjdx(\nabla_{v_{j}}f_{j}(v),w_{j})=\int_{\Omega}\nabla H_{j}(\nabla v_{j})\cdot\nabla w_{j}+\partial_{z_{j}}F_{j}(v)w_{j}-h_{j}w_{j}dx

and

j=1N(vjfj(v)vjfj(w),vjwj)θΩ|vw|2𝑑x.\sum^{N}_{j=1}(\nabla_{v_{j}}f_{j}(v)-\nabla_{v_{j}}f_{j}(w),v_{j}-w_{j})\geq\theta\int_{\Omega}|\nabla v-\nabla w|^{2}dx.

Moreover, vv is a Nash equilibrium of f1,,fNf_{1},\dots,f_{N} if and only if it is a weak solution of

{div(Hj(vj))+zjFj(v)=hjinΩvj=0onΩ\begin{cases}-\textup{div}(\nabla H_{j}(\nabla v_{j}))+\partial_{z_{j}}F_{j}(v)=h_{j}&\textup{in}\;\Omega\\ \hskip 135.14455ptv_{j}=0&\textup{on}\;\partial\Omega\end{cases} (4.6)

(j=1,,N)(j=1,\dots,N).

Remark 4.2.

In this statement, (,)(\cdot,\cdot) denotes the natural pairing between H1(Ω)H^{-1}(\Omega) and H01(Ω)H^{1}_{0}(\Omega).

We now can state a result involving the approximation of Nash equilibrium for f1,,fNf_{1},\dots,f_{N}.

Proposition 4.3.

Let u0(H2(Ω)H01(Ω))Nu^{0}\in(H^{2}(\Omega)\cap H^{1}_{0}(\Omega))^{N}. There is a unique Lipschitz continuous

u:[0,)L2(Ω)N;tu(,t)u:[0,\infty)\rightarrow L^{2}(\Omega)^{N};t\mapsto u(\cdot,t)

such that u(,t)(H2(Ω)H01(Ω))Nu(\cdot,t)\in(H^{2}(\Omega)\cap H^{1}_{0}(\Omega))^{N} for each t0t\geq 0 and

{tuj=div(Hj(uj))zjFj(u)+hjinΩ×(0,)uj=0onΩ×(0,)uj=uj0onΩ×{0}\begin{cases}\partial_{t}u_{j}=\textup{div}(\nabla H_{j}(\nabla u_{j}))-\partial_{z_{j}}F_{j}(u)+h_{j}\;\;&\textup{in}\;\Omega\times(0,\infty)\\ \;\;\;u_{j}=0\;\;&\textup{on}\;\partial\Omega\times(0,\infty)\\ \;\;\;u_{j}=u^{0}_{j}\;\;&\textup{on}\;\Omega\times\{0\}\end{cases} (4.7)

(j=1,,N)(j=1,\dots,N). Furthermore, there is λ>0\lambda>0 such that

Ω|u(x,t)v(x)|2𝑑xe2λtΩ|u0(x)v(x)|2𝑑x\int_{\Omega}|u(x,t)-v(x)|^{2}dx\leq e^{-2\lambda t}\int_{\Omega}|u^{0}(x)-v(x)|^{2}dx (4.8)

for t0t\geq 0. Here vv is the unique Nash equilibrium of f1,,fNf_{1},\dots,f_{N}.

Remark 4.4.

For almost every t0t\geq 0, the PDE in (4.7) holds almost everywhere in Ω\Omega; the boundary condition holds in the trace sense; and u(,0)=u0u(\cdot,0)=u^{0} as L2(Ω)NL^{2}(\Omega)^{N} functions.

Proof.

Suppose that ζL2(Ω)N\zeta\in L^{2}(\Omega)^{N} and vH01(Ω)Nv\in H^{1}_{0}(\Omega)^{N} is a weak solution of

{div(Hj(vj))+zjFj(v)=hj+ζjinΩvj=0onΩ.\begin{cases}-\textup{div}(\nabla H_{j}(\nabla v_{j}))+\partial_{z_{j}}F_{j}(v)=h_{j}+\zeta_{j}&\textup{in}\;\Omega\\ \hskip 135.14455ptv_{j}=0&\textup{on}\;\partial\Omega.\end{cases} (4.9)

As Ω\partial\Omega is assumed to be smooth and in view of the hypothesis (4.3), elliptic regularity (Theorem 1, section 8.3 of [12]) implies v(H2(Ω)H01(Ω))Nv\in(H^{2}(\Omega)\cap H^{1}_{0}(\Omega))^{N}. Therefore,

{vH01(Ω)N:v1f1(v),,vNfN(v)L2(Ω)}=(H2(Ω)H01(Ω))N.\displaystyle\left\{v\in H^{1}_{0}(\Omega)^{N}:\nabla_{v_{1}}f_{1}(v),\dots,\nabla_{v_{N}}f_{N}(v)\in L^{2}(\Omega)\right\}=(H^{2}(\Omega)\cap H^{1}_{0}(\Omega))^{N}. (4.10)

Theorem 1.1 and Corollary 2.6 now allow us to conclude. ∎

4.3 Another example with H=L2(Ω)H=L^{2}(\Omega)

Define

fj(v)=ΩLj(v1,,vN)hjvjdxf_{j}(v)=\int_{\Omega}L_{j}(\nabla v_{1},\dots,\nabla v_{N})-h_{j}v_{j}\;dx

for vH01(Ω)Nv\in H^{1}_{0}(\Omega)^{N}. Here h1,,hNL2(Ω)h_{1},\dots,h_{N}\in L^{2}(\Omega) and each Lj:NdL_{j}:\mathbb{R}^{Nd}\rightarrow\mathbb{R} is smooth with

θ|pq|2j=1N(pjLj(p)pjLj(q))(pjqj)1θ|pq|2\displaystyle\theta|p-q|^{2}\leq\sum^{N}_{j=1}(\nabla_{p_{j}}L_{j}(p)-\nabla_{p_{j}}L_{j}(q))\cdot(p_{j}-q_{j})\leq\frac{1}{\theta}|p-q|^{2} (4.11)

for all p,qNdp,q\in\mathbb{R}^{Nd} and some θ(0,1]\theta\in(0,1]. We’ll also assume there is C>0C>0 such that

|pjLj(p)|C(1+|p|)|\nabla_{p_{j}}L_{j}(p)|\leq C(1+|p|) (4.12)

for each pNdp\in\mathbb{R}^{Nd} and j=1,,Nj=1,\dots,N.

Direct computation gives the following lemma.

Lemma 4.5.

For each v,wH01(Ω)Nv,w\in H^{1}_{0}(\Omega)^{N} and j=1,,Nj=1,\dots,N,

vjfj(v)={vjfj(v)}\partial_{v_{j}}f_{j}(v)=\{\nabla_{v_{j}}f_{j}(v)\}

with

(vjfj(v),wj)=ΩpjLj(v1,,vN)wjhjwjdx(\nabla_{v_{j}}f_{j}(v),w_{j})=\int_{\Omega}\nabla_{p_{j}}L_{j}(\nabla v_{1},\dots,\nabla v_{N})\cdot\nabla w_{j}-h_{j}w_{j}\;dx (4.13)

and

j=1N(vjfj(v)vjfj(w),vjwj)θΩ|vw|2𝑑x.\sum^{N}_{j=1}(\nabla_{v_{j}}f_{j}(v)-\nabla_{v_{j}}f_{j}(w),v_{j}-w_{j})\geq\theta\int_{\Omega}|\nabla v-\nabla w|^{2}dx.

Moreover,

vjfj(v+tw)vjfj(v)\nabla_{v_{j}}f_{j}(v+tw)\rightharpoonup\nabla_{v_{j}}f_{j}(v)

as t0+t\rightarrow 0^{+}, and vv is a Nash equilibrium of f1,,fNf_{1},\dots,f_{N} if and only if it is a weak solution of

{div(pjLj(v1,,vN))=hjinΩvj=0onΩ\begin{cases}-\textup{div}(\nabla_{p_{j}}L_{j}(\nabla v_{1},\dots,\nabla v_{N}))=h_{j}&\textup{in}\;\Omega\\ \hskip 135.86795ptv_{j}=0&\textup{on}\;\partial\Omega\end{cases} (4.14)

(j=1,,N)(j=1,\dots,N).

With this lemma, we can now establish the subsequent assertion about the Nash equilibrium of f1,,fNf_{1},\dots,f_{N}.

Proposition 4.6.

Let u0(H2(Ω)H01(Ω))Nu^{0}\in(H^{2}(\Omega)\cap H^{1}_{0}(\Omega))^{N}. There is a unique Lipschitz continuous

u:[0,)L2(Ω)N;tu(,t)u:[0,\infty)\rightarrow L^{2}(\Omega)^{N};t\mapsto u(\cdot,t)

such that u(,t)(H2(Ω)H01(Ω))Nu(\cdot,t)\in(H^{2}(\Omega)\cap H^{1}_{0}(\Omega))^{N} for each t0t\geq 0 and

{tuj=div(pjLj(u1,,uN))+hjinΩ×(0,)uj=0onΩ×(0,)uj=uj0onΩ×{0}\begin{cases}\partial_{t}u_{j}=\textup{div}(\nabla_{p_{j}}L_{j}(\nabla u_{1},\dots,\nabla u_{N}))+h_{j}\;\;&\textup{in}\;\Omega\times(0,\infty)\\ \;\;\;u_{j}=0\;\;&\textup{on}\;\partial\Omega\times(0,\infty)\\ \;\;\;u_{j}=u^{0}_{j}\;\;&\textup{on}\;\Omega\times\{0\}\end{cases} (4.15)

(j=1,,N(j=1,\dots,N). Furthermore, there is λ>0\lambda>0 such that

Ω|u(x,t)v(x)|2𝑑xe2λtΩ|u0(x)v(x)|2𝑑x\int_{\Omega}|u(x,t)-v(x)|^{2}dx\leq e^{-2\lambda t}\int_{\Omega}|u^{0}(x)-v(x)|^{2}dx (4.16)

for t0t\geq 0. Here vv is the unique Nash equilibrium of f1,,fNf_{1},\dots,f_{N}.

Proof.

Suppose that ξL2(Ω)N\xi\in L^{2}(\Omega)^{N} and vH01(Ω)Nv\in H^{1}_{0}(\Omega)^{N} is a weak solution of

{div(pjLj(v1,,vN))=hj+ξjinΩvj=0onΩ.\begin{cases}-\textup{div}(\nabla_{p_{j}}L_{j}(\nabla v_{1},\dots,\nabla v_{N}))=h_{j}+\xi_{j}&\textup{in}\;\Omega\\ \hskip 135.86795ptv_{j}=0&\textup{on}\;\partial\Omega.\end{cases} (4.17)

By the uniform estimate (4.11), elliptic regularity implies v(H2(Ω)H01(Ω))Nv\in(H^{2}(\Omega)\cap H^{1}_{0}(\Omega))^{N} (see the remark at the end of section 9.1 in [12]). As a result,

{vH01(Ω)N:v1f1(v),,vNfN(v)L2(Ω)}=(H2(Ω)H01(Ω))N.\displaystyle\left\{v\in H^{1}_{0}(\Omega)^{N}:\nabla_{v_{1}}f_{1}(v),\dots,\nabla_{v_{N}}f_{N}(v)\in L^{2}(\Omega)\right\}=(H^{2}(\Omega)\cap H^{1}_{0}(\Omega))^{N}. (4.18)

We then conclude by Theorem 2.7 and Remark 2.8. ∎

4.4 H=H1(Ω)H=H^{-1}(\Omega)

For a given vL2(Ω)Nv\in L^{2}(\Omega)^{N}, set

fj(v)=ΩFj(v)𝑑xhj,vjf_{j}(v)=\int_{\Omega}F_{j}(v)dx-\langle h_{j},v_{j}\rangle

for j=1,,Nj=1,\dots,N. Here h1,,hNH1(Ω)h_{1},\dots,h_{N}\in H^{-1}(\Omega), and ,\langle\cdot,\cdot\rangle denotes the inner product in H1(Ω)H^{-1}(\Omega). We will assume F1,,FN:NF_{1},\dots,F_{N}:\mathbb{R}^{N}\rightarrow\mathbb{R} are smooth and satisfy (4.5) and

{|zkzjFj(z)|Cθ|zw|2i=1N(ziFi(z)ziFi(w))(ziwi)\begin{cases}|\partial_{z_{k}}\partial_{z_{j}}F_{j}(z)|\leq C\\ \\ \displaystyle\theta|z-w|^{2}\leq\sum^{N}_{i=1}(\partial_{z_{i}}F_{i}(z)-\partial_{z_{i}}F_{i}(w))(z_{i}-w_{i})\end{cases} (4.19)

for z,wNz,w\in\mathbb{R}^{N}, j,k=1,,Nj,k=1,\dots,N, and some C,θ>0C,\theta>0. For convenience, we will also assume

zjFj(0)=0\partial_{z_{j}}F_{j}(0)=0 (4.20)

for j=1,,Nj=1,\dots,N.

The following observations can be verified by routine computations.

Lemma 4.7.

For each v,wL2(Ω)Nv,w\in L^{2}(\Omega)^{N} and j=1,,Nj=1,\dots,N,

vjfj(v)={vjfj(v)}\partial_{v_{j}}f_{j}(v)=\{\nabla_{v_{j}}f_{j}(v)\}

with

vjfj(v)=zjFj(v)(Δ)1hj.\nabla_{v_{j}}f_{j}(v)=\partial_{z_{j}}F_{j}(v)-(-\Delta)^{-1}h_{j}. (4.21)

Moreover, vH01(Ω)Nv\in H^{1}_{0}(\Omega)^{N} is a Nash equilibrium of f1,,fNf_{1},\dots,f_{N} if and only if it is a weak solution of

{Δ(zjFj(v))=hjinΩvj=0onΩ\begin{cases}-\Delta(\partial_{z_{j}}F_{j}(v))=h_{j}&\textup{in}\;\Omega\\ \hskip 60.70653ptv_{j}=0&\textup{on}\;\partial\Omega\end{cases} (4.22)

(j=1,,N)(j=1,\dots,N).

By Proposition A.1 in the appendix, there are smooth functions G1,,GNG_{1},\dots,G_{N} such that

{zjFj(z)=yjif and only ifyjGj(y)=zj|Gj(y)|B(1+|y|2)and|yjGj(y)|B(1+|y|)0i=1N(yiGi(y)yiGi(z))(yizi)|ykyjGj(y)|1θ\begin{cases}\partial_{z_{j}}F_{j}(z)=y_{j}\;\text{if and only if}\;\partial_{y_{j}}G_{j}(y)=z_{j}\\ \\ |G_{j}(y)|\leq B(1+|y|^{2})\;\text{and}\quad|\partial_{y_{j}}G_{j}(y)|\leq B(1+|y|)\\ \\ \displaystyle 0\leq\sum^{N}_{i=1}\left(\partial_{y_{i}}G_{i}(y)-\partial_{y_{i}}G_{i}(z)\right)(y_{i}-z_{i})\\ \\ \displaystyle|\partial_{y_{k}}\partial_{y_{j}}G_{j}(y)|\leq\frac{1}{\theta}\end{cases} (4.23)

for all y,zNy,z\in\mathbb{R}^{N}, j,k=1,,Nj,k=1,\dots,N and some constant BB. Notice that in view of (4.20)

yjGj(0)=0\partial_{y_{j}}G_{j}(0)=0 (4.24)

for j=1,,Nj=1,\dots,N. Thus, it is possible to solve (4.22) by choosing

wj=(Δ)1hjH01(Ω)w_{j}=(-\Delta)^{-1}h_{j}\in H^{1}_{0}(\Omega)

for j=1,,Nj=1,\dots,N and then setting

vj=yjGj(w)v_{j}=\partial_{y_{j}}G_{j}(w)

for j=1,,Nj=1,\dots,N. In particular, we have just shown that f1,,fNf_{1},\dots,f_{N} has a Nash equilibrium, and it is not hard to see it is unique.

Let us now see how to approximate this Nash equilibrium.

Proposition 4.8.

Let u0H01(Ω)Nu^{0}\in H^{1}_{0}(\Omega)^{N}. There is a unique Lipschitz continuous

u:[0,)H1(Ω)N;tu(,t)u:[0,\infty)\rightarrow H^{-1}(\Omega)^{N};t\mapsto u(\cdot,t)

such that u(,t)H01(Ω)Nu(\cdot,t)\in H^{1}_{0}(\Omega)^{N} for each t0t\geq 0 and

{tuj=Δ(zjFj(u))hjinΩ×(0,)uj=0onΩ×(0,)uj=uj0onΩ×{0}\begin{cases}\partial_{t}u_{j}=\Delta(\partial_{z_{j}}F_{j}(u))-h_{j}\;\;&\textup{in}\;\Omega\times(0,\infty)\\ \;\;\;u_{j}=0\;\;&\textup{on}\;\partial\Omega\times(0,\infty)\\ \;\;\;u_{j}=u^{0}_{j}\;\;&\textup{on}\;\Omega\times\{0\}\end{cases} (4.25)

(j=1,,N(j=1,\dots,N). Furthermore, there is λ>0\lambda>0 such that

u(,t)veλtu0v\|u(\cdot,t)-v\|\leq e^{-\lambda t}\|u^{0}-v\| (4.26)

for t0t\geq 0. Here \|\cdot\| denotes the norm on H1(Ω)NH^{-1}(\Omega)^{N} and vv is the Nash equilibrium of f1,,fNf_{1},\dots,f_{N}.

Proof.

Define A:H1(Ω)N2H1(Ω)NA:H^{-1}(\Omega)^{N}\rightarrow 2^{H^{-1}(\Omega)^{N}} as

Av={{(Δ(z1F1(v))+h1,,Δ(zNFN(v))+hN)},vH01(Ω)N,otherwiseAv=\begin{cases}\{(-\Delta(\partial_{z_{1}}F_{1}(v))+h_{1},\dots,-\Delta(\partial_{z_{N}}F_{N}(v))+h_{N})\},&v\in H^{1}_{0}(\Omega)^{N}\\ \emptyset,&\text{otherwise}\end{cases}

for vH1(Ω)Nv\in H^{-1}(\Omega)^{N}. We note that D(A)=H01(Ω)ND(A)=H^{1}_{0}(\Omega)^{N}, and 0Av0\in Av if and only if vH01(Ω)Nv\in H^{1}_{0}(\Omega)^{N} is a Nash equilibrium for f1,,fNf_{1},\dots,f_{N}. By the second inequality listed in (4.19),

AvAw,vw\displaystyle\langle Av-Aw,v-w\rangle =Ωj=1N(zjFj(v)zjFj(w))(vjwj)dx\displaystyle=\int_{\Omega}\sum^{N}_{j=1}(\partial_{z_{j}}F_{j}(v)-\partial_{z_{j}}F_{j}(w))(v_{j}-w_{j})dx (4.27)
θΩ|vw|2𝑑x\displaystyle\geq\theta\int_{\Omega}|v-w|^{2}dx (4.28)
λvw2\displaystyle\geq\lambda\|v-w\|^{2} (4.29)

for v,wH01(Ω)Nv,w\in H^{1}_{0}(\Omega)^{N} and some λ>0\lambda>0.

As a result, it suffices to check that AA is maximal. This amounts to finding a weak solution vH01(Ω)Nv\in H^{1}_{0}(\Omega)^{N} of

{vjΔ(zjFj(v))=hj+ζjinΩvj=0onΩ\begin{cases}v_{j}-\Delta(\partial_{z_{j}}F_{j}(v))=h_{j}+\zeta_{j}&\textup{in}\;\Omega\\ \hskip 75.88371ptv_{j}=0&\textup{on}\;\partial\Omega\end{cases} (4.30)

for a given ζH1(Ω)N\zeta\in H^{-1}(\Omega)^{N}. To this end, we consider the auxiliary problem of finding wH01(Ω)Nw\in H^{1}_{0}(\Omega)^{N} such that

{yjGj(w)Δwj=hj+ζjinΩwj=0onΩ.\begin{cases}\partial_{y_{j}}G_{j}(w)-\Delta w_{j}=h_{j}+\zeta_{j}&\textup{in}\;\Omega\\ \hskip 70.82428ptw_{j}=0&\textup{on}\;\partial\Omega.\end{cases} (4.31)

If we can find ww, then vH01(Ω)Nv\in H^{1}_{0}(\Omega)^{N} defined by

vj=yjGj(w)v_{j}=\partial_{y_{j}}G_{j}(w)

(j=1,,N)(j=1,\dots,N) is a solution of (4.30).

Notice that (4.31) has a solution if and only if

gj(w)=ΩGj(w)+12|wj|2dx(hj+ζj,wj)g_{j}(w)=\int_{\Omega}G_{j}(w)+\frac{1}{2}|\nabla w_{j}|^{2}dx-(h_{j}+\zeta_{j},w_{j}) (4.32)

(wH01(Ω)Nw\in H^{1}_{0}(\Omega)^{N}, j=1,,Nj=1,\dots,N) has a Nash equilibrium. Here (,)(\cdot,\cdot) is the natural pairing between H1(Ω)H^{-1}(\Omega) and H01(Ω)H^{1}_{0}(\Omega). Using (4.23), it is straightforward to check that g1,,gNg_{1},\dots,g_{N} satisfies the properties listed in (1.9) with X1==XN=V=H01(Ω)X_{1}=\dots=X_{N}=V=H^{1}_{0}(\Omega). As a result, Corollary 2.6 implies that g1,,gNg_{1},\dots,g_{N} has a Nash equilibrium. ∎

4.5 H=H01(Ω)H=H^{1}_{0}(\Omega)

Set

fj(v)=ΩFj(Δv)hjvjdxf_{j}(v)=\int_{\Omega}F_{j}(\Delta v)-\nabla h_{j}\cdot\nabla v_{j}\;dx

for v(H2(Ω)H01(Ω))Nv\in(H^{2}(\Omega)\cap H^{1}_{0}(\Omega))^{N} and j=1,,Nj=1,\dots,N. Here h1,,hNH01(Ω)h_{1},\dots,h_{N}\in H^{1}_{0}(\Omega) and

Δv=(Δv1,,ΔvN).\Delta v=(\Delta v_{1},\dots,\Delta v_{N}).

We will suppose F1,,FN:NF_{1},\dots,F_{N}:\mathbb{R}^{N}\rightarrow\mathbb{R} are smooth and satisfy (4.19) and (4.20). A few key observations regarding f1,fNf_{1},\dots f_{N} are as follows.

Lemma 4.9.

For each v,w(H2(Ω)H01(Ω))Nv,w\in(H^{2}(\Omega)\cap H^{1}_{0}(\Omega))^{N} and j=1,,Nj=1,\dots,N,

vjfj(v)={vjfj(v)}\partial_{v_{j}}f_{j}(v)=\{\nabla_{v_{j}}f_{j}(v)\}

with

(vjfj(v),wj)=ΩzjFj(Δv)Δwjhjwjdx.(\nabla_{v_{j}}f_{j}(v),w_{j})=\int_{\Omega}\partial_{z_{j}}F_{j}(\Delta v)\Delta w_{j}-\nabla h_{j}\cdot\nabla w_{j}dx.

Moreover, vv is a Nash equilibrium of f1,,fNf_{1},\dots,f_{N} if and only if it is a solution of

{zjFj(Δv)=hjinΩvj=0onΩ\begin{cases}-\partial_{z_{j}}F_{j}(\Delta v)=h_{j}&\textup{in}\;\Omega\\ \hskip 51.31218ptv_{j}=0&\textup{on}\;\partial\Omega\end{cases} (4.33)

(j=1,,N)(j=1,\dots,N).

Remark 4.10.

Above, (,)(\cdot,\cdot) denotes the natural pairing between (H2(Ω)H01(Ω))(H^{2}(\Omega)\cap H^{1}_{0}(\Omega))^{*} and H2(Ω)H01(Ω)H^{2}(\Omega)\cap H^{1}_{0}(\Omega).

Using the functions G1,,GNG_{1},\dots,G_{N} from the previous subsection, we can solve

{Δvj=yjGj(h)inΩvj=0onΩ,\begin{cases}-\Delta v_{j}=-\partial_{y_{j}}G_{j}(-h)&\textup{in}\;\Omega\\ \hskip 18.7898ptv_{j}=0&\textup{on}\;\partial\Omega,\end{cases} (4.34)

for j=1,,Nj=1,\dots,N to obtain a solution of (4.33). It follows that f1,,fNf_{1},\dots,f_{N} has a unique Nash equilibrium vv, which belongs to the space

𝒟={v(H2(Ω)H01(Ω))N:ΔvH01(Ω)N}.{\cal D}=\left\{v\in(H^{2}(\Omega)\cap H^{1}_{0}(\Omega))^{N}:\Delta v\in H^{1}_{0}(\Omega)^{N}\right\}.

As with our previous examples, we can approximate this Nash equilibrium with a continuous time flow.

Proposition 4.11.

Let u0𝒟u^{0}\in{\cal D}. There is a unique Lipschitz continuous

u:[0,)H01(Ω)N;tu(,t)u:[0,\infty)\rightarrow H^{1}_{0}(\Omega)^{N};t\mapsto u(\cdot,t)

such that u(,t)𝒟u(\cdot,t)\in{\cal D} for each t0t\geq 0 and

{tuj=zjFj(Δu)+hjinΩ×(0,)uj=0onΩ×(0,)uj=uj0onΩ×{0}\begin{cases}\partial_{t}u_{j}=\partial_{z_{j}}F_{j}(\Delta u)+h_{j}\;\;&\textup{in}\;\Omega\times(0,\infty)\\ \;\;\;u_{j}=0\;\;&\textup{on}\;\partial\Omega\times(0,\infty)\\ \;\;\;u_{j}=u^{0}_{j}\;\;&\textup{on}\;\Omega\times\{0\}\end{cases} (4.35)

(j=1,,N(j=1,\dots,N). Furthermore, there is λ>0\lambda>0 such that

Ω|u(x,t)v(x)|2𝑑xe2λtΩ|u0(x)v(x)|2𝑑x\int_{\Omega}|\nabla u(x,t)-\nabla v(x)|^{2}dx\leq e^{-2\lambda t}\int_{\Omega}|\nabla u^{0}(x)-\nabla v(x)|^{2}dx (4.36)

for t0t\geq 0. Here vv is the unique Nash equilibrium of f1,,fNf_{1},\dots,f_{N}.

Proof.

Define A:H01(Ω)N2H01(Ω)NA:H^{1}_{0}(\Omega)^{N}\rightarrow 2^{H^{1}_{0}(\Omega)^{N}} by

Av={{(z1F1(Δv)h1,,zNFN(Δv)hN)},v𝒟,otherwiseAv=\begin{cases}\{\left(-\partial_{z_{1}}F_{1}(\Delta v)-h_{1},\dots,-\partial_{z_{N}}F_{N}(\Delta v)-h_{N}\right)\},&v\in{\cal D}\\ \emptyset,&\text{otherwise}\end{cases} (4.37)

for vH01(Ω)Nv\in H^{1}_{0}(\Omega)^{N}. Note that D(A)=𝒟D(A)={\cal D}, and 0Av0\in Av if and only if v𝒟v\in{\cal D} is a Nash equilibrium for f1,,fNf_{1},\dots,f_{N}. In view of (4.19),

AvAw,vw\displaystyle\langle Av-Aw,v-w\rangle =Ωj=1N(zjFj(Δv)zjFj(Δw))(ΔvjΔwj)dx\displaystyle=\int_{\Omega}\sum^{N}_{j=1}(\partial_{z_{j}}F_{j}(\Delta v)-\partial_{z_{j}}F_{j}(\Delta w))(\Delta v_{j}-\Delta w_{j})dx (4.38)
θΩ|ΔvΔw|2𝑑x\displaystyle\geq\theta\int_{\Omega}|\Delta v-\Delta w|^{2}dx (4.39)
λΩ|vw|2𝑑x.\displaystyle\geq\lambda\int_{\Omega}|\nabla v-\nabla w|^{2}dx. (4.40)

for v,w𝒟v,w\in{\cal D} and some λ>0\lambda>0.

Therefore, it suffices to show

{vjzjFj(Δv)=hj+ζjinΩvj=0onΩ\begin{cases}v_{j}-\partial_{z_{j}}F_{j}(\Delta v)=h_{j}+\zeta_{j}&\textup{in}\;\Omega\\ \hskip 67.21056ptv_{j}=0&\textup{on}\;\partial\Omega\end{cases} (4.41)

has a solution v(H2(Ω)H01(Ω))Nv\in(H^{2}(\Omega)\cap H^{1}_{0}(\Omega))^{N} for a given ζH01(Ω)N\zeta\in H^{1}_{0}(\Omega)^{N}. We note that any such solution vv will automatically belong to 𝒟{\cal D}. Moreover, this is equivalent to finding a weak solution vH01(Ω)Nv\in H^{1}_{0}(\Omega)^{N} of

{yjGj(vhζ)Δvj=0inΩvj=0onΩ.\begin{cases}\partial_{y_{j}}G_{j}(v-h-\zeta)-\Delta v_{j}=0&\textup{in}\;\Omega\\ \hskip 112.0187ptv_{j}=0&\textup{on}\;\partial\Omega.\end{cases} (4.42)

We emphasize that elliptic regularity implies that any such weak solution actually belongs to (H2(Ω)H01(Ω))N(H^{2}(\Omega)\cap H^{1}_{0}(\Omega))^{N}.

Any vH01(Ω)Nv\in H^{1}_{0}(\Omega)^{N} which solves (4.42) is a Nash equilibrium of g1,,gNg_{1},\dots,g_{N} defined by

gj(v)=ΩGj(vhζ)+12|vj|2dx.g_{j}(v)=\int_{\Omega}G_{j}(v-h-\zeta)+\frac{1}{2}|\nabla v_{j}|^{2}dx. (4.43)

As we did when considering the functions defined in (4.32), we can apply Corollary 2.6 and conclude that g1,,gNg_{1},\dots,g_{N} has a unique Nash equilibrium. ∎

Appendix A Dual functions

In this appendix, we will verify a technical assertion that was used to establish the existence of a Nash equilibrium in a few of the examples we considered above.

Proposition A.1.

Suppose F1,,FN:NF_{1},\dots,F_{N}:\mathbb{R}^{N}\rightarrow\mathbb{R} are smooth and satisfy

|Fj(z)|C(1+|z|2)|F_{j}(z)|\leq C(1+|z|^{2}) (A.1)

for zNz\in\mathbb{R}^{N} and j=1,,Nj=1,\dots,N and

θ|zw|2j=1N(zjFj(z)zjFj(w))(zjwj)\theta|z-w|^{2}\leq\sum^{N}_{j=1}\left(\partial_{z_{j}}F_{j}(z)-\partial_{z_{j}}F_{j}(w)\right)(z_{j}-w_{j}) (A.2)

for z,wNz,w\in\mathbb{R}^{N} and some θ>0\theta>0. There are smooth functions G1,,GN:NG_{1},\dots,G_{N}:\mathbb{R}^{N}\rightarrow\mathbb{R} with the following properties.

  1. (i)

    zjFj(z)=yj\partial_{z_{j}}F_{j}(z)=y_{j} if and only if yjGj(y)=zj\partial_{y_{j}}G_{j}(y)=z_{j}.

  2. (ii)

    For all j,k=1,,Nj,k=1,\dots,N and yNy\in\mathbb{R}^{N},

    |ykyjGj(y)|1θ.|\partial_{y_{k}}\partial_{y_{j}}G_{j}(y)|\leq\frac{1}{\theta}. (A.3)
  3. (iii)

    For each y,wNy,w\in\mathbb{R}^{N},

    0j=1N(yjGj(y)yjGj(w))(yjwj).0\leq\sum^{N}_{j=1}\left(\partial_{y_{j}}G_{j}(y)-\partial_{y_{j}}G_{j}(w)\right)(y_{j}-w_{j}). (A.4)
  4. (iv)

    There is a constant BB such that

    |Gj(y)|B(1+|y|2)and|yjGj(y)|B(1+|y|)|G_{j}(y)|\leq B(1+|y|^{2})\;\text{and}\quad|\partial_{y_{j}}G_{j}(y)|\leq B(1+|y|) (A.5)

    for yNy\in\mathbb{R}^{N} and j=1,,Nj=1,\dots,N.

Proof.

First we note that the mapping of N\mathbb{R}^{N}

z(z1F1(z),,zNFN(z))z\mapsto(\partial_{z_{1}}F_{1}(z),\dots,\partial_{z_{N}}F_{N}(z)) (A.6)

is invertible. This follows from the theorem due to Browder [8] and Minty [23] as (A.2) implies that this map is both monotone and coercive; invertibility is a consequence of Hadamard’s global invertibility theorem (Theorem A of [17]), as well. The condition (A.2) also gives

θ|v|2j,k=1NzkzjFj(z)vkvj\theta|v|^{2}\leq\sum^{N}_{j,k=1}\partial_{z_{k}}\partial_{z_{j}}F_{j}(z)v_{k}v_{j} (A.7)

for each z,vNz,v\in\mathbb{R}^{N}.

Now fix yNy\in\mathbb{R}^{N} and let zNz\in\mathbb{R}^{N} be the unique Nash equilibrium of

NxFj(x)yjxj\mathbb{R}^{N}\ni x\mapsto F_{j}(x)-y_{j}x_{j} (A.8)

(j=1,,N)(j=1,\dots,N). Such a Nash equilibrium exists for these NN functions by Corollary 2.6. As a result,

zjFj(z)=yj,j=1,,N.\partial_{z_{j}}F_{j}(z)=y_{j},\quad j=1,\dots,N. (A.9)

Since (A.6) is invertible and smooth, (A.9) determines zz as a smooth function of yy. Let us now define

Gj(y):=zjyjFj(z)G_{j}(y):=z_{j}y_{j}-F_{j}(z) (A.10)

for j=1,,Nj=1,\dots,N. Differentiating with respect to yjy_{j} and using (A.9), we find

yjGj(y)=zj.\partial_{y_{j}}G_{j}(y)=z_{j}.

This proves (i)(i) and that

y(y1G1(y),,yNGN(y))y\mapsto(\partial_{y_{1}}G_{1}(y),\dots,\partial_{y_{N}}G_{N}(y)) (A.11)

is the inverse mapping of (A.6).

It follows from part (i)(i) and the inverse function theorem that whenever yjGj(y)=zj\partial_{y_{j}}G_{j}(y)=z_{j} and zjFj(z)=yj\partial_{z_{j}}F_{j}(z)=y_{j} for j=1,,Nj=1,\dots,N, the matrices

(zkzjFj(z))j,k=1N and (ykyjGj(y))j,k=1N(\partial_{z_{k}}\partial_{z_{j}}F_{j}(z))^{N}_{j,k=1}\text{ and }(\partial_{y_{k}}\partial_{y_{j}}G_{j}(y))^{N}_{j,k=1}

are inverses. Consequently, for a given wNw\in\mathbb{R}^{N}, we can choose vNv\in\mathbb{R}^{N} defined as

vj=k=1NykyjGj(y)wkv_{j}=\sum^{N}_{k=1}\partial_{y_{k}}\partial_{y_{j}}G_{j}(y)w_{k}

in (A.7) to get

0\displaystyle 0 θj=1N(k=1NykyjGj(y)wk)2\displaystyle\leq\theta\sum^{N}_{j=1}\left(\sum^{N}_{k=1}\partial_{y_{k}}\partial_{y_{j}}G_{j}(y)w_{k}\right)^{2}
j,k=1NykyjGj(y)wkwj\displaystyle\leq\sum^{N}_{j,k=1}\partial_{y_{k}}\partial_{y_{j}}G_{j}(y)w_{k}w_{j} (A.12)
=j=1N(k=1NykyjGj(y)wk)wj\displaystyle=\sum^{N}_{j=1}\left(\sum^{N}_{k=1}\partial_{y_{k}}\partial_{y_{j}}G_{j}(y)w_{k}\right)w_{j} (A.13)
j=1N(k=1NykyjGj(y)wk)2|w|.\displaystyle\leq\sqrt{\sum^{N}_{j=1}\left(\sum^{N}_{k=1}\partial_{y_{k}}\partial_{y_{j}}G_{j}(y)w_{k}\right)^{2}}\cdot|w|. (A.14)

As a result,

j=1N(k=1NykyjGj(y)wk)21θ2|w|2.\sum^{N}_{j=1}\left(\sum^{N}_{k=1}\partial_{y_{k}}\partial_{y_{j}}G_{j}(y)w_{k}\right)^{2}\leq\frac{1}{\theta^{2}}|w|^{2}. (A.15)

Upon selecting w=ekw=e_{k}, the kkth standard unit vector in N\mathbb{R}^{N}, we find

(ykyjGj(y))2=(=1NyyjGj(y)w)2i=1N(=1NyyiGi(y)w)21θ2(\partial_{y_{k}}\partial_{y_{j}}G_{j}(y))^{2}=\left(\sum^{N}_{\ell=1}\partial_{y_{\ell}}\partial_{y_{j}}G_{j}(y)w_{\ell}\right)^{2}\leq\sum^{N}_{i=1}\left(\sum^{N}_{\ell=1}\partial_{y_{\ell}}\partial_{y_{i}}G_{i}(y)w_{\ell}\right)^{2}\leq\frac{1}{\theta^{2}}

for j,k=1,,Nj,k=1,\dots,N. We conclude assertion (ii)(ii).

Part (iii)(iii) follows directly from (A). Indeed, for any y,wNy,w\in\mathbb{R}^{N}, we can use the fundamental theorem of calculus to derive

j=1N(yjGj(y)yjGj(w))(yjwj)\displaystyle\sum^{N}_{j=1}\left(\partial_{y_{j}}G_{j}(y)-\partial_{y_{j}}G_{j}(w)\right)(y_{j}-w_{j}) (A.16)
=01(j,k=1NykyjGj(w+t(yw))(ykwk)(yjwj))𝑑t\displaystyle=\int^{1}_{0}\left(\sum^{N}_{j,k=1}\partial_{y_{k}}\partial_{y_{j}}G_{j}(w+t(y-w))(y_{k}-w_{k})(y_{j}-w_{j})\right)dt (A.17)
0.\displaystyle\geq 0. (A.18)

We are left to verify assertion (iv)(iv). First note that by part (ii)(ii),

|yjGj(y)|\displaystyle|\partial_{y_{j}}G_{j}(y)| |yjGj(y)yjGj(0)|+|yjGj(0)|\displaystyle\leq|\partial_{y_{j}}G_{j}(y)-\partial_{y_{j}}G_{j}(0)|+|\partial_{y_{j}}G_{j}(0)| (A.19)
=|01k=1NykyjGj(ty)ykdt|+|yjGj(0)|\displaystyle=\left|\int^{1}_{0}\sum^{N}_{k=1}\partial_{y_{k}}\partial_{y_{j}}G_{j}(ty)y_{k}dt\right|+|\partial_{y_{j}}G_{j}(0)| (A.20)
k=1N(01|ykyjGj(ty)|𝑑t)|yk|+|yjGj(0)|\displaystyle\leq\sum^{N}_{k=1}\left(\int^{1}_{0}\left|\partial_{y_{k}}\partial_{y_{j}}G_{j}(ty)\right|dt\right)|y_{k}|+|\partial_{y_{j}}G_{j}(0)| (A.21)
1θk=1N|yk|+|yjGj(0)|\displaystyle\leq\frac{1}{\theta}\sum^{N}_{k=1}|y_{k}|+|\partial_{y_{j}}G_{j}(0)| (A.22)
D(1+|y|)\displaystyle\leq D(1+|y|) (A.23)

for some D>0D>0 independent of j=1,,Nj=1,\dots,N and yy.

For a given yNy\in\mathbb{R}^{N} there is a unique zz such that zjFj(z)=yj\partial_{z_{j}}F_{j}(z)=y_{j} and yjGj(y)=zj\partial_{y_{j}}G_{j}(y)=z_{j} for j=1,,Nj=1,\dots,N. We just showed above that

|zj|=|yjGj(y)|D(1+|y|).|z_{j}|=|\partial_{y_{j}}G_{j}(y)|\leq D(1+|y|).

In view of hypothesis, (A.1) and our definition of GjG_{j} (A.10), we additionally have

|Gj(y)|\displaystyle|G_{j}(y)| =|zjyjFj(z)|\displaystyle=|z_{j}y_{j}-F_{j}(z)| (A.24)
|zj||yj|+|Fj(z)|\displaystyle\leq|z_{j}||y_{j}|+|F_{j}(z)| (A.25)
D(1+|y|)|y|+C(1+|z|2)\displaystyle\leq D(1+|y|)|y|+C(1+|z|^{2}) (A.26)
D(1+|y|)|y|+C(1+ND2(1+|y|)2)\displaystyle\leq D(1+|y|)|y|+C(1+ND^{2}(1+|y|)^{2}) (A.27)
B(1+|y|2)\displaystyle\leq B(1+|y|^{2}) (A.28)

for an appropriately chosen constant BB. ∎

References

  • [1] Abdullah Alotaibi, Patrick L. Combettes, and Naseer Shahzad. Solving coupled composite monotone inclusions by successive Fejér approximations of their Kuhn-Tucker set. SIAM J. Optim., 24(4):2076–2095, 2014.
  • [2] Jean-Pierre Aubin and Arrigo Cellina. Differential inclusions, volume 264 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1984. Set-valued maps and viability theory.
  • [3] J. B. Baillon and H. Brézis. Une remarque sur le comportement asymptotique des semigroupes non linéaires. Houston J. Math., 2(1):5–7, 1976.
  • [4] Jonathan M. Borwein and Qiji J. Zhu. Techniques of variational analysis, volume 20 of CMS Books in Mathematics/Ouvrages de Mathématiques de la SMC. Springer-Verlag, New York, 2005.
  • [5] H. Brézis. Opérateurs maximaux monotones et semi-groupes de contractions dans les espaces de Hilbert. North-Holland Publishing Co., Amsterdam-London; American Elsevier Publishing Co., Inc., New York, 1973. North-Holland Mathematics Studies, No. 5. Notas de Matemática (50).
  • [6] Luis M. Briceño Arias. A Douglas-Rachford splitting method for solving equilibrium problems. Nonlinear Anal., 75(16):6053–6059, 2012.
  • [7] Luis M. Briceño Arias and Patrick L. Combettes. Monotone operator methods for Nash equilibria in non-potential games. In Computational and analytical mathematics, volume 50 of Springer Proc. Math. Stat., pages 143–159. Springer, New York, 2013.
  • [8] Felix E. Browder. Nonlinear elliptic boundary value problems. Bull. Amer. Math. Soc., 69:862–874, 1963.
  • [9] Constantinos Daskalakis, Paul W. Goldberg, and Christos H. Papadimitriou. The complexity of computing a Nash equilibrium. SIAM J. Comput., 39(1):195–259, 2009.
  • [10] Constantinos Daskalakis, Aranyak Mehta, and Christos Papadimitriou. A note on approximate Nash equilibria. Theoret. Comput. Sci., 410(17):1581–1588, 2009.
  • [11] Jonathan Eckstein and B. F. Svaiter. General projective splitting methods for sums of maximal monotone operators. SIAM J. Control Optim., 48(2):787–811, 2009.
  • [12] Lawrence C. Evans. Partial differential equations, volume 19 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, second edition, 2010.
  • [13] Francisco Facchinei and Jong-Shi Pang. Finite-dimensional variational inequalities and complementarity problems. Vol. I. Springer Series in Operations Research. Springer-Verlag, New York, 2003.
  • [14] Ky Fan. Fixed-point and minimax theorems in locally convex topological linear spaces. Proc. Nat. Acad. Sci. U.S.A., 38:121–126, 1952.
  • [15] Sjur Didrik Flam. Paths to constrained Nash equilibria. Appl. Math. Optim., 27(3):275–289, 1993.
  • [16] I. L. Glicksberg. A further generalization of the Kakutani fixed theorem, with application to Nash equilibrium points. Proc. Amer. Math. Soc., 3:170–174, 1952.
  • [17] W. B. Gordon. On the diffeomorphisms of Euclidean space. Amer. Math. Monthly, 79:755–759, 1972.
  • [18] Patrick R. Johnstone and Jonathan Eckstein. Convergence rates for projective splitting. SIAM J. Optim., 29(3):1931–1957, 2019.
  • [19] Shizuo Kakutani. A generalization of Brouwer’s fixed point theorem. Duke Math. J., 8:457–459, 1941.
  • [20] David Kinderlehrer and Guido Stampacchia. An introduction to variational inequalities and their applications, volume 31 of Classics in Applied Mathematics. Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2000. Reprint of the 1980 original.
  • [21] Richard J. Lipton, Evangelos Markakis, and Aranyak Mehta. Playing large games using simple strategies. In Proceedings of the 4th ACM Conference on Electronic Commerce, EC’03, pages 36–41, New York, NY, USA, 2003. Association for Computing Machinery.
  • [22] George J. Minty. Monotone (nonlinear) operators in Hilbert space. Duke Math. J., 29:341–346, 1962.
  • [23] George J. Minty. on a “monotonicity” method for the solution of non-linear equations in Banach spaces. Proc. Nat. Acad. Sci. U.S.A., 50:1038–1041, 1963.
  • [24] John Nash. Non-cooperative games. Ann. of Math. (2), 54:286–295, 1951.
  • [25] John F. Nash, Jr. Equilibrium points in nn-person games. Proc. Nat. Acad. Sci. U.S.A., 36:48–49, 1950.
  • [26] J. B. Rosen. Existence and uniqueness of equilibrium points for concave nn-person games. Econometrica, 33:520–534, 1965.