Keywords

These keywords were added by machine and not by the authors. This process is experimental and the keywords may be updated as the learning algorithm improves.

FormalPara Introduction

This article is an attempt to present a brief introduction to viscosity solutions of first-order partial differential equations (PDE for short) and to review some aspects of the large time behavior of solutions of Hamilton–Jacobi equations with Neumann boundary conditions.

The notion of viscosity solution was introduced in [20] (see also [18]) by Crandall and Lions, and it has been widely accepted as the right notion of generalized solutions of the first-order PDE of the Hamilton–Jacobi type and fully nonlinear (possibly degenerate) elliptic or parabolic PDE. There have already been many nice contributions to overview of viscosity solutions of first-order and/or second-order partial differential equations. The following list touches just a few of them [2, 6, 15, 19, 29, 31, 41, 42].

This article is meant to serve as a quick introduction for graduate students or young researchers to viscosity solutions and is, of course, an outcome of the lectures delivered by the author at the CIME school as well as at Waseda University, Collège de France, Kumamoto University, King Abdulaziz University and University of Tokyo. For its easy readability, it contains some of very basics of mathematical analysis which are usually left aside to other textbooks.

The first section is an introduction to viscosity solutions of first-order partial differential equations. As a motivation to viscosity solutions we take up an optimal control problem and show that the value function of the control problem is characterized as a unique viscosity solution of the associated Bellman equation. This choice is essentially the same as used in the book [42] by Lions as well as in [2, 6, 29].

In Sects. 25, we develop the theory of viscosity solutions of Hamilton–Jacobi equations with the linear Neumann boundary condition together with the corresponding optimal control problems, which we follow [8, 38, 39]. In Sect. 6, following [38], we show the convergence of the solution of Hamilton–Jacobi equation of evolution type with the linear Neumann boundary condition to a solution of the stationary problem.

The approach here to the convergence result depends heavily on the variational formula for solutions, that is, the representation of solutions as the value function of the associated control problem. There is another approach, due to [3], based on the asymptotic monotonicity of a certain functional of the solutions as time goes to infinity, which is called the PDE approach. The PDE approach does not depend on the variational formula for the solutions and provides a very simple proof of the convergence with sharper hypotheses. The approach taken here may be called the dynamical or optimal control one. This approach requires the convexity of the Hamiltonian, so that one can associate it with an optimal control problem. Although it requires lots of steps before establishing the convergence result, its merit is that one can get an interpretation to the convergence result through the optimal control representation.

The topics covered in this article are very close to the ones discussed by Barles [4]. Both are to present an introduction to viscosity solutions and to discuss the large time asymptotics for solutions of Hamilton–Jacobi equations. This article has probably a more elementary flavor than [4] in the part of the introduction to viscosity solutions, and the paper [4] describes the PDE-viscosity approach to the large time asymptotics while this article concentrates on the dynamical or optimal control approach.

The reference list covers only those papers which the author more or less consulted while he was writing this article, and it is far from a complete list of those which have contributed to the developments of the subject.

The author would like to thank the course directors, Paola Loreti and Nicoletta Tchou, for their encouragement and patience while he was preparing this article.

He would also like to thank his colleagues and students for pointing out many misprints and mistakes in earlier versions of these lecture notes.

Notation:

  • When \(\mathcal{F}\) is a set of real-valued functions on X, \(\sup \mathcal{F}\) and \(\inf \mathcal{F}\) denote the functions on X given, respectively, by

    $$\displaystyle{(\sup \mathcal{F})(x) :=\sup \{ f(x)\, :\, f \in \mathcal{F}\}\ \ \text{ and }\ \ (\inf \mathcal{F})(x) :=\inf \{ f(x)\, :\, f \in \mathcal{F}\}.}$$
  • For any \(a,b \in \mathbb{R}\), we write a ∧ b = min{a, b} and \(a \vee b =\max \{ a,\,b\}\). Also, we write \(a_{+} = a \vee 0\) and \(a_{-} = (-a)_{+}\).

  • A function ω ∈ C([0, R)), with 0 < R ≤ , is called a modulus if it is nondecreasing and satisfies ω(0) = 0.

  • For any \(x = (x_{1},\ldots,x_{n}),y = (y_{1},\ldots,y_{n}) \in {\mathbb{R}}^{n}\), xy denotes the Euclidean inner product \(x_{1}y_{1} + \cdots + x_{n}y_{n}\) of x and y.

  • For any \(x,y \in {\mathbb{R}}^{n}\) the line segment between x and y is denoted by \([x,y] :=\{ (1 - t)x + ty\, :\, t \in [0,\,1]\}\).

  • For \(k \in \mathbb{N}\) and \(\Omega \subset {\mathbb{R}}^{n}\), \({C}^{k}(\Omega, {\mathbb{R}}^{m})\) (or simply, \({C}^{k}(\Omega, {\mathbb{R}}^{m})\)) denotes the collection of functions \(f\, :\, \Omega \rightarrow {\mathbb{R}}^{m}\) (not necessarily open), each of which has an open neighborhood U of Ω and a function g ∈ C k(U) such that f(x) = g(x) for all x ∈ Ω.

  • For \(f \in C(\Omega, {\mathbb{R}}^{m})\), where \(\Omega \subset {\mathbb{R}}^{n}\), the support of f is defined as the closure of {x ∈ Ω : f(x) ≠ 0} and is denoted by suppf.

  • UC(X) (resp., BUC(X)) denotes the space of all uniformly continuous (resp., bounded, uniformly continuous) functions in a metric space X.

  • We write 1 E for the characteristic function of the set E. That is, 1 E (x) = 1 if x ∈ E and 1 E (x) = 0 otherwise.

  • The sup-norm of function f on a set Ω is denoted by \(\|f\|_{\infty,\Omega } =\| f\|_{\infty } :=\sup _{\Omega }\vert f\vert \).

  • We write \(\mathbb{R}_{+}\) for the interval (0, ).

  • For any interval \(J \subset \mathbb{R}\), \(\mathrm{AC}(J, {\mathbb{R}}^{m})\) denotes the space of all absolutely continuous functions in J with value in \({\mathbb{R}}^{m}\).

  • Given a convex Hamiltonian \(H \in C(\overline{\Omega } \times {\mathbb{R}}^{n})\), where \(\Omega \subset {\mathbb{R}}^{n}\) is an open set, we denote by L the Lagrangian given by

    $$\displaystyle{L(x,\xi ) =\sup _{p\in {\mathbb{R}}^{n}}(\xi \cdot p - H(x,p))\ \ \ \text{ for }(x,\xi ) \in \overline{\Omega } \times {\mathbb{R}}^{n}.}$$
  • Let \(\Omega \subset {\mathbb{R}}^{n}\) be an open subset of \({\mathbb{R}}^{n}\), \(g \in C(\partial \Omega, \mathbb{R})\), t > 0 and \((\eta,v,l) \in {L}^{1}([0,\,t], {\mathbb{R}}^{n} \times {\mathbb{R}}^{n} \times \mathbb{R})\) such that \(\eta (s) \in \overline{\Omega }\) for all s ∈ [0, t] and l(s) = 0 whenever η(s) ∈ Ω. We write

    $$\displaystyle{\mathcal{L}(t,\eta,v,l) =\int _{ 0}^{t}[L(\eta (s),-v(s)) + g(\eta (s))l(s)]\mathrm{d}s.}$$

1 Introduction to Viscosity Solutions

We give the definition of viscosity solutions of first-order PDE and study their basic properties.

1.1 Hamilton–Jacobi Equations

Let Ω be an open subset of \({\mathbb{R}}^{n}\). Given a function \(H : \Omega \times {\mathbb{R}}^{n} \rightarrow \mathbb{R}\), we consider the PDE

$$\displaystyle{ H(x,\mathit{Du}(x)) = 0\ \ \mathrm{in}\ \Omega, }$$
(1)

where Du denotes the gradient of u, that is,

$$\displaystyle{\mathit{Du} := (u_{x_{1}},u_{x_{2}},\ldots,u_{x_{n}}) \equiv (\partial u/\partial x_{1},\ldots,\partial u/\partial x_{n}).}$$

We also consider the PDE

$$\displaystyle{ u_{t}(x,t) + H(x,D_{x}u(x,t)) = 0\ \ \mathrm{in}\ \Omega \times (0,\infty ). }$$
(2)

Here the variable t may be regarded as the time variable and u t denotes the time derivative ∂u ∕ ∂t. The variable x is then regarded as the space variable and D x u (or, Du) denotes the gradient of u in the space variable x.

The PDE of the type of (1) or (2) are called Hamilton–Jacobi equations. A more concrete example of (1) is given by

$$\displaystyle{\vert \mathit{Du}(x)\vert = k(x),}$$

which appears in geometrical optics and describes the surface front of propagating waves. Hamilton–Jacobi equations arising in Mechanics have the form

$$\displaystyle{\vert \mathit{Du}(x){\vert }^{2} + V (x) = 0,}$$

where the terms | Du(x) | 2 and V (x) correspond to the kinetic and potential energies, respectively.

More generally, the PDE of the form

$$\displaystyle{ F(x,u(x),\mathit{Du}(x)) = 0\ \ \mathrm{in}\ \Omega }$$
(3)

may be called Hamilton–Jacobi equations.

1.2 An Optimal Control Problem

We consider the function

$$\displaystyle{X = X(t) = (X_{1}(t),X_{2}(t),\ldots,X_{n}(t)) \in {\mathbb{R}}^{n}}$$

of time \(t \in \mathbb{R}\), and

$$\displaystyle{\dot{X} =\dot{ X}(t) = \frac{\mathrm{d}X} {\mathrm{d}t} (t)}$$

denotes its derivative. Let \(A \subset {\mathbb{R}}^{m}\) be a given set, let \(g : {\mathbb{R}}^{n} \times A \rightarrow {\mathbb{R}}^{n}\), \(f : {\mathbb{R}}^{n} \times A \rightarrow \mathbb{R}\) be given functions and λ > 0 be a given constant. We denote by A the set of all Lebesgue measurable α : [0, ) → A.

Fix any \(x \in {\mathbb{R}}^{n}\) and α ∈ A, and consider the initial value problem for the ordinary differential equation (for short, ODE)

$$\displaystyle{ \left \{\begin{array}{ll} &\dot{X}(t) = g(X(t),\alpha (t))\ \ \text{ for a.e. }t > 0,\\ &X(0) = x. \end{array} \right. }$$
(4)

The solution of (4) will be denoted by \(X = X(t) = X(t;x,\alpha )\). The solution X(t) may depend significantly on choices of α ∈ A. Next we introduce the functional

$$\displaystyle{ J(x,\alpha ) =\int _{ 0}^{\infty }f(X(t),\alpha (t))\mathrm{{e}}^{-\lambda t}\,\mathrm{d}t, }$$
(5)

a function of x and α ∈ A, which serves a criterion to decide which choice of α is better. The best value of the functional J is given by

$$\displaystyle{ V (x) =\inf _{\alpha \in \mathbf{A}}J(x,\alpha ). }$$
(6)

This is an optimization problem, and the main theme is to select a control α = α x  ∈ A so that

$$\displaystyle{V (x) = J(x,\alpha ).}$$

Such a control α is called an optimal control. The ODE in (4) is called the dynamics or state equation, the functional J given by (5) is called the cost functional, and the function V given by (6) is called the value function. The function f or t↦e − λt f(X(t), α(t)) is called the running cost and λ is called the discount rate.

In what follows, we assume that f, g are bounded continuous functions on \({\mathbb{R}}^{n} \times A\) and moreover, they satisfy the Lipschitz condition, i.e., there exists a constant M > 0 such that

$$\displaystyle{ \left \{\begin{array}{ll} &\vert f(x,a)\vert \leq M,\qquad \vert g(x,a)\vert \leq M, \\ &\vert f(x,a) - f(y,a)\vert \leq M\vert x - y\vert, \\ &\vert g(x,a) - g(y,a)\vert \leq M\vert x - y\vert. \end{array} \right. }$$
(7)

A basic result in ODE theory guarantees that the initial value problem (4) has a unique solution X(t).

There are two basic approaches in optimal control theory:

  1. 1.

    Pontryagin’s Maximum Principle Approach.

  2. 2.

    Bellman’s Dynamic Programming Approach.

Both of approaches have been introduced and developed since 1950s.

Pontryagin’s maximum principle gives a necessary condition for the optimality of controls and provides a powerful method to design an optimal control.

Bellman’s approach associates the optimization problem with a PDE, called the Bellman equation. In the problem, where the value function V is given by (6), the corresponding Bellman equation is the following.

$$\displaystyle{ \lambda V (x) + H(x,DV (x)) = 0\ \ \ \mathrm{in}\ {\mathbb{R}}^{n}, }$$
(8)

where H is a function given by

$$\displaystyle{H(x,p) =\sup _{a\in A}\{ - g(x,a) \cdot p - f(x,a)\},}$$

with xy denoting the Euclidean inner product in \({\mathbb{R}}^{n}\). Bellman’s idea is to characterize the value function V by the Bellman equation, to use the characterization to compute the value function and to design an optimal control. To see how it works, we assume that (8) has a smooth bounded solution V and compute formally as follows. First of all, we choose a function \(a\, :\, {\mathbb{R}}^{n} \rightarrow A\) so that

$$\displaystyle{H(x,DV (x)) = -g(x,a(x)) \cdot DV (x) - f(x,a(x)),}$$

and solve the initial value problem

$$\displaystyle{\dot{X}(t) = g(X(t),a(X(t))),\qquad X(0) = x,}$$

where x is a fixed point in \({\mathbb{R}}^{n}\). Next, writing α(t) = a(X(t)), we have

$$\displaystyle\begin{array}{rcl} 0& =& \,\int _{0}^{\infty }\mathrm{{e}}^{-\lambda t}\big(\lambda V (X(t)) + H(X(t),DV (X(t)))\big)\,\mathrm{d}t {}\\ & =& \,\int _{0}^{\infty }\mathrm{{e}}^{-\lambda t}\big(\lambda V (X(t)) - g(X(t),\alpha (t)) \cdot DV (X(t)) - f(X(t),\alpha (t))\big)\,\mathrm{d}t {}\\ & =& \,\int _{0}^{\infty }\left (-\frac{\mathrm{d}} {\mathrm{d}t}\mathrm{{e}}^{-\lambda t}V (X(t)) -\mathrm{ {e}}^{\lambda t}f(X(t),\alpha (t))\right )\,\mathrm{d}t {}\\ & =& \,V (X(0)) -\int _{0}^{\infty }\mathrm{{e}}^{-\lambda t}f(X(t),\alpha (t))\,\mathrm{d}t. {}\\ \end{array}$$

Thus we have

$$\displaystyle{V (x) = J(x,\alpha ).}$$

If PDE (8) characterizes the value function, that is, the solution V is the value function, then the above equality says that the control α(t) = a(X(t)) is an optimal control, which we are looking for.

In Bellman’s approach PDE plays a central role, and we discuss this approach in what follows. The first remark is that the value function may not be differentiable at some points. A simple example is as follows.

Example 1.1.

We consider the case where n = 1, \(A = [-1,1] \subset \mathbb{R}\), \(f(x,a) =\mathrm{ {e}}^{-{x}^{2} }\), g(x, a) = a and λ = 1. Let X(t) be the solution of (4) for some control α ∈ A, which means just to satisfy

$$\displaystyle{\vert \dot{X}(t)\vert \leq 1\ \ \ \ \text{ a.e. }t > 0.}$$

Let V be the value function given by (6). Then it is clear that \(V (-x) = V (x)\) for all \(x \in \mathbb{R}\) and that

$$\displaystyle{V (x) =\int _{ 0}^{\infty }\mathrm{{e}}^{-t-{(x+t)}^{2} }\,\mathrm{d}t =\mathrm{ {e}}^{x}\int _{ x}^{\infty }\mathrm{{e}}^{-t-{t}^{2} }\,\mathrm{d}t\ \ \ \mathrm{if}\ \ x > 0.}$$

For x > 0, one gets

$$\displaystyle{V ^{\prime}(x) =\mathrm{ {e}}^{x}\int _{ x}^{\infty }\mathrm{{e}}^{-t-{t}^{2} }\,\mathrm{d}t -\mathrm{ {e}}^{-{x}^{2} },}$$

and

$$\displaystyle{V ^{\prime}(0+) =\int _{ 0}^{\infty }\mathrm{{e}}^{-t-{t}^{2} }\,\mathrm{d}t - 1 <\int _{ 0}^{\infty }\mathrm{{e}}^{-t}\,\mathrm{d}t - 1 = 0.}$$

This together with the symmetry property, \(V (-x) = V (x)\) for all \(x \in \mathbb{R}\), shows that V is not differentiable at x = 0.

Value functions in optimal control do not have enough regularity to satisfy, in the classical sense, the corresponding Bellman equations in general as the above example shows.

We introduce the notion of viscosity solution of the first-order PDE

$$\displaystyle{ F(x,u(x),\mathit{Du}(x)) = 0\ \ \mathrm{in}\ \Omega, }$$
(FE)

where \(F\, :\, \Omega \times \mathbb{R} \times {\mathbb{R}}^{n} \rightarrow \mathbb{R}\) is a given continuous function.

Definition 1.1.

  1. (i)

    We call u ∈ C(Ω) a viscosity subsolution of (FE) if

    $$\displaystyle{\left \{\begin{array}{ll} &\phi \in {C}^{1}(\Omega ),\ z \in \Omega,\ \max _{ \Omega }(u-\phi ) = (u-\phi )(z) \\ &\Rightarrow F(z,u(z),D\phi (z)) \leq 0.\end{array} \right.}$$
  2. (ii)

    We call u ∈ C(Ω) a viscosity supersolution of (FE) if

    $$\displaystyle{\left \{\begin{array}{ll} &\phi \in {C}^{1}(\Omega ),\ z \in \Omega,\ \min _{ \Omega }(u-\phi ) = (u-\phi )(z) \\ &\Rightarrow F(z,u(z),D\phi (z)) \geq 0.\end{array} \right.}$$
  3. (iii)

    We call u ∈ C(Ω) a viscosity solution of (FE) if u is both a viscosity subsolution and supersolution of (FE).

The viscosity subsolution or supersolution property is checked through smooth functions ϕ in the above definition, and such smooth functions ϕ are called test functions.

Remark 1.1.

If we set \({F}^{-}(x,r,p) = -F(x,-r,-p)\), then it is obvious that u ∈ C(Ω) is a viscosity subsolution (resp., supersolution) of (FE) if and only if \({u}^{-}(x) := -u(x)\) is a viscosity supersolution (resp., subsolution) of

$$\displaystyle{{F}^{-}(x,{u}^{-}(x),{\mathit{Du}}^{-}(x)) = 0\ \ \mathrm{in}\ \Omega.}$$

Note also that \({({F}^{-})}^{-} = F\) and \({({u}^{-})}^{-} = u\). With these observations, one property for viscosity subsolutions can be phrased as a property for viscosity supersolutions. In other words, every proposition concerning viscosity subsolutions has a counterpart for viscosity supersolutions.

Remark 1.2.

It is easily seen by adding constants to test functions that u ∈ C(Ω) is a viscosity subsolution of (FE) if and only if

$$\displaystyle{\left \{\begin{array}{ll} &\phi \in {C}^{1}(\Omega ),\ z \in \Omega,\ \max _{ \Omega }(u-\phi ) = (u-\phi )(z) = 0 \\ &\Rightarrow F(z,\phi (z),D\phi (z)) \leq 0.\end{array} \right.}$$

One can easily formulate a counterpart of this proposition for viscosity supersolutions.

Remark 1.3.

It is easy to see by an argument based on a partition of unity (see Appendix A.1) that u ∈ C(Ω) is a viscosity subsolution of (FE) if and only if

$$\displaystyle{\left \{\begin{array}{ll} &\phi \in {C}^{1}(\Omega ),\ z \in \Omega,\ u -\phi \text{ attains a local maximum at }z \\ &\Rightarrow F(z,\phi (z),D\phi (z)) \leq 0.\end{array} \right.}$$

Remark 1.4.

It is easily seen that u ∈ C(Ω) is a viscosity subsolution of (FE) if and only if

$$\displaystyle{\left \{\begin{array}{ll} &\phi \in {C}^{1}(\Omega ),\ z \in \Omega,\ u -\phi \text{ attains a strict maximum at }z \\ &\Rightarrow F(z,\phi (z),D\phi (z)) \leq 0.\end{array} \right.}$$

Similarly, one may replace “strict maximum” by “strict local maximum” in the statement. The idea to show these is to replace the function ϕ by \(\phi (x) + \vert x - z{\vert }^{2}\) when needed.

Remark 1.5.

The condition, ϕ ∈ C 1(Ω), can be replaced by the condition, ϕ ∈ C (Ω) in the above definition. The argument in the following example explains how to see this equivalence.

Example 1.2 (Vanishing viscosity method). 

The term “viscosity solution” originates to the vanishing viscosity method, which is one of classical methods to construct solutions of first-order PDE.

Consider the second-order PDE

$$\displaystyle{ -\varepsilon \Delta {u}^{\varepsilon } + F(x,{u}^{\varepsilon }(x),{\mathit{Du}}^{\varepsilon }(x)) = 0\ \ \mathrm{in}\ \Omega, }$$
(9)

where \(\varepsilon > 0\) is a parameter to be sent to zero later on, Ω is an open subset of \({\mathbb{R}}^{n}\), F is a continuous function on \(\Omega \times \mathbb{R} \times {\mathbb{R}}^{n}\) and Δ denotes the Laplacian

$$\displaystyle{\Delta = \frac{{\partial }^{2}} {\partial x_{1}^{2}} + \cdots + \frac{{\partial }^{2}} {\partial x_{n}^{2}}.}$$

We assume that functions \({u}^{\varepsilon } \in {C}^{2}(\Omega )\), with \(\varepsilon \in (0,1)\), and u ∈ C(Ω) are given and that

$$\displaystyle{\lim _{\varepsilon \rightarrow 0}{u}^{\varepsilon }(x) = u(x)\ \ \ \ \text{locally uniformly on }\Omega.}$$

Then the claim is that u is a viscosity solution of

$$\displaystyle{ F(x,u(x),\mathit{Du}(x)) = 0\ \ \mathrm{in}\ \Omega. }$$
(FE)

In what follows, we just check that u is a viscosity subsolution of (FE). For this, we assume that

$$\displaystyle{\phi \in {C}^{1}(\Omega ),\quad \hat{x} \in \Omega,\quad \max _{ \Omega }(u-\phi ) = (u-\phi )(\hat{x}),}$$

and moreover, this maximum is a strict maximum of u − ϕ. We need to show that

$$\displaystyle{ F(\hat{x},u(\hat{x}),D\phi (\hat{x})) \leq 0. }$$
(10)

First of all, we assume that ϕ ∈ C 2(Ω), and show that (10) holds. Fix an r > 0 so that \(\overline{B}_{r}(\hat{x}) \subset \Omega \). Let \(x_{\varepsilon }\) be a maximum point over \(\overline{B}_{r}(\hat{x})\) of the function \({u}^{\varepsilon }-\phi\). We may choose a sequence \(\{\varepsilon _{j}\}_{j\in \mathbb{N}} \subset (0,\,1)\) so that \(\lim _{j\rightarrow \infty }\varepsilon _{j} = 0\) and \(\lim _{j\rightarrow \infty }x_{\varepsilon _{j}} = y\) for some \(y \in \overline{B}_{r}(\hat{x})\). Observe that

$$\displaystyle\begin{array}{rcl} (u-\phi )(\hat{x})& \leq & \,({u}^{\varepsilon _{j}}-\phi )(\hat{x}) +\| u - {u}^{\varepsilon _{j}}\|_{\infty,B_{ r}(\hat{x})} {}\\ & \leq & \,({u}^{\varepsilon _{j}}-\phi )(x_{\varepsilon _{ j}}) +\| u - {u}^{\varepsilon _{j} }\|_{\infty,B_{r}(\hat{x})} {}\\ & \leq & \,(u-\phi )(x_{\varepsilon _{j}}) + 2\|{u}^{\varepsilon _{j}} - u\|_{\infty,B_{ r}(\hat{x})} {}\\ & \rightarrow & \,(u-\phi )(y)\ \ \ \text{ as }j \rightarrow \infty. {}\\ \end{array}$$

Accordingly, since \(\hat{x}\) is a strict maximum point of u − ϕ, we see that \(y =\hat{ x}\). Hence, if j is sufficiently large, then \(x_{\varepsilon _{j}} \in B_{r}(\hat{x})\). By the maximum principle from Advanced Calculus, we find that

$$\displaystyle{ \frac{\partial } {\partial x_{i}}({u}^{\varepsilon _{j}}-\phi )(x_{\varepsilon _{ j}}) = 0\ \ \text{ and }\ \ \frac{{\partial }^{2}} {\partial x_{i}^{2}}({u}^{\varepsilon _{j}}-\phi )(x_{\varepsilon _{ j}}) \leq 0\ \ \text{ for all }\ i = 1,2,\ldots,n.}$$

Hence, we get

$$\displaystyle{{\mathit{Du}}^{\varepsilon _{j}}(x_{\varepsilon _{ j}}) = D\phi (x_{\varepsilon _{j}}),\quad \Delta {u}^{\varepsilon _{j} }(x_{\varepsilon _{j}}) \leq \Delta \phi (x_{\varepsilon _{j}}).}$$

These together with (9) yield

$$\displaystyle{-\varepsilon _{j}\Delta \phi (x_{\varepsilon _{j}}) + F(x_{\varepsilon _{j}},{u}^{\varepsilon _{j}}(x_{\varepsilon _{ j}}),D\phi (x_{\varepsilon _{j}})) \leq 0.}$$

Sending j →  now ensures that (10) holds.

Finally we show that the C 2 regularity of ϕ can be relaxed, so that (10) holds for all ϕ ∈ C 1(Ω). Let r > 0 be the constant as above, and choose a sequence \(\{\phi _{k}\} \subset {C}^{\infty }(\Omega )\) so that

$$\displaystyle{\lim _{k\rightarrow \infty }\phi _{k}(x) =\phi (x)\ \ \ \text{ uniformly on }B_{r}(\hat{x}).}$$

Let \(\{y_{k}\} \subset \overline{B}_{r}(\hat{x})\) be a sequence consisting of a maximum point of u − ϕ k . An argument similar to the above yields

$$\displaystyle{\lim _{k\rightarrow \infty }y_{k} =\hat{ x}.}$$

If k is sufficiently large, then we have \(y_{k} \in B_{r}(\hat{x})\) and, due to (10) valid for C 2 test functions,

$$\displaystyle{F(y_{k},u(y_{k}),D\phi _{k}(y_{k})) \leq 0.}$$

Sending k →  allows us to conclude that (10) holds.

1.3 Characterization of the Value Function

In this subsection we are concerned with the characterization of the value function V by the Bellman equation

$$\displaystyle{ \lambda V (x) + H(x,DV (x)) = 0\ \ \mathrm{in}\ {\mathbb{R}}^{n}, }$$
(11)

where λ is a positive constant and

$$\displaystyle{H(x,p) =\sup _{a\in A}\{ - g(x,a) \cdot p - f(x,a)\}.}$$

Recall that

$$\displaystyle{V (x) =\inf _{\alpha \in \mathbf{A}}J(x,\alpha ),}$$

and

$$\displaystyle{J(x,\alpha ) =\int _{ 0}^{\infty }f(X(t),\alpha (t))\mathrm{{e}}^{-\lambda t}\,\mathrm{d}t,}$$

where X(t) = X(t; x, α) denotes the solution of the initial value problem

$$\displaystyle{\left \{\begin{array}{ll} &\dot{X}(t) = g(X(t),\alpha (t))\ \ \text{ for a.e. }t > 0,\\ &X(0) = x. \end{array} \right.}$$

Recall also that for all \((x,a) \in {\mathbb{R}}^{n} \times A\) and some constant M > 0,

$$\displaystyle{ \left \{\begin{array}{ll} &\vert f(x,a)\vert \leq M,\qquad \vert g(x,a)\vert \leq M, \\ &\vert f(x,a) - f(y,a)\vert \leq M\vert x - y\vert, \\ &\vert g(x,a) - g(y,a)\vert \leq M\vert x - y\vert.\end{array} \right. }$$
(12)

The following lemma will be used without mentioning, the proof of which may be an easy exercise.

Lemma 1.1.

Let \(h,k\, :\, A \rightarrow \mathbb{R}\) be bounded functions. Then

$$\displaystyle{\left \vert \sup _{a\in A}h(a) -\sup _{a\in A}k(a)\right \vert \vee \left \vert \inf _{a\in A}h(a) -\,\inf _{a\in A}k(a)\right \vert \leq \sup _{a\in A}\vert h(a) - k(a)\vert.}$$

In view of the above lemma, the following lemma is an easy consequence of (12), and the detail of the proof is left to the reader.

Lemma 1.2.

The Hamiltonian H satisfies the following inequalities:

$$\displaystyle\begin{array}{rcl} & & \vert H(x,p) - H(y,p)\vert \leq M\vert x - y\vert (\vert p\vert + 1)\quad \text{ for all }\ x,y,p \in {\mathbb{R}}^{n}, {}\\ & & \vert H(x,p) - H(x,q)\vert \leq M\vert p - q\vert \qquad \qquad \,\,\,\text{ for all }\ x,p,q \in {\mathbb{R}}^{n}. {}\\ \end{array}$$

In particular, we have \(H \in C({\mathbb{R}}^{n} \times {\mathbb{R}}^{n})\).

Proposition 1.1.

The inequality

$$\displaystyle{\vert V (x)\vert \leq \frac{M} {\lambda } }$$

holds for all \(x \in {\mathbb{R}}^{n}\) . Hence, the value function V is bounded on \({\mathbb{R}}^{n}\).

Proof.

For any \((x,\alpha ) \in {\mathbb{R}}^{n} \times \mathbf{A}\), we have

$$\displaystyle{\vert J(x,\alpha )\vert \leq \int _{0}^{\infty }\mathrm{{e}}^{-\lambda t}\vert f(X(t),\alpha (t))\vert \,\mathrm{d}t \leq M\int _{ 0}^{\infty }\mathrm{{e}}^{-\lambda t}\,\mathrm{d}t = \frac{M} {\lambda }.}$$

Applying Lemma 1.1 yields

$$\displaystyle{\vert V (x)\vert \leq \sup _{\alpha \in \mathbf{A}}\vert J(x,\alpha )\vert \leq \frac{M} {\lambda }.}$$

 □ 

Proposition 1.2.

The function V is Hölder continuous on \({\mathbb{R}}^{n}\).

Proof.

Fix any \(x,y \in {\mathbb{R}}^{n}\). For any α ∈ A, we estimate the difference of J(x, α) and J(y, α). To begin with, we estimate the difference of X(t) : = X(t; x, α) and Y (t) : = X(t; y, α). Since

$$\displaystyle\begin{array}{rcl} \vert \dot{X}(t) -\dot{ Y }(t)\vert & =& \,\vert g(X(t),\alpha (t)) - g(Y (t),\alpha (t))\vert {}\\ &\leq & \,M\vert X(t) - Y (t)\vert \ \ \ \text{ for a.e. }t \geq 0, {}\\ \end{array}$$

we find that

$$\displaystyle\begin{array}{rcl} \vert X(t) - Y (t)\vert & \leq & \,\vert X(0) - Y (0)\vert +\int _{ 0}^{t}\vert \dot{X}(s) -\dot{ Y }(s)\vert \,\mathrm{d}s {}\\ & \leq & \,\vert x - y\vert + M\int _{0}^{t}\vert X(s) - Y (s)\vert \,\mathrm{d}s\ \ \ \text{ for all }\ t \geq 0. {}\\ \end{array}$$

By applying Gronwall’s inequality, we get

$$\displaystyle{\vert X(t) - Y (t)\vert \leq \vert x - y\vert \,\mathrm{{e}}^{\mathit{Mt}}\ \ \ \text{ for all }\ t \geq 0.}$$

Next, since

$$\displaystyle{\vert J(x,\alpha ) - J(y,\alpha )\vert \leq \int _{0}^{\infty }\mathrm{{e}}^{-\lambda s}\vert f(X(s),\alpha (s)) - f(Y (s),\alpha (s))\vert \,\mathrm{d}s,}$$

if λ > M, then we have

$$\displaystyle\begin{array}{rcl} \vert J(x,\alpha ) - J(y,\alpha )\vert & \leq & \int _{0}^{\infty }\mathrm{{e}}^{-\lambda s}M\vert X(s) - Y (s)\vert \,\mathrm{d}s {}\\ & \leq & M\int _{0}^{\infty }\mathrm{{e}}^{-\lambda s}\vert x - y\vert \mathrm{{e}}^{Ms}\,\mathrm{d}s = \frac{M\vert x - y\vert } {\lambda -M}, {}\\ \end{array}$$

and

$$\displaystyle{ \vert V (x) - V (y)\vert \leq \frac{M} {\lambda -M}\vert x - y\vert. }$$
(13)

If 0 < λ < M, then we select 0 < θ < 1 so that θM < λ, and calculate

$$\displaystyle\begin{array}{rcl} \vert f(\xi,a) - f(\eta,a)\vert & \leq & \,\vert f(\xi,a) - f(\eta,a){\vert }^{\theta +(1-\theta )} {}\\ & \leq & \,{(M\vert \xi -\eta \vert )}^{\theta }{(2M)}^{1-\theta }\ \ \text{ for all }\ \xi,\eta \in {\mathbb{R}}^{n},a \in A, {}\\ \end{array}$$

and

$$\displaystyle\begin{array}{rcl} \vert J(x,\alpha ) - J(y,\alpha )\vert & \leq & \,{(2M)}^{1-\theta }\int _{ 0}^{\infty }\mathrm{{e}}^{-\lambda s}{(M\vert X(s) - Y (s)\vert )}^{\theta }\,\mathrm{d}s {}\\ & \leq & \,{(2M)}^{1-\theta }\int _{ 0}^{\infty }\mathrm{{e}}^{-\lambda s}{(M\vert x - y\vert )}^{\theta }\mathrm{{e}}^{\theta Ms}\,\mathrm{d}s {}\\ & \leq & \,2M\vert x - y{\vert }^{\theta }\int _{0}^{\infty }\mathrm{{e}}^{-(\lambda -\theta M)s}\,\mathrm{d}s = \frac{2M\vert x - y{\vert }^{\theta }} {\lambda -\theta M}, {}\\ \end{array}$$

which shows that

$$\displaystyle{ \vert V (x) - V (y)\vert \leq \frac{2M\vert x - y{\vert }^{\theta }} {\lambda -\theta M}. }$$
(14)

Thus we conclude from (13) and (14) that V is Hölder continuous on \({\mathbb{R}}^{n}\). □ 

Proposition 1.3 (Dynamic programming principle). 

Let 0 < τ < ∞ and \(x \in {\mathbb{R}}^{n}\) . Then

$$\displaystyle{V (x) =\inf _{\alpha \in \mathbf{A}}\Big(\int _{0}^{\tau }\mathrm{{e}}^{-\lambda t}f(X(t),\alpha (t))\,\mathrm{d}t +\mathrm{ {e}}^{-\lambda \tau }V (X(\tau ))\Big),}$$

where X(t) denotes X(t;x,α).

Proof.

Let 0 < τ <  and \(x \in {\mathbb{R}}^{n}\). Fix γ ∈ A. We have

$$\displaystyle\begin{array}{rcl} J(x,\gamma )& =& \,\int _{0}^{\tau }\mathrm{{e}}^{-\lambda t}f(X(t),\gamma (t))\,\mathrm{d}t +\int _{ \tau }^{\infty }\mathrm{{e}}^{-\lambda t}f(X(t),\gamma (t))\,\mathrm{d}t \\ & =& \,\int _{0}^{\tau }\mathrm{{e}}^{-\lambda t}f(X(t),\alpha (t))\,\mathrm{d}t +\mathrm{ {e}}^{-\lambda \tau }\int _{ 0}^{\infty }\mathrm{{e}}^{-\lambda t}f(Y (t),\beta (t))\,\mathrm{d}t,{}\end{array}$$
(15)

where

$$\displaystyle\begin{array}{rcl} & & X(t) = X(t;x,\gamma ),\quad \alpha (t) :=\gamma (t),\quad \beta (t) :=\gamma (t+\tau ), {}\\ & & Y (t) := X(t+\tau ) = X(t;X(\tau ),\beta ). {}\\ \end{array}$$

By (15), we get

$$\displaystyle{J(x,\gamma ) \geq \int _{0}^{\tau }\mathrm{{e}}^{-\lambda t}f(X(t),\alpha (t))\,\mathrm{d}t +\mathrm{ {e}}^{-\lambda \tau }V (X(\tau )),}$$

from which we have

$$\displaystyle{J(x,\gamma ) \geq \inf _{\alpha \in \mathbf{A}}\Big(\int _{0}^{\tau }\mathrm{{e}}^{-\lambda t}f(X(t),\alpha (t))\,\mathrm{d}t +\mathrm{ {e}}^{-\lambda \tau }V (X(\tau ))\Big).}$$

Consequently,

$$\displaystyle{ V (x) \geq \inf _{\alpha \in \mathbf{A}}\Big(\int _{0}^{\tau }\mathrm{{e}}^{-\lambda t}f(X(t),\alpha (t))\,\mathrm{d}t +\mathrm{ {e}}^{-\lambda \tau }V (X(\tau ))\Big). }$$
(16)

Now, let α, β ∈ A. Define γ ∈ A by

$$\displaystyle{\gamma (t) = \left \{\begin{array}{@{}l@{\quad }l@{}} \alpha (t) \quad &\mathrm{if}\ \ 0 \leq t \leq \tau,\\ \beta (t-\tau )\quad &\mathrm{if } \ \ \tau < t. \end{array} \right.}$$

Set

$$\displaystyle{X(t) := X(t;x,\alpha )\ \ \ \text{ and }\ \ \ Y (t) := X(t;X(\tau ),\beta ).}$$

We have

$$\displaystyle{\left \{\begin{array}{ll} &X(t) = X(t;x,\gamma )\ \ \text{ and }\ \ \alpha (t) =\gamma (t)\ \ \ \text{ for all }\ t \in [0,\,\tau ],\\ &\beta (t) =\gamma (t+\tau )\ \ \text{ and } \ \ Y (t) = X(t+\tau )\ \ \ \text{ for all } \ t \geq 0. \end{array} \right.}$$

Hence, we have (15) and therefore,

$$\displaystyle{V (x) \leq \int _{0}^{\tau }\mathrm{{e}}^{-\lambda t}f(X(t),\alpha (t))\,\mathrm{d}t +\mathrm{ {e}}^{-\lambda \tau }J(X(\tau ),\beta ).}$$

Moreover, we get

$$\displaystyle\begin{array}{rcl} V (x) \leq \int _{0}^{\tau }\mathrm{{e}}^{-\lambda t}f(X(t),\alpha (t))\,\mathrm{d}t +\mathrm{ {e}}^{-\lambda \tau }V (X(\tau )),& & {}\\ \end{array}$$

and

$$\displaystyle\begin{array}{rcl} V (x) \leq \inf _{\alpha \in \mathbf{A}}\Big(\int _{0}^{\tau }\mathrm{{e}}^{-\lambda t}f(X(t),\alpha (t))\,\mathrm{d}t +\mathrm{ {e}}^{-\lambda \tau }V (X(\tau ))\Big).& &{}\end{array}$$
(17)

Combining (16) and (17) completes the proof. □ 

Theorem 1.1.

The value function V is a viscosity solution of (11).

Proof. (Subsolution property) Let \(\phi \in {C}^{1}({\mathbb{R}}^{n})\) and \(\hat{x} \in {\mathbb{R}}^{n}\), and assume that

$$\displaystyle{(V -\phi )(\hat{x}) =\max _{{\mathbb{R}}^{n}}(V -\phi ) = 0.}$$

Fix any a ∈ A and set α(t) : = a, \(X(t) := X(t;\hat{x},\alpha )\). Let 0 < h < . Now, since V ≤ ϕ, \(V (\hat{x}) =\phi (\hat{x})\), by Proposition 1.3 we get

$$\displaystyle\begin{array}{rcl} \phi (\hat{x})& =& \,V (\hat{x}) \leq \int _{0}^{h}\mathrm{{e}}^{-\lambda t}f(X(t),\alpha (t))\,\mathrm{d}t +\mathrm{ {e}}^{-\lambda h}V (X(h)) {}\\ & \leq & \,\int _{0}^{h}\mathrm{{e}}^{-\lambda t}f(X(t),\alpha (t))\,\mathrm{d}t +\mathrm{ {e}}^{-\lambda h}\phi (X(h)). {}\\ \end{array}$$

From this, we get

$$\displaystyle\begin{array}{rcl} 0& \leq & \,\int _{0}^{h}\mathrm{{e}}^{-\lambda t}f(X(t),a)\,\mathrm{d}t +\int _{ 0}^{h} \frac{\mathrm{d}} {\mathrm{d}t}(\mathrm{{e}}^{-\lambda t}\phi (X(t)))\,\mathrm{d}t \\ & =& \,\int _{0}^{h}\mathrm{{e}}^{-\lambda t}\left (f(X(t),a) -\lambda \phi (X(t)) + D\phi (X(t)) \cdot \dot{ X}(t)\right )\,\mathrm{d}t \\ & =& \,\int _{0}^{h}\mathrm{{e}}^{-\lambda t}\big(f(X(t),a) -\lambda \phi (X(t)) + D\phi (X(t)) \cdot g(X(t),a)\big)\,\mathrm{d}t.{}\end{array}$$
(18)

Noting that

$$\displaystyle{ \vert X(t) -\hat{ x}\vert =\Big \vert \int _{0}^{t}\dot{X}(s)\,\mathrm{d}s\Big\vert \leq \int _{ 0}^{t}\vert g(X(s),a)\vert \,\mathrm{d}s \leq M\int _{ 0}^{t}\,\mathrm{d}s = \mathit{Mt}, }$$
(19)

dividing (18) by h and sending h → 0, we find that

$$\displaystyle{0 \leq -\lambda \phi (\hat{x}) + f(\hat{x},a) + g(\hat{x},a) \cdot D\phi (\hat{x}).}$$

Since a ∈ A is arbitrary, we have \(\lambda \phi (\hat{x}) + H(\hat{x},D\phi (\hat{x})) \leq 0\).

(Supersolution property) Let \(\phi \in {C}^{1}({\mathbb{R}}^{n})\) and \(\hat{x} \in {\mathbb{R}}^{n}\), and assume that

$$\displaystyle{(V -\phi )(\hat{x}) =\min _{{\mathbb{R}}^{n}}(V -\phi ) = 0.}$$

Fix \(\varepsilon > 0\) and h > 0. By Proposition 1.3, we may choose α ∈ A so that

$$\displaystyle{V (\hat{x}) +\varepsilon h >\int _{ 0}^{h}\mathrm{{e}}^{-\lambda t}f(X(t),\alpha (t))\,\mathrm{d}t +\mathrm{ {e}}^{-\lambda h}V (X(h)),}$$

where \(X(t) := X(t;\hat{x},\alpha )\). Since V ≥ ϕ in \({\mathbb{R}}^{n}\) and \(V (\hat{x}) =\phi (\hat{x})\), we get

$$\displaystyle{\phi (\hat{x}) +\varepsilon h >\int _{ 0}^{h}\mathrm{{e}}^{-\lambda t}f(X(t),\alpha (t))\,\mathrm{d}t +\mathrm{ {e}}^{-\lambda h}\phi (X(h)).}$$

Hence we get

$$\displaystyle\begin{array}{rcl} 0& \geq & \,\int _{0}^{h}\mathrm{{e}}^{-\lambda t}f(X(t),\alpha (t))\,\mathrm{d}t +\int _{ 0}^{h} \frac{\mathrm{d}} {\mathrm{d}t}(\mathrm{{e}}^{-\lambda t}\phi (X(t)))\,\mathrm{d}t -\varepsilon h {}\\ & =& \,\int _{0}^{h}\mathrm{{e}}^{-\lambda t}\left (f(X(t),\alpha (t)) -\lambda \phi (X(t)) + D\phi (X(t)) \cdot \dot{ X}(t)\right )\,\mathrm{d}t -\varepsilon h {}\\ & =& \,\int _{0}^{h}\mathrm{{e}}^{-\lambda t}\big(f(X(t),\alpha (t)) -\lambda \phi (X(t)) + D\phi (X(t)) \cdot g(X(t),\alpha (t))\big)\,\mathrm{d}t -\varepsilon h. {}\\ \end{array}$$

By the definition of H, we get

$$\displaystyle{ \int _{0}^{h}\mathrm{{e}}^{-\lambda t}(\lambda \phi (X(t)) + H(X(t),D\phi (t))\,\mathrm{d}t +\varepsilon h > 0. }$$
(20)

As in (19), we have

$$\displaystyle{\vert X(t) -\hat{ x}\vert \leq \mathit{Mt}.}$$

Dividing (20) by h and sending h → 0 yield

$$\displaystyle{\lambda \phi (\hat{x}) + H(\hat{x},D\phi (\hat{x}))+\varepsilon \geq 0,}$$

from which we get \(\lambda \phi (\hat{x}) + H(\hat{x},D\phi (\hat{x})) \geq 0\). The proof is now complete. □ 

Theorem 1.2.

Let \(u \in \mathrm{BUC}({\mathbb{R}}^{n})\) and \(v \in \mathrm{BUC}({\mathbb{R}}^{n})\) be a viscosity subsolution and supersolution of (11) , respectively. Then u ≤ v in \({\mathbb{R}}^{n}\).

Proof.

Let \(\varepsilon > 0\), and define \(u_{\varepsilon } \in C({\mathbb{R}}^{n})\) by \(u_{\varepsilon }(x) = u(x) -\varepsilon (\langle x\rangle + M)\), where \(\langle x\rangle = {(\vert x{\vert }^{2} + 1)}^{1/2}\). A formal calculation

$$\displaystyle\begin{array}{rcl} u_{\varepsilon }(x) + H(x,\mathit{Du}_{\varepsilon }(x))& \leq & \,u(x) -\varepsilon M + H(x,\mathit{Du}(x)) +\varepsilon M\vert D\langle x\rangle \vert {}\\ &\leq & \,u(x) + H(x,\mathit{Du}(x)) \leq 0 {}\\ \end{array}$$

reveals that \(u_{\varepsilon }\) is a viscosity subsolution of (11), which can be easily justified.

We show that the inequality \(u_{\varepsilon } \leq v\) holds, from which we deduce that u ≤ v is valid. To do this, we assume that \(\sup _{{\mathbb{R}}^{n}}(u_{\varepsilon } - v) > 0\) and will get a contradiction. Since

$$\displaystyle{\lim _{\vert x\vert \rightarrow \infty }(u_{\varepsilon } - v)(x) = -\infty,}$$

we may choose a constant R > 0 so that

$$\displaystyle{\sup _{{\mathbb{R}}^{n}\setminus B_{R}}(u_{\varepsilon } - v) < 0.}$$

The function \(u_{\varepsilon } - v \in C(\overline{B}_{R})\) then attains a maximum at a point in B R , but not at any point in ∂B R .

Let α > 1 and consider the function

$$\displaystyle{\Phi (x,y) = u_{\varepsilon }(x) - v(y) -\alpha \vert x - y{\vert }^{2}}$$

on \(K := \overline{B}_{R} \times \overline{B}_{R}\). Since Φ ∈ C(K), Φ attains a maximum at a point in K. Let \((x_{\alpha },y_{\alpha }) \in K\) be its maximum point. Because K is compact, we may choose a sequence {α j } ⊂ (1, ) diverging to infinity so that for some \((\hat{x},\hat{y}) \in K\),

$$\displaystyle{(x_{\alpha _{j}},y_{\alpha _{j}}) \rightarrow (\hat{x},\hat{y})\ \ \ \text{ as }j \rightarrow \infty.}$$

Note that

$$\displaystyle\begin{array}{rcl} 0<\max _{\overline{B}_{ R}}(u_{\varepsilon } - v) =\max _{x\in \overline{B}_{ R}}\Phi (x,x) \leq \Phi (x_{\alpha },y_{\alpha }) \\ & & = u_{\varepsilon }(x_{\alpha }) - v(y_{\alpha }) -\alpha \vert x_{\alpha } - y_{\alpha }{\vert }^{2}, {}\end{array}$$
(21)

from which we get

$$\displaystyle{\alpha \vert x_{\alpha } - y_{\alpha }{\vert }^{2} \leq \sup _{{ \mathbb{R}}^{n}}u_{\varepsilon } +\sup _{{\mathbb{R}}^{n}}(-v).}$$

We infer from this that \(\hat{x} =\hat{ y}\). Once again by (21), we get

$$\displaystyle{\max _{\overline{B}_{ R}}(u_{\varepsilon } - v) \leq u_{\varepsilon }(x_{\alpha }) - v(y_{\alpha }).}$$

Setting α = α j and sending j →  in the above, since \(u,\,v \in C({\mathbb{R}}^{n})\), we see that

$$\displaystyle\begin{array}{rcl} \max _{\overline{B}_{ R}}(u_{\varepsilon } - v)& \leq & \lim _{\alpha =\alpha _{j},j\rightarrow \infty }(u_{\varepsilon }(x_{\alpha }) - v(y_{\alpha })) {}\\ & =& u_{\varepsilon }(\hat{x}) - v(\hat{x}). {}\\ \end{array}$$

That is, the point \(\hat{x}\) is a maximum point of \(u_{\varepsilon } - v\). By (21), we have

$$\displaystyle{\alpha \vert x_{\alpha } - y_{\alpha }{\vert }^{2} \leq u_{\varepsilon }(x_{\alpha }) - v(y_{\alpha }) -\max _{ \overline{B}_{R}}(u - v),}$$

and hence

$$\displaystyle{\lim _{\alpha =\alpha _{j},j\rightarrow \infty }\alpha \vert x_{\alpha } - y_{\alpha }{\vert }^{2} = 0.}$$

Since \(\hat{x}\) is a maximum point of \(u_{\varepsilon } - v\), by our choice of R we see that \(\hat{x} \in B_{R}\). Accordingly, if α = α j and j is sufficiently large, then \(x_{\alpha },y_{\alpha } \in B_{R}\). By the viscosity property of \(u_{\varepsilon }\) and v, for α = α j and \(j \in \mathbb{N}\) large enough, we have

$$\displaystyle{u_{\varepsilon }(x_{\alpha }) + H(x_{\alpha },2\alpha (x_{\alpha } - y_{\alpha })) \leq 0,\qquad v(y_{\alpha }) + H(y_{\alpha },2\alpha (x_{\alpha } - y_{\alpha })) \geq 0.}$$

Subtracting one from the other yields

$$\displaystyle{u_{\varepsilon }(x_{\alpha }) - v(y_{\alpha }) \leq H(y_{\alpha },2\alpha (x_{\alpha } - y_{\alpha })) - H(x_{\alpha },2\alpha (x_{\alpha } - y_{\alpha })).}$$

Using one of the properties of H from Lemma 1.2, we obtain

$$\displaystyle{u_{\varepsilon }(x_{\alpha }) - v(y_{\alpha }) \leq M\vert x_{\alpha } - y_{\alpha }\vert (2\alpha \vert x_{\alpha } - y_{\alpha }\vert + 1).}$$

Sending α = α j  → , we get

$$\displaystyle{u_{\varepsilon }(\hat{x}) - v(\hat{x}) \leq 0,}$$

which is a contradiction. □ 

1.4 Semicontinuous Viscosity Solutions and the Perron Method

Let u, v ∈ C(Ω) be a viscosity subsolutions of (FE) and set

$$\displaystyle{w(x) =\max \{ u(x),v(x)\}\ \ \text{ for }\ x \in \Omega.}$$

It is easy to see that w is a viscosity subsolution of (FE). Indeed, if ϕ ∈ C 1(Ω), y ∈ Ω and w − ϕ has a maximum at y, then we have either w(y) = u(y) and \((u-\phi )(x) \leq (w-\phi )(x) \leq (w-\phi )(y) = (u-\phi )(y)\) for all x ∈ Ω, or w(y) = v(y) and \((v-\phi )(x) \leq (v-\phi )(y)\), from which we get F(y, w(y), (y)) ≤ 0. If \(\{u_{k}\}_{k\in \mathbb{N}} \subset C(\Omega )\) is a uniformly bounded sequence of viscosity subsolutions of (FE), then the function w given by \(w(x) =\sup _{k}u_{k}(x)\) defines a bounded function on Ω but it may not be continuous, a situation that the notion of viscosity subsolution does not apply.

We are thus led to extend the notion of viscosity solution to that for discontinuous functions.

Let \(U \subset {\mathbb{R}}^{n}\), and recall that a function \(f : U \rightarrow \mathbb{R} \cup \{-\infty,\infty \} = [-\infty,\,\infty ]\) is upper semicontinuous if

$$\displaystyle{\limsup _{y\rightarrow x}f(y) \leq f(x)\ \ \text{ for all }\ x \in U.}$$

The totality of all such upper semicontinuous functions f will be denoted by USC(U). Similarly, we denote by LSC(U) the space of all lower semicontinuous functions on U. That is, \(\mathrm{LSC}(U) := -\mathrm{USC}(U) =\{ -f\, :\, f \in \mathrm{USC}(U)\}\).

Some basic observations regarding semicontinuity are the following three propositions.

Proposition 1.4.

Let f : U → [−∞, ∞]. Then, f ∈USC (U) if and only if the set {x ∈ U : f(x) < a} is a relatively open subset of U for any \(a \in \mathbb{R}\).

Proposition 1.5.

If   \(\mathcal{F}\subset \mathrm{LSC}(U)\) , then   \(\sup \mathcal{F}\in \mathrm{LSC}(U)\) . Similarly, if   \(\mathcal{F}\subset \mathrm{USC}(U)\) , then   \(\inf \mathcal{F}\in \mathrm{USC}(U)\).

Proposition 1.6.

Let K be a compact subset of \({\mathbb{R}}^{n}\) and f ∈USC (K). Then f attains a maximum. Here the maximum value may be either −∞ or ∞.

Next, we define the upper (resp., lower) semicontinuous envelopesf  ∗  (resp., f  ∗ ) of f : U → [ − , ] by

$$\displaystyle{{f}^{{\ast}}(x) =\lim _{ r\rightarrow 0+}\sup \{f(y)\, :\, y \in U \cap B_{r}(x)\}}$$

(resp., \(f_{{\ast}} = -{(-f)}^{{\ast}}\)  or, equivalently,  \(f_{{\ast}}(x) =\lim _{r\rightarrow 0+}\inf \{f(y)\, :\, y \in U \cap B_{r}(x)\}\)).

Proposition 1.7.

Let f : U → [−∞, ∞]. Then we have f USC (U), f LSC (U) and

$$\displaystyle{{f}^{{\ast}}(x) =\min \{ g(x)\, :\, g \in \mathrm{USC}(U),\ g \geq f\}\ \ \text{ for all }\ x \in U.}$$

A consequence of the above proposition is that if f ∈ USC(U), then f  ∗  = f in U. Similarly, f  ∗  = f in U if f ∈ LSC(U).

We go back to

$$\displaystyle{ F(x,u(x),\mathit{Du}(x)) = 0\ \ \mathrm{in}\ \Omega. }$$
(FE)

Here we assume neither that \(F\, :\, \Omega \times \mathbb{R} \times {\mathbb{R}}^{n} \rightarrow \mathbb{R}\) is continuous nor that \(\Omega \subset {\mathbb{R}}^{n}\) is open. We just assume that \(F\, :\, \Omega \times \mathbb{R} \times {\mathbb{R}}^{n} \rightarrow \mathbb{R}\) is locally bounded and that Ω is a subset of \({\mathbb{R}}^{n}\).

Definition 1.2.

  1. (i)

    A locally bounded function \(u\, :\, \Omega \rightarrow \mathbb{R}\) is called a viscosity subsolution (resp., supersolution) of (FE) if

    $$\displaystyle{\left \{\begin{array}{ll} &\phi \in {C}^{1}(\Omega ),\ z \in \Omega,\ \max _{ \Omega }({u}^{{\ast}}-\phi ) = ({u}^{{\ast}}-\phi )(z) \\ &\Rightarrow F_{{\ast}}(z,{u}^{{\ast}}(z),D\phi (z)) \leq 0\end{array} \right.}$$
    $$\displaystyle{\left (resp.,\quad \left \{\begin{array}{ll} &\phi \in {C}^{1}(\Omega ),\ z \in \Omega,\ \min _{ \Omega }(u_{{\ast}}-\phi ) = (u_{{\ast}}-\phi )(z) \\ &\Rightarrow{F}^{{\ast}}(z,u_{{\ast}}(z),D\phi (z)) \geq 0\end{array} \right.\quad \right ).}$$
  2. (ii)

    A locally bounded function \(u\, :\, \Omega \rightarrow \mathbb{R}\) is a viscosity solution of (FE) if it is both a viscosity subsolution and supersolution of (FE).

We warn here that the envelopes F  ∗  and F  ∗  are taken in the full variables. For instance, if \(\xi \in \Omega \times \mathbb{R} \times {\mathbb{R}}^{n}\), then

$$\displaystyle{F_{{\ast}}(\xi ) =\lim _{r\rightarrow 0+}\inf \{F(\eta )\, :\,\eta \in \Omega \times \mathbb{R} \times {\mathbb{R}}^{n},\ \vert \eta -\xi \vert < r\}.}$$

We say conveniently that u is a viscosity solution (or subsolution) of F(x, u(x), Du(x)) ≤ 0 in Ω if u is a viscosity subsolution of (FE). Similarly, we say that u is a viscosity solution (or supersolution) of F(x, u(x), Du(x)) ≥ 0 in Ω if u is a viscosity supersolution of (FE). Also, we say that u satisfies F(x, u(x), Du(x)) ≤ 0 in Ω (resp., F(x, u(x), Du(x)) ≥ 0 in Ω) in the viscosity sense if u is a viscosity subsolution (resp., supersolution) of (FE).

Once we fix a PDE, like (FE), on a set Ω, we denote by \({\mathcal{S}}^{-}\) and \({\mathcal{S}}^{+}\) the sets of all its viscosity subsolutions and supersolutions, respectively.

The above definition differs from the one in [19]. As is explained in [19], the above one allows the following situation: let Ω be a nonempty open subset of \({\mathbb{R}}^{n}\) and suppose that the Hamilton–Jacobi equation (1) has a continuous solution u ∈ C(Ω). Choose two dense subsets U and V of Ω such that U ∩ V =  and UV ≠ Ω. Select a function \(v\, :\, \Omega \rightarrow \mathbb{R}\) so that v(x) = u(x) if x ∈ U, \(v(x) = u(x) + 1\) if x ∈ V and v(x) ∈ [u(x), u(x) + 1] if x ∈ Ω ∖ (UV ). Then we have v  ∗ (x) = u(x) and \({v}^{{\ast}}(x) = u(x) + 1\) for all x ∈ Ω. Consequently, v is a viscosity solution of (1). If UV ≠ Ω, then there are infinitely many choices of such functions v.

The same remarks as Remarks 1.1–1.4 are valid for the above generalized definition.

Definition 1.3.

Let \(\Omega \subset {\mathbb{R}}^{n}\) and \(u\, :\, \Omega \rightarrow \mathbb{R}\). The subdifferential D  −  u(x) and superdifferential D  +  u(x) of the function u at x ∈ Ω are defined, respectively, by

$$\displaystyle\begin{array}{rcl}{ D}^{-}u(x)& =& \,\{p \in {\mathbb{R}}^{n}\, :\, u(x + h) \geq u(x) + p \cdot h + o(\vert h\vert )\ \text{ as }\ x + h \in \Omega,\,h \rightarrow 0\}, {}\\ {D}^{+}u(x)& =& \,\{p \in {\mathbb{R}}^{n}\, :\, u(x + h) \leq u(x) + p \cdot h + o(\vert h\vert )\ \text{ as }\ x + h \in \Omega,\,h \rightarrow 0\}, {}\\ \end{array}$$

where o( | h | ) denotes a function on an interval (0, δ), with δ > 0, having the property: \(\lim _{h\rightarrow 0}o(\vert h\vert )/\vert h\vert = 0\).

We remark that \({D}^{-}u(x) = -{D}^{+}(-u)(x)\). If u is a convex function in \({\mathbb{R}}^{n}\) and p ∈ D  −  u(x) for some \(x,p \in {\mathbb{R}}^{n}\), then

$$\displaystyle{u(x + h) \geq u(x) + p \cdot h\ \ \ \text{ for all }\ h \in {\mathbb{R}}^{n}.}$$

See Proposition B.1 for the above claim. In convex analysis, D  −  u(x) is usually denoted by ∂u(x).

Proposition 1.8.

Let \(\Omega \subset {\mathbb{R}}^{n}\) and \(u\, :\, \Omega \rightarrow \mathbb{R}\) be locally bounded. Let x ∈ Ω. Then

$$\displaystyle{{D}^{+}u(x) =\{ D\phi (x)\, :\,\phi \in {C}^{1}(\Omega ),\ u -\phi \text{ attains a maximum at }x\}.}$$

As a consequence of the above proposition, we have the following: if u is locally bounded in Ω, then

$$\displaystyle\begin{array}{rcl}{ D}^{-}u(x)& =& \,-{D}^{+}(-u)(x) {}\\ & =& \,-\{D\phi (x)\, :\,\phi \in {C}^{1}(\Omega ),\ -u -\phi \text{ attains a maximum at }x\} {}\\ & =& \,\{D\phi (x)\, :\,\phi \in {C}^{1}(\Omega ),\ u -\phi \text{ attains a minimum at }x\}. {}\\ \end{array}$$

Corollary 1.1.

Let \(\Omega \subset {\mathbb{R}}^{n}\) . Let \(F\, :\, \Omega \times \mathbb{R} \times {\mathbb{R}}^{n} \rightarrow \mathbb{R}\) and \(u\, :\, \Omega \rightarrow \mathbb{R}\) be locally bounded. Then u is a viscosity subsolution (resp., supersolution) of (FE) if and only if

$$\displaystyle\begin{array}{rcl} F_{{\ast}}(x,{u}^{{\ast}}(x),p) \leq & \,0\ \ \ \text{ for all }x \in \Omega,\,p \in {D}^{+}{u}^{{\ast}}(x) & {}\\ (\text{ resp., }\ {F}^{{\ast}}(x,u_{ {\ast}}(x),p) \geq & \,0\ \ \ \ \text{ for all }x \in \Omega,\,p \in {D}^{-}u_{ {\ast}}(x)\ ).& {}\\ \end{array}$$

This corollary (or Remark 1.3) says that the viscosity properties of a function, i.e., the properties that the function be a viscosity subsolution, supersolution, or solution are of local nature. For instance, under the hypotheses of Corollary 1.1, the function u is a viscosity subsolution of (FE) if and only if for each x ∈ Ω there exists an open neighborhood U x , in \({\mathbb{R}}^{n}\), of x such that u is a viscosity subsolution of (FE) in U x  ∩ Ω.

Proof.

Let ϕ ∈ C 1(Ω) and y ∈ Ω, and assume that u − ϕ has a maximum at y. Then

$$\displaystyle{(u-\phi )(y + h) \leq (u-\phi )(y)\ \ \mathrm{if}\ y + h \in \Omega,}$$

and hence, as y + h ∈ Ω,  h → 0,

$$\displaystyle{u(y + h) \leq u(y) +\phi (y + h) -\phi (y) = u(y) + D\phi (y) \cdot h + o(\vert h\vert ).}$$

This shows that

$$\displaystyle{\{D\phi (y)\, :\,\phi \in {C}^{1}(\Omega ),\ u -\phi \text{ attains a maximum at }y\} \subset {D}^{+}u(y).}$$

Next let y ∈ Ω and p ∈ D  +  u(y). Then we have

$$\displaystyle{u(y + h) \leq u(y) + p \cdot h +\omega (\vert h\vert )\vert h\vert \ \ \mathrm{if}\ y + h \in \Omega \text{ and }\vert h\vert \leq \delta }$$

for some constant δ > 0 and a function ω ∈ C([0, δ]) satisfying ω(0) = 0. We may choose ω to be nondecreasing in [0, δ]. In the above inequality, we want to replace the term ω( | h | ) | h | by a C 1 function ψ(h) having the property: ψ(h) = o( | h | ). Following [23], we define the function \(\gamma \,:\, [0,\delta /2] \rightarrow \mathbb{R}\) by

$$\displaystyle{\gamma (r) =\int _{ 0}^{2r}\omega (t)\,\mathrm{d}t.}$$

Noting that

$$\displaystyle{\gamma (r) \geq \int _{r}^{2r}\omega (t)\,\mathrm{d}t \geq \omega (r)r\ \ \ \text{ for }r \in [0,\,\delta /2],}$$

we see that

$$\displaystyle{u(y + h) \leq u(y) + p \cdot h +\gamma (\vert h\vert )\ \ \mathrm{if}\ y + h \in \Omega \text{ and }\vert h\vert \leq \delta /2.}$$

It immediate to see that γ ∈ C 1([0, δ ∕ 2]) and \(\gamma (0) =\gamma^{\prime}(0) = 0\). We set ψ(h) = γ( | h | ) for h ∈ B δ ∕ 2(0). Then \(\psi \in {C}^{1}(B_{\delta /2}(0))\), ψ(0) = 0 and (0) = 0. It is now clear that if we set

$$\displaystyle{\phi (x) = u(y) + p \cdot (x - y) +\psi (x - y)\ \ \ \text{ for }x \in B_{\delta /2}(y),}$$

then the function u − ϕ attains a maximum over Ω ∩ B δ ∕ 2(y) at y and (y) = p. □ 

Now, we discuss a couple of stability results concerning viscosity solutions.

Proposition 1.9.

Let \(\{u_{\varepsilon }\}_{\varepsilon \in (0,1)} \subset {\mathcal{S}}^{-}\) . Assume that Ω is locally compact and \(\{u_{\varepsilon }\}\) converges locally uniformly to a function u in Ω as \(\varepsilon \rightarrow 0\) . Then \(u \in {\mathcal{S}}^{-}\).

Proof.

Let ϕ ∈ C 1(Ω). Assume that u  ∗  − ϕ attains a strict maximum at \(\hat{x} \in \Omega \). We choose a constant r > 0 so that \(K := \overline{B}_{r}(\hat{x}) \cap \Omega \) is compact. For each \(\varepsilon \in (0,\,1)\), we choose a maximum point (over K) \(x_{\varepsilon }\) of \(u_{\varepsilon }^{{\ast}}-\phi\).

Next, we choose a sequence \(\{\varepsilon _{j}\} \subset (0,\,1)\) converging to zero such that \(x_{\varepsilon _{j}} \rightarrow z\) for some z ∈ K as j → . Next, observe in view of the choice of \(x_{\varepsilon }\) that

$$\displaystyle\begin{array}{rcl} ({u}^{{\ast}}-\phi )(x_{\varepsilon _{ j}})& \geq & \,(u_{\varepsilon _{j}}^{{\ast}}-\phi )(x_{\varepsilon _{ j}}) -\| {u}^{{\ast}}- u_{\varepsilon _{ j}}^{{\ast}}\|_{ \infty,K} {}\\ & \geq & \,({u}^{{\ast}}-\phi )(x_{\varepsilon _{ j}}) - 2\|{u}^{{\ast}}- u_{\varepsilon _{ j}}^{{\ast}}\|_{ \infty,K} {}\\ & \geq & \,({u}^{{\ast}}-\phi )(\hat{x}) - 2\|{u}^{{\ast}}- u_{\varepsilon _{ j}}^{{\ast}}\|_{ \infty,K}. {}\\ \end{array}$$

Sending j →  yields

$$\displaystyle{({u}^{{\ast}}-\phi )(z) \geq \limsup (u_{\varepsilon _{ j}}^{{\ast}}-\phi )(x_{\varepsilon _{ j}}) \geq \liminf _{j\rightarrow \infty }(u_{\varepsilon _{j}}^{{\ast}}-\phi )(x_{\varepsilon _{ j}}) \geq ({u}^{{\ast}}-\phi )(\hat{x}),}$$

which shows that \(z =\hat{ x}\) and \(\lim _{j\rightarrow \infty }u_{\varepsilon _{j}}^{{\ast}}(x_{\varepsilon _{j}}) = {u}^{{\ast}}(\hat{x})\). For \(j \in \mathbb{N}\) sufficiently large, we have \(x_{\varepsilon _{j}} \in B_{r}(\hat{x})\) and, since \(u_{\varepsilon _{j}} \in {\mathcal{S}}^{-}\),

$$\displaystyle{F_{{\ast}}(x_{\varepsilon _{j}},u_{\varepsilon _{j}}^{{\ast}}(x_{\varepsilon _{ j}}),D\phi (x_{\varepsilon _{j}})) \leq 0.}$$

If we send j → , we find that \(u \in {\mathcal{S}}^{-}\). □ 

Proposition 1.10.

Let Ω be locally compact. Let \(\mathcal{F}\subset {\mathcal{S}}^{-}\) . That is, \(\mathcal{F}\) is a family of viscosity subsolutions of (FE) . Assume that \(\sup \mathcal{F}\) is locally bounded in Ω. Then we have \(\sup \mathcal{F}\in {\mathcal{S}}^{-}\).

Remark 1.6.

By definition, the set Ω is locally compact if for any x ∈ Ω, there exists a constant r > 0 such that \(\Omega \cap \overline{B}_{r}(x)\) is compact. For instance, every open subset and closed subset of \({\mathbb{R}}^{n}\) are locally compact. The set \(A := (0,\,1) \times [0,\,1] \subset {\mathbb{R}}^{2}\) is locally compact, but the set A ∪{ (0, 0)} is not locally compact.

Remark 1.7.

Similarly to Remark 1.5, if Ω is locally compact, then the C 1 regularity of the test functions in the Definition 1.2 can be replaced by the C regularity.

Proof.

Set \(u =\sup \mathcal{F}\). Let ϕ ∈ C 1(Ω) and \(\hat{x} \in \Omega \), and assume that

$$\displaystyle{\max _{\Omega }({u}^{{\ast}}-\phi ) = ({u}^{{\ast}}-\phi )(\hat{x}) = 0.}$$

We assume moreover that \(\hat{x}\) is a strict maximum point of u  ∗  − ϕ. That is, we have (u  ∗  − ϕ)(x) < 0 for all \(x\not =\hat{x}\). Choose a constant r > 0 so that \(W := \Omega \cap \overline{B}_{r}(\hat{x})\) is compact.

By the definition of u  ∗ , there are sequences \(\{y_{k}\} \subset W\) and \(\{v_{k}\} \subset \mathcal{F}\) such that

$$\displaystyle{y_{k} \rightarrow \hat{ x},\qquad v_{k}(y_{k}) \rightarrow {u}^{{\ast}}(\hat{x})\ \ \text{ as }\ k \rightarrow \infty.}$$

Since W is compact, for each \(k \in \mathbb{N}\) we may choose a point x k  ∈ W such that

$$\displaystyle{\max _{W}(v_{k}^{{\ast}}-\phi ) = (v_{ k}^{{\ast}}-\phi )(x_{ k}).}$$

By passing to a subsequence if necessary, we may assume that \(\{x_{k}\}\) converges to a point z ∈ W. We then have

$$\displaystyle\begin{array}{rcl} 0& =& \,({u}^{{\ast}}-\phi )(\hat{x}) \geq ({u}^{{\ast}}-\phi )(x_{ k}) \geq (v_{k}^{{\ast}}-\phi )(x_{ k}) {}\\ & \geq & \,(v_{k}^{{\ast}}-\phi )(y_{ k}) \geq (v_{k}-\phi )(y_{k}) \rightarrow ({u}^{{\ast}}-\phi )(\hat{x}) = 0, {}\\ \end{array}$$

and consequently

$$\displaystyle{\lim _{k\rightarrow \infty }{u}^{{\ast}}(x_{ k}) =\lim _{k\rightarrow \infty }v_{k}^{{\ast}}(x_{ k}) = {u}^{{\ast}}(\hat{x}).}$$

In particular, we see that

$$\displaystyle{({u}^{{\ast}}-\phi )(z) \geq \lim _{ k\rightarrow \infty }({u}^{{\ast}}-\phi )(x_{ k}) = 0,}$$

which shows that \(z =\hat{ x}\). That is, \(\lim _{k\rightarrow \infty }x_{k} =\hat{ x}\).

Thus, we have \(x_{k} \in B_{r}(\hat{x})\) for sufficiently large \(k \in \mathbb{N}\). Since \(v_{k} \in {\mathcal{S}}^{-}\), we get

$$\displaystyle{F_{{\ast}}(x_{k},v_{k}^{{\ast}}(x_{ k}),D\phi (x_{k})) \leq 0}$$

if k is large enough. Hence, sending k →  yields

$$\displaystyle{F_{{\ast}}(\hat{x},{u}^{{\ast}}(\hat{x}),D\phi (\hat{x})) \leq 0,}$$

which proves that \(u \in {\mathcal{S}}^{-}\). □ 

Theorem 1.3.

Let Ω be a locally compact subset of \(\,{\mathbb{R}}^{n}\) . Let \(\,\{u_{\varepsilon }\}_{\varepsilon \in (0,1)}\,\) and \(\{F_{\varepsilon }\}_{\varepsilon \in (0,1)}\) be locally uniformly bounded collections of functions on Ω and \(\Omega \times \mathbb{R} \times {\mathbb{R}}^{n}\) , respectively. Assume that for each \(\varepsilon \in (0,\,1)\) ,\(u_{\varepsilon }\) is a viscosity subsolution of

$$\displaystyle{F_{\varepsilon }(x,u_{\varepsilon }(x),\mathit{Du}_{\varepsilon }(x)) \leq 0\ \ \ \mathrm{in}\ \ \Omega.}$$

Set

$$\displaystyle\begin{array}{rcl} \bar{u}(x)& =& \,\lim _{r\rightarrow 0+}\sup \{u_{\varepsilon }(y)\, :\, y \in B_{r}(x) \cap \Omega,\,\varepsilon \in (0,\,r)\}, {}\\ \underline{F}(\xi )& =& \,\lim _{r\rightarrow 0+}\inf \{F_{\varepsilon }(\eta )\, :\,\eta \in \Omega \times \mathbb{R} \times {\mathbb{R}}^{n},\,\vert \eta -\xi \vert < r,\,\varepsilon \in (0,\,r)\}. {}\\ \end{array}$$

Then \(\bar{u}\) is a viscosity subsolution of

$$\displaystyle{\underline{F}(x,\bar{u}(x),D\bar{u}(x)) \leq 0\ \ \ \mathrm{in}\ \ \Omega.}$$

Remark 1.8.

The function \(\bar{u}\) is upper semicontinuous in Ω. Indeed, we have

$$\displaystyle{\bar{u}(y) \leq \sup \{ u_{\varepsilon }(z)\, :\, z \in B_{r}(x) \cap \Omega,\,\varepsilon \in (0,\,r)\}}$$

for all x ∈ Ω and y ∈ B r (x) ∩ Ω. This yields

$$\displaystyle{\limsup _{\Omega \ni y\rightarrow x}\bar{u}(y) \leq \sup \{ u_{\varepsilon }(z)\, :\, z \in B_{r}(x) \cap \Omega,\,\varepsilon \in (0,\,r)\}}$$

for all x ∈ Ω. Hence,

$$\displaystyle{\limsup _{\Omega \ni y\rightarrow x}\bar{u}(y) \leq \bar{ u}(x)\ \ \ \text{ for all }x \in \Omega.}$$

Similarly, the function \(\underline{F}\) is lower semicontinuous in \(\Omega \times \mathbb{R} \times {\mathbb{R}}^{n}\).

Proof.

It is easily seen that for all x ∈ Ω, r > 0 and y ∈ B r (x) ∩ Ω,

$$\displaystyle{u_{\varepsilon }^{{\ast}}(y) \leq \sup \{ u_{\varepsilon }(z)\, :\, z \in B_{ r}(x) \cap \Omega \}.}$$

From this we deduce that

$$\displaystyle{\bar{u}(x) =\lim _{r\rightarrow 0+}\sup \{u_{\varepsilon }^{{\ast}}(y)\, :\, y \in B_{ r}(x) \cap \Omega,\,0 <\varepsilon < r\}\ \ \ \text{ for all }\ x \in \Omega.}$$

Hence, we may assume by replacing \(u_{\varepsilon }\) by \(u_{\varepsilon }^{{\ast}}\) if necessary that \(u_{\varepsilon } \in \mathrm{USC}(\Omega )\). Similarly, we may assume that \(F_{\varepsilon } \in \mathrm{LSC}(\Omega \times \mathbb{R} \times {\mathbb{R}}^{n})\).

Let ϕ ∈ C 1(Ω) and \(\hat{x} \in \Omega \). Assume that \(\bar{u}-\phi\) has a strict maximum at \(\hat{x}\). Let r > 0 be a constant such that \(\overline{B}_{r}(\hat{x}) \cap \Omega \) is compact.

For each \(k \in \mathbb{N}\) we choose \(y_{k} \in B_{r/k}(\hat{x}) \cap \Omega \) and \(\varepsilon _{k} \in (0,\,1/k)\) so that

$$\displaystyle{\vert \bar{u}(\hat{x}) - u_{\varepsilon _{k}}(y_{k})\vert < 1/k,}$$

and then choose a maximum point \(x_{k} \in B_{r}(\hat{x}) \cap \Omega \) of \(u_{\varepsilon _{k}}-\phi\) over \(\overline{B}_{r}(\hat{x}) \cap \Omega \).

Since

$$\displaystyle{(u_{\varepsilon _{k}}-\phi )(x_{k}) \geq (u_{\varepsilon _{k}}-\phi )(y_{k}),}$$

we get

$$\displaystyle{\limsup _{k\rightarrow 0}(u_{\varepsilon _{k}}-\phi )(x_{k}) \geq (\bar{u}-\phi )(\hat{x}),}$$

which implies that

$$\displaystyle{\lim _{k\rightarrow \infty }x_{k} =\hat{ x}\ \ \text{ and }\ \ \lim _{k\rightarrow \infty }u_{\varepsilon _{k}}(x_{k}) =\bar{ u}(\hat{x}).}$$

If \(k \in \mathbb{N}\) is sufficiently large, we have \(x_{k} \in B_{r}(\hat{x}) \cap \Omega \) and hence

$$\displaystyle{F_{\varepsilon _{k}}(x_{k},u_{\varepsilon _{k}}(x_{k}),D\phi (x_{k})) \leq 0.}$$

Thus, we get

$$\displaystyle{\underline{F}(\hat{x},\bar{u}(\hat{x}),D\phi (\hat{x})) \leq 0.}$$

 □ 

Proposition 1.9 can be seen now as a direct consequence of the above theorem. The following proposition is a consequence of the above theorem as well.

Proposition 1.11.

Let Ω be locally compact. Let \(\{u_{k}\}\) be a sequence of viscosity subsolutions of (FE).Assume that \(\{u_{k}\} \subset \mathrm{USC}(\Omega )\) and that \(\{u_{k}\}\) is a nonincreasing sequence of functions on Ω, i.e., u k (x) ≥ u k+1 (x) for all x ∈ Ω and \(k \in \mathbb{N}\) . Set

$$\displaystyle{u(x) =\lim _{k\rightarrow \infty }u_{k}(x)\ \ \ \text{ for }\ x \in \Omega.}$$

Assume that u is locally bounded on Ω. Then \(u \in {\mathcal{S}}^{-}\).

Let us introduce the (outer) normal cone N(z, Ω) at z ∈ Ω by

$$\displaystyle{N(z,\Omega ) =\{ p \in {\mathbb{R}}^{n}\, :\, 0 \geq p \cdot (x - z) + o(\vert x - z\vert )\ \text{ as }\ \Omega \ni x \rightarrow z\}.}$$

Another definition equivalent to the above is the following:

$$\displaystyle{N(z,\Omega ) = -{D}^{+}\mathbf{1}_{ \Omega }(z),}$$

where 1 Ω denotes the characteristic function of Ω. Note that if z ∈ Ω is an interior point of Ω, then N(z, Ω) = { 0}.

We say that (FE) or the pair (F, Ω) is proper if F(x, r, p + q) ≥ F(x, r, p) for all \((x,r,p) \in \Omega \times \mathbb{R} \times {\mathbb{R}}^{n}\) and all q ∈ N(x, Ω).

Proposition 1.12.

Assume that (FE)is proper. If u ∈ C 1 (Ω) is a classical subsolution of (FE),then \(u \in {\mathcal{S}}^{-}\).

Proof.

Let ϕ ∈ C 1(Ω) and assume that u − ϕ attains a maximum at z ∈ Ω. We may assume by extending the domain of definition of u and ϕ that u and ϕ are defined and of class C 1 in B r (z) for some r > 0. By reselecting r > 0 small enough if needed, we may assume that

$$\displaystyle{(u-\phi )(x) < (u-\phi )(z) + 1\ \ \ \text{ for all }x \in B_{r}(z).}$$

It is clear that the function \(u -\phi +\mathbf{1}_{\Omega }\) attains a maximum over B r (z) at z, which shows that \(D\phi (z) -\mathit{Du}(z) \in {D}^{+}\mathbf{1}_{\Omega }(z)\). Setting \(q = -D\phi (z) + \mathit{Du}(z)\), we have \(\mathit{Du}(z) = D\phi (z) + q\) and

$$\displaystyle{0 \geq F(z,u(z),D\phi (z) + q) \geq F(z,u(z),D\phi (z)) \geq F_{{\ast}}(z,u(z),D\phi (z)),}$$

which completes the proof. □ 

Proposition 1.13 (Perron method). 

Let \(\mathcal{F}\) be a nonempty subset of \({\mathcal{S}}^{-}\) having the properties:

  1. (P1)

    \(\sup \mathcal{F}\in \mathcal{F}\).

  2. (P2)

    If \(v \in \mathcal{F}\) and \(v\not\in {\mathcal{S}}^{+}\) , then there exists a \(w \in \mathcal{F}\) such that w(y) > v(y) at some point y ∈ Ω.

Then \(\sup \mathcal{F}\in \mathcal{S}\).

Proof.

We have \(\sup \mathcal{F}\in \mathcal{F}\subset {\mathcal{S}}^{-}\). That is, \(\sup \mathcal{F}\in {\mathcal{S}}^{-}\). If we suppose that \(\sup \mathcal{F}\not\in {\mathcal{S}}^{+}\), then, by (P2), we have \(w \in \mathcal{F}\) such that \(w(y) > (\sup \mathcal{F})(y)\) for some y ∈ Ω, which contradicts the definition of \(\sup \mathcal{F}\). Hence, \(\sup \mathcal{F}\in {\mathcal{S}}^{+}\). □ 

Theorem 1.4.

Assume that Ω is locally compact and that (FE) is proper. Let \(f \in \mathrm{LSC}(\Omega ) \cap {\mathcal{S}}^{-}\) and \(g \in \mathrm{USC}(\Omega ) \cap {\mathcal{S}}^{+}\) . Assume that f ≤ g in Ω. Set

$$\displaystyle{\mathcal{F} =\{ v \in {\mathcal{S}}^{-}\, :\, f \leq v \leq g\mathrm{in}\ \Omega \}.}$$

Then \(\sup \mathcal{F}\in \mathcal{S}\).

In the above theorem, the semicontinuity requirement on f, g is “opposite” in a sense: the lower (resp., upper) semicontinuity for the subsolution f (resp., supersolution g). This choice of semicontinuities is convenient in practice since in the construction of supersolution f, for instance, one often takes the infimum of a collection of continuous supersolutions and the resulting function is automatically upper semicontinuous.

Of course, under the same hypotheses of the above theorem, we have following conclusion as well: if we set \({\mathcal{F}}^{+} =\{ v \in {\mathcal{S}}^{+}\, :\, f \leq v \leq g\mathrm{in}\ \Omega \}\), then \(\inf {\mathcal{F}}^{+} \in \mathcal{S}\).

Lemma 1.3.

Assume that Ω is locally compact and that (FE) is proper. Let \(u \in {\mathcal{S}}^{-}\) and y ∈ Ω, and assume that u is not a viscosity supersolution of (FE) at y, that is,

$$\displaystyle{{F}^{{\ast}}(y,u_{ {\ast}}(y),p) < 0\ \ \ \text{ for some }\ p \in {D}^{-}u_{ {\ast}}(y).}$$

Let \(\varepsilon > 0\) and U be a neighborhood of y. Then there exists a \(v \in {\mathcal{S}}^{-}\) such that

$$\displaystyle{ \left \{\begin{array}{llll} &u(x) \leq v(x) \leq \max \{ u(x),u_{{\ast}}(y)+\varepsilon \}&&\text{ for all }\ x \in \Omega, \\ &v = u &&\mathrm{in}\ \ \Omega \setminus U, \\ &v_{{\ast}}(y) > u_{{\ast}}(y). \end{array} \right. }$$
(22)

Furthermore, if u is continuous at y, then there exist an open neighborhood V of y and a constant δ > 0 such that v is a viscosity subsolution of

$$\displaystyle{ F(x,v(x),Dv(x)) = -\delta \ \ \ \mathrm{in}\ V \cap \Omega. }$$
(23)

Proof.

By assumption, there exists a function ϕ ∈ C 1(Ω) such that u  ∗ (y) = ϕ(y), u  ∗ (x) > ϕ(x) for all x ≠ y and

$$\displaystyle{ {F}^{{\ast}}(y,u_{ {\ast}}(y),D\phi (y)) < 0. }$$

Thanks to the upper semicontinuity of F  ∗ , there exists a \(\delta \in (0,\,\varepsilon )\) such that

$$\displaystyle{ {F}^{{\ast}}(x,\phi (x) + t,D\phi (x)) < -\delta \ \ \ \text{ for all }\ (x,t) \in (\overline{B}_{\delta }(y) \cap \Omega ) \times [0,\,\delta ], }$$
(24)

and \(\overline{B}_{\delta }(y) \cap \Omega \) is a compact subset of U.

By replacing δ > 0 by a smaller number if needed, we may assume that

$$\displaystyle{ \phi (x)+\delta \leq u_{{\ast}}(y) +\varepsilon \ \ \ \text{ for all }\ x \in \overline{B}_{\delta }(y) \cap \Omega. }$$
(25)

Since u  ∗  − ϕ attains a strict minimum at y and the minimum value is zero, if \((\Omega \cap \overline{B}_{\delta }(y)) \setminus B_{\delta /2}(y)\not =\varnothing \), then the constant

$$\displaystyle{m :=\min _{(\Omega \cap \overline{B}_{\delta }(y))\setminus B_{\delta /2}(y)}(u_{{\ast}}-\phi )}$$

is positive. Of course, in this case, we have

$$\displaystyle{u_{{\ast}}(x) \geq \phi (x) + m\ \ \text{ for all }x \in (\Omega \cap \overline{B}_{\delta }(y)) \setminus B_{\delta /2}(y).}$$

Set λ = min{m, δ} if \((\Omega \cap \overline{B}_{\delta }(y)) \setminus B_{\delta /2}(y)\not =\varnothing \) and λ = δ otherwise, and observe that

$$\displaystyle{ u_{{\ast}}(x) \geq \phi (x) +\lambda \ \ \text{ for all }x \in (\Omega \cap \overline{B}_{\delta }(y)) \setminus B_{\delta /2}(y). }$$
(26)

We define \(v\, :\, \Omega \rightarrow \mathbb{R}\) by

$$\displaystyle{v(x) = \left \{\begin{array}{@{}l@{\quad }l@{}} \max \{u(x),\phi (x)+\lambda \}\quad &\mathrm{if}\ x \in B_{\delta }(y),\\ u(x) \quad &\mathrm{if } \ x\not\in B_{\delta }(y). \end{array} \right.}$$

If we set \(\psi (x) =\phi (x)+\lambda\) for x ∈ B δ (y) ∩ Ω, by (24), ψ is a classical subsolution of (FE) in B δ (y) ∩ Ω. Since (FE) is proper, ψ is a viscosity subsolution of (FE) in B δ (y) ∩ Ω. Hence, by Proposition 1.10, we see that v is a viscosity subsolution of (FE) in B δ (y) ∩ Ω.

According to (26) and the definition of v, we have

$$\displaystyle{v(x) = u(x)\ \ \text{ for all }x \in \Omega \setminus B_{\delta /2}(y),}$$

and, hence, v is a viscosity subsolution of (FE) in \(\Omega \setminus \overline{B}_{\delta /2}(y)\) Thus, we find that \(v \in {\mathcal{S}}^{-}\).

Since v = u in Ω ∖ B δ (y) by the definition of v, it follows that v = u in Ω ∖ U. It is clear by the definition of v that v ≥ u in Ω. Moreover, by (25) we get

$$\displaystyle{v(x) \leq \max \{ u(x),\,u_{{\ast}}(y) +\varepsilon \}\ \ \ \text{ for all }x \in \Omega \cap B_{\delta }(y).}$$

Also, observe that

$$\displaystyle{v_{{\ast}}(y) =\max \{ u_{{\ast}}(y),\,u_{{\ast}}(y)+\lambda \} = u_{{\ast}}(y)+\lambda > u_{{\ast}}(y).}$$

Thus, (22) is valid.

Now, we assume that u is continuous at y. Then we find an open neighborhood V ⊂ B δ (y) of y such that

$$\displaystyle{u(x) <\phi (x) +\lambda \ \ \ \text{ for all }\ x \in V \cap \Omega,}$$

and hence, we have \(v(x) =\phi (x)+\lambda\) for all x ∈ V ∩ Ω. Now, by (24) we see that v is a classical (and hence viscosity) subsolution of (23). □ 

Proof (Theorem 1.4). 

We have \(\mathcal{F}\not =\varnothing \) since \(f \in \mathcal{F}\). In view of Proposition 1.13, we need only to show that the set \(\mathcal{F}\) satisfies (P1) and (P2).

By Proposition 1.10, we see immediately that \(\mathcal{F}\) satisfies (P1).

To check property (P2), let \(v \in \mathcal{F}\) be not a viscosity supersolution of (FE). There is a point y ∈ Ω where v is not a viscosity supersolution of (FE). That is, for some \(p \in {D}^{-}v_{{\ast}}(y)\), we have

$$\displaystyle{ {F}^{{\ast}}(y,v_{ {\ast}}(y),p) < 0. }$$
(27)

Noting v  ∗  ≤ g  ∗  in Ω, there are two possibilities: v  ∗ (y) = g  ∗ (y) or \(v_{{\ast}}(y) < g_{{\ast}}(y)\). If \(v_{{\ast}}(y) = g_{{\ast}}(y)\), then \(p \in {D}^{-}g_{{\ast}}(y)\). Since \(g \in {\mathcal{S}}^{+}\), we have

$$\displaystyle{{F}^{{\ast}}(y,g_{ {\ast}}(y),p) \geq 0,}$$

which contradicts (27). If \(v_{{\ast}}(y) < g_{{\ast}}(y)\), then we choose a constant \(\varepsilon > 0\) and a neighborhood V of y so that

$$\displaystyle{ v_{{\ast}}(y)+\varepsilon < g_{{\ast}}(x)\ \ \ \text{ for all }x \in V \cap \Omega. }$$
(28)

Now, Lemma 1.3 guarantees that there exist  \(w \in {\mathcal{S}}^{-}\)  such that  \(v \leq w \leq \max \{ v,v_{{\ast}}(y)+\varepsilon \}\) in Ω, v = w in Ω ∖ V and \(w_{{\ast}}(y) > v_{{\ast}}(y)\). For any x ∈ Ω ∩ V, by (28) we have

$$\displaystyle{w(x) \leq \max \{ v(x),g_{{\ast}}(x)\} \leq g(x).}$$

For any x ∈ Ω ∖ V, we have

$$\displaystyle{w(x) = v(x) \leq g(x).}$$

Thus, we find that \(w \in \mathcal{F}\). Since \(w_{{\ast}}(y) > v_{{\ast}}(y)\), it is clear that w(z) > v(z) at some point z ∈ Ω. Hence, \(\mathcal{F}\) satisfies (P2). □ 

1.5 An Example

We illustrate the use of the stability properties established in the previous subsection by studying the solvability of the Dirichlet problem for the eikonal equation

$$\displaystyle{ \vert \mathit{Du}(x)\vert = k(x)\mathrm{in}\ \ \Omega, }$$
(29)
$$\displaystyle{ u(x) = 0\mathrm{on}\ \ \partial \Omega, }$$
(30)

where Ω is a bounded, open, connected subset of \({\mathbb{R}}^{n}\) and \(k \in C(\overline{\Omega })\) is a positive function in \(\overline{\Omega }\), i.e., k(x) > 0 for all \(x \in \overline{\Omega }\).

Note that the constant function f(x) : = 0 is a classical subsolution of (29). Set \(M =\max _{\overline{\Omega }}k\). We observe that for each y ∈ ∂Ω the function \(g_{y}(x) := M\vert x - y\vert \) is a classical supersolution of (29). We set

$$\displaystyle{g(x) =\inf \{ g_{y}(x)\, :\, y \in \partial \Omega \}\ \ \ \text{ for }\ x \in \overline{\Omega }.}$$

By Proposition 1.10 (its version for supersolutions), we see that g is a viscosity supersolution of (29). Also, by applying Lemma 1.1, we find that g is Lipschitz continuous in \(\overline{\Omega }\).

An application of Theorem 1.4 ensures that there is a viscosity solution \(u\, :\, \Omega \rightarrow \mathbb{R}\) of (29) such that f ≤ u ≤ g in Ω. Since \(f(x) = g(x) = 0\) on ∂Ω, if we set u(x) = 0 for x ∈ ∂Ω, then the resulting function u is continuous at points on the boundary ∂Ω and satisfies the Dirichlet condition (30) in the classical sense.

Note that u  ∗  ≤ g in Ω, which clearly implies that u = u  ∗  ∈ USC(Ω). Now, if we use the next proposition, we find that u is locally Lipschitz continuous in Ω and conclude that \(u \in C(\overline{\Omega })\). Thus, the Dirichlet problem (29)–(30) has a viscosity solution \(u \in C(\overline{\Omega })\) which satisfies (30) in the classical (or pointwise) sense.

Proposition 1.14.

Let R > 0, C > 0 and u ∈USC (B R ). Assume that u is a viscosity solution of

$$\displaystyle{\vert \mathit{Du}(x)\vert \leq C\ \ \ \mathrm{in}\ \ B_{R}.}$$

Then u is Lipschitz continuous in B R with C being a Lipschitz bound. That is, \(\vert u(x) - u(y)\vert \leq C\vert x - y\vert \) for all x,y ∈ B R .

Proof.

Fix any z ∈ B R and set \(r = (R -\vert z\vert )/4\). Fix any y ∈ B r (z). Note that B 3r (y) ⊂ B R . Choose a function f ∈ C 1([0, 3r)) so that f(t) = t for all t ∈ [0, 2r], f′(t) ≥ 1 for all t ∈ [0, 3r) and \(\lim _{t\rightarrow 3r-}f(t) = \infty \). Fix any \(\varepsilon > 0\), and we claim that

$$\displaystyle{ u(x) \leq v(x) := u(y) + (C+\varepsilon )f(\vert x - y\vert )\ \ \ \text{ for all }\ x \in B_{3r}(y). }$$
(31)

Indeed, if this were not the case, we would find a point ξ ∈ B 3r (y) ∖ { y} such that u − v attains a maximum at ξ, which yields together with the viscosity property of u

$$\displaystyle{C \geq \vert Dv(\xi )\vert = (C+\varepsilon )f^{\prime}(\vert \xi - y\vert ) \geq C +\varepsilon.}$$

This is a contradiction. Thus we have (31).

Note that if x ∈ B r (z), then x ∈ B 2r (y) and \(f(\vert x - y\vert ) = \vert x - y\vert \). Hence, from (31), we get

$$\displaystyle{u(x) - u(y) \leq (C+\varepsilon )\vert x - y\vert \ \ \ \text{ for all }x,y \in B_{r}(z).}$$

By symmetry, we see that

$$\displaystyle{\vert u(x) - u(y)\vert \leq (C+\varepsilon )\vert x - y\vert \ \ \ \text{ for all }x,y \in B_{r}(z),}$$

from which we deduce that

$$\displaystyle{ \vert u(x) - u(y)\vert \leq C\vert x - y\vert \ \ \ \text{ for all }x,y \in B_{r}(z), }$$
(32)

Now, let x, y ∈ B R be arbitrary points. Set \(r = \frac{1} {4}\min \{R -\vert x\vert,\,R -\vert y\vert \}\), and choose a finite sequence \(\{z_{i}\}_{i=0}^{N}\) of points on the line segment [x, y] so that z 0 = x, z N  = y, \(\vert z_{i} - z_{i-1}\vert < r\) for all i = 1, , N and \(\sum _{i=1}^{N}\vert z_{i} - z_{i-1}\vert = \vert x - y\vert \). By (32), we get

$$\displaystyle{\vert u(z_{i}) - u(z_{i-1})\vert \leq C\vert z_{i} - z_{i-1}\vert \ \ \ \text{ for all }i = 1,\ldots,N.}$$

Summing these over i = 1, , N yields the desired inequality. □ 

1.6 Sup-convolutions

Sup-convolutions and inf-convolutions are basic and important tools for regularizing or analyzing viscosity solutions. In this subsection, we recall some properties of sup-convolutions.

Let \(u\, :\, {\mathbb{R}}^{n} \rightarrow \mathbb{R}\) be a bounded function and \(\varepsilon \in \mathbb{R}_{+}\). The standard sup-convolution \({u}^{\varepsilon }\, :\, {\mathbb{R}}^{n} \rightarrow \mathbb{R}\) and inf-convolution \(u_{\varepsilon }\, :\, {\mathbb{R}}^{n} \rightarrow \mathbb{R}\) are defined, respectively, by

$$\displaystyle{{u}^{\varepsilon }(x) =\sup _{y\in {\mathbb{R}}^{n}}\left (u(y) -\frac{1} {2\varepsilon }\vert y - x{\vert }^{2}\right )\ \ \text{ and }\ \ u_{\varepsilon }(x) =\inf _{ y\in {\mathbb{R}}^{n}}\left (u(y) + \frac{1} {2\varepsilon }\vert y - x{\vert }^{2}\right ).}$$

Note that

$$\displaystyle{u_{\varepsilon }(x) = -\sup \left (-u(y) -\frac{1} {2\varepsilon }\vert y - x{\vert }^{2}\right ) = -{(-u)}^{\varepsilon }(x).}$$

This relation immediately allows us to interpret a property of sup-convolutions into the corresponding property of inf-convolutions.

In what follows we assume that u is bounded and upper semicontinuous in \({\mathbb{R}}^{n}\). Let M > 0 be a constant such that | u(x) | ≤ M for all \(x \in {\mathbb{R}}^{n}\).

Proposition 1.15.

  1. (i)

    We have

    $$\displaystyle{-M \leq u(x) \leq {u}^{\varepsilon }(x) \leq M\ \ \ \text{ for all }\ x \in {\mathbb{R}}^{n}.}$$
  2. (ii)

    Let \(x \in {\mathbb{R}}^{n}\) and \(p \in {D}^{+}{u}^{\varepsilon }(x)\) . Then

    $$\displaystyle{\vert p\vert \leq 2\sqrt{\frac{M} {\varepsilon }} \ \ \ \text{ and }\ \ \ p \in {D}^{+}u(x +\varepsilon p).}$$

Another important property of sup-convolutions is that the sup-convolution \({u}^{\varepsilon }\) is semiconvex in \({\mathbb{R}}^{n}\). More precisely, the function

$$\displaystyle{{u}^{\varepsilon }(x) + \frac{1} {2\varepsilon }\vert x{\vert }^{2} =\sup _{ y\in {\mathbb{R}}^{n}}\left (u(y) -\frac{1} {2\varepsilon }\vert y{\vert }^{2} + \frac{1} {\varepsilon } y \cdot x\right )}$$

is convex in \({\mathbb{R}}^{n}\) (see Appendix A.2) as is clear from the form of the right hand side of the above identity.

Proof.

To show assertion (i), we just check that for all \(x \in {\mathbb{R}}^{n}\),

$$\displaystyle{{u}^{\varepsilon }(x) \leq \sup _{y\in {\mathbb{R}}^{n}}u(y) \leq M,}$$

and

$$\displaystyle{{u}^{\varepsilon }(x) \geq u(x) \geq -M.}$$

Next, we prove assertion (ii). Let \(\hat{x} \in {\mathbb{R}}^{n}\) and \(\hat{p} \in {D}^{+}{u}^{\varepsilon }(\hat{x})\). Choose a point \(\hat{y} \in {\mathbb{R}}^{n}\) so that

$$\displaystyle{{u}^{\varepsilon }(\hat{x}) = u(\hat{y}) -\frac{1} {2\varepsilon }\vert \hat{y} -\hat{ x}{\vert }^{2}.}$$

(Such a point \(\hat{y}\) always exists under our assumptions on u.) It is immediate to see that

$$\displaystyle{ \frac{1} {2\varepsilon }\vert \hat{y} -\hat{ x}{\vert }^{2} = u(\hat{y}) - {u}^{\varepsilon }(\hat{x}) \leq 2M. }$$
(33)

We may choose a function \(\phi \in {C}^{1}({\mathbb{R}}^{n})\) so that \(D\phi (\hat{x}) =\hat{ p}\) and \(\max _{{\mathbb{R}}^{n}}({u}^{\varepsilon }-\phi ) = ({u}^{\varepsilon }-\phi )(\hat{x})\). Observe that the function

$$\displaystyle{{\mathbb{R}}^{2n} \ni (x,y)\mapsto u(y) -\frac{1} {2\varepsilon }\vert y - x{\vert }^{2} -\phi (x)}$$

attains a maximum at \((\hat{x},\hat{y})\). Hence, both the functions

$$\displaystyle{{\mathbb{R}}^{n} \ni x\mapsto -\frac{1} {2\varepsilon }\vert \hat{y} - x{\vert }^{2} -\phi (x)}$$

and

$$\displaystyle{{\mathbb{R}}^{n} \ni x\mapsto u(x +\hat{ y} -\hat{ x}) -\phi (x)}$$

attain maximum values at \(\hat{x}\). Therefore, we find that

$$\displaystyle{\frac{1} {\varepsilon } (\hat{x} -\hat{ y}) + D\phi (\hat{x}) = 0\ \ \text{ and }\ \ D\phi (\hat{x}) \in {D}^{+}u(\hat{y}),}$$

which shows that

$$\displaystyle{\hat{p} = \frac{1} {\varepsilon } (\hat{y} -\hat{ x}) \in {D}^{+}u(\hat{y}).}$$

From this, we get  \(\hat{y} =\hat{ x} +\varepsilon \hat{ p}\), and, moreover,  \(\hat{p} \in {D}^{+}u(\hat{x} +\varepsilon \hat{ p})\). Also, using (33), we get  \(\vert \hat{p}\vert \leq 2\sqrt{M/\varepsilon }\). Thus we see that (ii) holds. □ 

The following observations illustrate a typical use of the above proposition. Let Ω be an open subset of \({\mathbb{R}}^{n}\). Let \(H\, :\, \Omega \times {\mathbb{R}}^{n} \rightarrow \mathbb{R}\) and \(u\, :\, \Omega \rightarrow \mathbb{R}\) be bounded and upper semicontinuous. Let M > 0 be a constant such that | u(x) | ≤ M for all x ∈ Ω. Let \(\varepsilon > 0\). Set \(\delta = 2\sqrt{\varepsilon M}\) and \(\Omega _{\delta } =\{ x \in \Omega \, :\, \mathrm{dist}(x,\partial \Omega ) >\delta \}\). Define \({u}^{\varepsilon }\) as above with u extended to \({\mathbb{R}}^{n}\) by setting \(u(x) = -M\) for \(x \in {\mathbb{R}}^{n} \setminus \Omega \). (Or, in a slightly different and more standard way, one may define \({u}^{\varepsilon }\) by

$$\displaystyle{{u}^{\varepsilon }(x) =\sup _{y\in \Omega }\left (u(y) -\frac{1} {2\varepsilon }\vert x - y{\vert }^{2}\right ).\ )}$$

By applying Proposition 1.15, we deduce that if u is a viscosity subsolution of

$$\displaystyle{H(x,\mathit{Du}(x)) \leq 0\ \ \ \mathrm{in}\ \ \Omega,}$$

then \({u}^{\varepsilon }\) is a viscosity subsolution of both

$$\displaystyle{ H(x +\varepsilon { \mathit{Du}}^{\varepsilon }(x),\,{\mathit{Du}}^{\varepsilon }(x)) \leq 0\ \ \ \mathrm{in}\ \ \Omega _{\delta }, }$$
(34)

and

$$\displaystyle{ \vert {\mathit{Du}}^{\varepsilon }(x)\vert \leq 2\sqrt{\frac{M} {\varepsilon }} \ \ \ \mathrm{in}\ \ \Omega _{\delta }. }$$
(35)

If we set

$$\displaystyle{G(x,p) =\inf _{y\in B_{\delta }}H(x + y,p)\ \ \ \text{ for }\ x \in \Omega _{\delta },}$$

then (34) implies that \({u}^{\varepsilon }\) is a viscosity subsolution of

$$\displaystyle{G(x,{\mathit{Du}}^{\varepsilon }(x)) \leq 0\ \ \ \mathrm{in}\ \Omega _{\delta }.}$$

If we apply Proposition 1.14 to \({u}^{\varepsilon }\), we see from (35) that \({u}^{\varepsilon }\) is locally Lipschitz in Ω δ .

2 Neumann Boundary Value Problems

We assume throughout this section and the rest of this article that \(\Omega \subset {\mathbb{R}}^{n}\) is open.

We will be concerned with the initial value problem for the Hamilton–Jacobi equation of evolution type

$$\displaystyle{\frac{\partial u} {\partial t} (x,t) + H(x,D_{x}u(x,t)) = 0\ \ \ \mathrm{in}\ \Omega \times (0,\,\infty ),}$$

and the asymptotic behavior of its solutions u(x, t) as t → .

The stationary problem associated with the above Hamilton–Jacobi equation is stated as

$$\displaystyle{ \left \{\begin{array}{ll} &H(x,\mathit{Du}(x)) = 0\ \ \ \mathrm{in}\ \Omega,\\ &\text{boundary condition on }\partial \Omega. \end{array} \right. }$$
(36)

In this article we will be focused on the Neumann boundary value problem among other possible choices of boundary conditions like periodic, Dirichlet, state-constraints boundary conditions.

We are thus given two functions \(\gamma \in C(\partial \Omega, {\mathbb{R}}^{n})\) and \(g \in C(\partial \Omega, \mathbb{R})\) which satisfy

$$\displaystyle{ \nu (x) \cdot \gamma (x) > 0\ \ \ \text{ for all }x \in \partial \Omega, }$$
(37)

where ν(x) denotes the outer unit normal vector at x, and the boundary condition posed on the unknown function u is stated as

$$\displaystyle{\gamma (x) \cdot \mathit{Du}(x) = g(x)\ \ \ \text{ for }x \in \partial \Omega.}$$

This condition is called the (inhomogeneous, linear) Neumann boundary condition. We remark that if \(u \in {C}^{1}(\overline{\Omega })\), then the directional derivative ∂u ∕ ∂γ of u in the direction of γ is given by

$$\displaystyle{\frac{\partial u} {\partial \gamma } (x) =\gamma (x) \cdot \mathit{Du}(x) =\lim _{t\rightarrow 0}\frac{u(x + t\gamma (x)) - u(x)} {t} \ \ \ \text{ for }x \in \partial \Omega.}$$

(Note here that u is assumed to be defined in a neighborhood of x.)

Our boundary value problem (36) is now stated precisely as

$$\displaystyle{ \left \{\begin{array}{llll} &H(x,\mathit{Du}(x)) = 0\ \ \ &&\mathrm{in}\ \Omega, \\ &\frac{\partial u} {\partial \gamma } (x) = g(x)\ \ \ &&\mathrm{on}\ \partial \Omega.\end{array} \right. }$$
(SNP)

Let U be an open subset of \({\mathbb{R}}^{n}\) such that U ∩ Ω ≠ . At this stage we briefly explain the viscosity formulation of a more general boundary value problem

$$\displaystyle{ \left \{\begin{array}{ll} &F(x,u(x),\mathit{Du}(x)) = 0\ \ \ \mathrm{in}\ U \cap \Omega,\\ &B(x, u(x), \mathit{Du } (x)) = 0\ \ \ \mathrm{on } \ U \cap \partial \Omega, \end{array} \right. }$$
(38)

where the functions \(F\, :\, (U \cap \overline{\Omega }) \times \mathbb{R} \times {\mathbb{R}}^{n} \rightarrow \mathbb{R}\), \(B\, :\, (U \cap \partial \Omega ) \times \mathbb{R} \times {\mathbb{R}}^{n} \rightarrow \mathbb{R}\) and \(u\, :\, (U \cap \overline{\Omega }) \rightarrow \mathbb{R}\) are assumed to be locally bounded in their domains of definition. The function u is said to be a viscosity subsolution of (38) if the following requirements are fulfilled:

$$\displaystyle{\left \{\begin{array}{ll} &\phi \in {C}^{1}(\overline{\Omega }),\quad \hat{x} \in \overline{\Omega },\quad \max _{\overline{\Omega }}({u}^{{\ast}}-\phi ) = ({u}^{{\ast}}-\phi )(\hat{x})\\ &\Rightarrow \\ &\begin{array}{llllll} &\text{(i)} &&F_{{\ast}}(\hat{x},{u}^{{\ast}}(\hat{x}),D\phi (\hat{x})) \leq 0 &&\mathrm{if}\ \ \hat{x} \in U \cap \Omega, \\ &\text{(ii)}&&F_{{\ast}}(\hat{x},{u}^{{\ast}}(\hat{x}),D\phi (\hat{x})) \wedge B_{{\ast}}(\hat{x},{u}^{{\ast}}(\hat{x}),D\phi (\hat{x})) \leq 0&&\mathrm{if}\ \ \hat{x} \in U \cap \partial \Omega.\end{array}\end{array} \right.}$$

The upper and lower semicontinuous envelopes are taken in all the variables. That is, for x ∈ U ∩ Ω, \(\xi \in (U \cap \overline{\Omega }) \times \mathbb{R} \times {\mathbb{R}}^{n}\) and \(\eta \in (U \cap \partial \Omega ) \times \mathbb{R} \times {\mathbb{R}}^{n}\),

$$\displaystyle{\begin{array}{ll} {u}^{{\ast}}(x) =&\,\lim _{r\rightarrow 0+}\sup \{u(y)\, :\, y \in B_{r}(x) \cap U \cap \overline{\Omega })\}, \\ F_{{\ast}}(\xi ) = &\,\lim _{r\rightarrow 0+}\inf \{F(X)\, :\, X \in (U \cap \Omega ) \times \mathbb{R} \times {\mathbb{R}}^{n},\,\vert X -\xi \vert < r\}, \\ B_{{\ast}}(\eta ) = &\,\lim _{r\rightarrow 0+}\inf \{B(Y )\, :\, Y \in (U \cap \partial \Omega ) \times \mathbb{R} \times {\mathbb{R}}^{n},\,\vert Y -\eta \vert < r\}.\end{array} }$$

The definition of viscosity supersolutions of the boundary value problem (38) is given by reversing the upper and lower positions of ∗ , the inequalities, and “sup” and “inf” (including ∧ and ∨ ), respectively. Then viscosity solutions of (38) are defined as those functions which are both viscosity subsolution and supersolution of (38).

Here, regarding the above definition of boundary value problems, we point out the following: define the function \(G\, :\, (U \cap \overline{\Omega }) \times \mathbb{R} \times {\mathbb{R}}^{n} \rightarrow \mathbb{R}\) by

$$\displaystyle{ G(x,u,p) = \left \{\begin{array}{@{}l@{\quad }l@{}} F(x,u,p)\quad &\mathrm{if}\ x \in \Omega,\\ B(x, u, p)\quad &\mathrm{if } \ \ x \in \partial \Omega, \end{array} \right. }$$
(39)

and note that the lower (resp., upper) semicontinuous envelope G  ∗  (resp., G  ∗ ) of G is given by

$$\displaystyle\begin{array}{rcl} & G_{{\ast}}(x,u,p) = \left \{\begin{array}{@{}l@{\quad }l@{}} F_{{\ast}}(x,u,p) \quad &\mathrm{if}\ x \in \Omega, \\ F_{{\ast}}(x,u,p) \wedge B_{{\ast}}(x,u,p)\quad &\mathrm{if}\ x \in \partial \Omega \end{array} \right. & {}\\ \Bigg(\text{resp.,}\ \ & {G}^{{\ast}}(x,u,p) = \left \{\begin{array}{@{}l@{\quad }l@{}} {F}^{{\ast}}(x,u,p) \quad &\mathrm{if}\ x \in \Omega, \\ {F}^{{\ast}}(x,u,p) \vee {B}^{{\ast}}(x,u,p)\quad &\mathrm{if}\ x \in \partial \Omega \end{array} \right.\Bigg).& {}\\ \end{array}$$

Thus, the above definition of viscosity subsolutions, supersolutions and solutions of (38) is the same as that of Definition 1.2 with F and Ω replaced by G defined by (39) and \(U \cap \overline{\Omega }\), respectively. Therefore, the propositions in Sect. 1.4 are valid as well to viscosity subsolutions, supersolutions and solutions of (38). In order to apply the above definition to (SNP), one may take \({\mathbb{R}}^{n}\) as U or any open neighborhood of \(\overline{\Omega }\).

In Sect. 1.4 we have introduced the notion of properness of PDE (FE). The following example concerns this property.

Example 2.1.

Consider the boundary value problem (38) in the case where n = 1, Ω = (0, 1), \(U = \mathbb{R}\), \(F(x,p) = p - 1\) and \(B(x,p) = p - 1\). The function u(x) = x on [0, 1] is a classical solution of (38). But this function u is not a viscosity subsolution of (38). Indeed, if we take the test function ϕ(x) = 2x, then u − ϕ takes a maximum at x = 0 while we have \(B(0,\phi ^{\prime}(0)) = F(0,\phi ^{\prime}(0)) = 2 - 1 = 1 > 0\). However, if we reverse the direction of derivative at 0 by replacing the above B by the function

$$\displaystyle{B(x,p) = \left \{\begin{array}{@{}l@{\quad }l@{}} p - 1 \quad &\text{ for }x = 1,\\ -p + 1\quad &\text{ for } x = 0, \end{array} \right.}$$

then the function u is a classical solution of (38) as well as a viscosity solution of (38).

Definition 2.1.

The domain Ω is said to be of class C 1 (or simply Ω ∈ C 1) if there is a function \(\rho \in {C}^{1}({\mathbb{R}}^{n})\) which satisfies

$$\displaystyle\begin{array}{rcl} & & \Omega =\{ x \in {\mathbb{R}}^{n}\, :\,\rho (x) < 0\}, {}\\ & & D\rho (x)\not =0\ \ \ \text{ for all }x \in \partial \Omega. {}\\ \end{array}$$

The functions ρ having the above properties are called defining functions of Ω.

Remark 2.1.

If ρ is chosen as in the above definition, then the outer unit normal vector ν(x) at x ∈ ∂Ω is given by

$$\displaystyle{\nu (x) = \frac{D\rho (x)} {\vert D\rho (x)\vert }.}$$

Indeed, we have

$$\displaystyle{N(x,\Omega ) =\{ t\nu (x)\, :\, t \geq 0\}\ \ \ \text{ for all }x \in \partial \Omega.}$$

To see this, observe that if t ≥ 0, then \(\mathbf{1}_{\overline{\Omega }} + t\rho\) as a function in \({\mathbb{R}}^{n}\) attains a local maximum at any point \(x \in \partial \Omega \), which shows that

$$\displaystyle{t\vert D\rho (x)\vert \nu (x) \in -{D}^{+}\mathbf{1}_{\overline{ \Omega }}(x) = N(x,\overline{\Omega }).}$$

Next, let z ∈ ∂Ω and \(\phi \in {C}^{1}({\mathbb{R}}^{n})\) be such that \(\mathbf{1}_{\overline{\Omega }}-\phi\) attains a strict maximum over \({\mathbb{R}}^{n}\) at z. Observe that − ϕ attains a strict maximum over \(\overline{\Omega }\) at x. Fix a constant r > 0 and, for each \(k \in \mathbb{N}\), choose a maximum (over \(\overline{B}_{r}(z))\) point \(x_{k} \in \overline{B}_{r}(z)\) of − ϕ −  2, and observe that \(-(\phi +{k\rho }^{2})(x_{k}) \geq -(\phi +{k\rho }^{2})(z) = -\phi (z)\) for all \(k \in \mathbb{N}\) and that x k  → z as k → . For \(k \in \mathbb{N}\) sufficiently large we have

$$\displaystyle{D(\phi +{k\rho }^{2})(x_{ k}) = 0,}$$

and hence

$$\displaystyle{D\phi (x_{k}) = -2k\rho (x_{k})D\rho (x_{k}),}$$

which shows in the limit as k →  that

$$\displaystyle{D\phi (z) = -tD\rho (z) = -t\vert D\rho (z)\vert \nu (z),}$$

where \(t =\lim _{k\rightarrow \infty }2k\rho (x_{k}) \in \mathbb{R}\). Noting that \(-(\phi +{k\rho }^{2})(x) < -\phi (x) \leq -\phi (z)\) for all x ∈ Ω, we find that \(x_{k}\not\in \overline{B}_{r}(z) \setminus \Omega \) for all \(k \in \mathbb{N}\). Hence, we have t ≥ 0. Thus, we see that \(N(z,\overline{\Omega }) \subset \{ t\nu (z)\, :\, t \geq 0\}\) and conclude that \(N(z,\overline{\Omega }) =\{ t\nu (z)\, :\, t \geq 0\}\)

Henceforth in this section we assume that Ω is of class C 1.

Proposition 2.1.

If \(u \in {C}^{1}(\overline{\Omega })\) is a classical solution (resp., subsolution, or supersolution) of (SNP),then it is a viscosity solution (resp., subsolution, or supersolution) of (SNP).

Proof.

Let G be the function given by (39), with B(x, u, p) = γ(x) ⋅p − g(x). According to the above discussion on the equivalence between the notion of viscosity solution for (SNP) and that for PDE G(x, Du(x)) = 0 in \(\overline{\Omega }\) and Proposition 1.12, it is enough to show that the pair \((G,\overline{\Omega })\) is proper. From the above remark, we know that for any x ∈ ∂Ω we have \(N(x,\overline{\Omega }) =\{ t\nu (x)\, :\, t \geq 0\}\) and

$$\displaystyle{G(x,p + t\nu (x)) =\gamma (x) \cdot (p + t\nu (x)) \geq \gamma (x) \cdot p = G(x,p)\ \ \ \text{ for all }t \geq 0.}$$

As we noted before, we have \(N(x,\overline{\Omega }) =\{ 0\}\) if x ∈ Ω. Thus, we have for all \(x \in \overline{\Omega }\) and all \(q \in N(x,\overline{\Omega })\),

$$\displaystyle{G(x,p + q) \geq G(x,p).}$$

 □ 

We may treat in the same way the evolution problem

$$\displaystyle{ \left \{\begin{array}{llll} &u_{t}(x,t) + H(x,t,D_{x}u(x,t)) = 0&&\mathrm{in}\ \ \Omega \times J, \\ &\frac{\partial u} {\partial \gamma } (x,t) = g(x,t) &&\mathrm{on}\ \ \partial \Omega \times J,\end{array} \right. }$$
(40)

where J is an open interval in \(\mathbb{R}\), \(H\, :\, \overline{\Omega } \times J \times {\mathbb{R}}^{n} \rightarrow \mathbb{R}\) and \(g\, :\, \partial \Omega \times J \rightarrow \mathbb{R}\). If we set \(\widetilde{\Omega } = \Omega \times \mathbb{R}\), \(U = {\mathbb{R}}^{n} \times J\),

$$\displaystyle{F(x,t,p,q) = q + H(x,p)\ \ \text{ for }\ (x,t,p,q) \in \overline{\Omega } \times J \times {\mathbb{R}}^{n} \times \mathbb{R},}$$

and

$$\displaystyle{B(x,t,p,q) =\gamma (x) \cdot p - g(x,t)\ \ \text{ for }\ (x,t,p,q) \in \partial \Omega \times J \times {\mathbb{R}}^{n} \times \mathbb{R},}$$

then the viscosity formulation for (38) applies to (40), with Ω replaced by \(\widetilde{\Omega }\).

We note here that if ρ is a defining function of Ω, then it, as a function of (x, t), is also a defining function of the “cylinder” \(\Omega \times \mathbb{R}\). Hence, if we set \(\tilde{\gamma }(x,t) = (\gamma (x),0)\) and \(\tilde{\nu }(x,t) = (\nu (x),0)\) for \((x,t) \in \partial (\Omega \times \mathbb{R}) = \partial \Omega \times \mathbb{R}\), then \(\tilde{\nu }(x,t)\) is the outer unit normal vector at \((x,t) \in \partial \Omega \times \mathbb{R}\). Moreover, if γ satisfies (37), then we have \(\tilde{\gamma }(x,t) \cdot \tilde{\nu } (x,t) =\gamma (x) \cdot \nu (x) > 0\) for all \((x,t) \in \partial \Omega \times \mathbb{R}\). Thus, as Proposition 2.1 says, if (37) holds, then any classical solution (resp., subsolution or supersolution) of (40) is a viscosity solution (resp., subsolution or supersolution) of (40).

Before closing this subsection, we add two lemmas concerning C 1 domains.

Lemma 2.1.

Let Ω be a bounded, open, connected subset of \(\,{\mathbb{R}}^{n}\) . Assume that Ω is of class C 1 . Then there exists a constant C > 0 and, for each \(x,y \in \overline{\Omega }\) with x≠y, a curve η ∈AC ([0, t(x,y)]), with t(x,y) > 0, such that t(x,y) ≤ C|x − y|, η(s) ∈ Ω for all s ∈ (0, t(x,y)), and \(\vert \dot{\eta }(s)\vert \leq 1\) for a.e. s ∈ [0, t(x,y)].

Lemma 2.2.

Let Ω be a bounded, open, connected subset of \(\,{\mathbb{R}}^{n}\) . Assume that Ω is of class C 1 . Let M > 0 and u ∈ C(Ω) be a viscosity subsolution of |Du(x)|≤ M in Ω. Then the function u is Lipschitz continuous in Ω.

The proof of these lemmas is given in Appendix A.3.

3 Initial-Boundary Value Problem for Hamilton–Jacobi Equations

We study the initial value problem for Hamilton–Jacobi equations with the Neumann boundary condition.

To make the situation clear, we collect our assumptions on Ω, γ and H.

  1. (A1)

    Ω is bounded open connected subset of \({\mathbb{R}}^{n}\).

  2. (A2)

    Ω is of class C 1.

  3. (A3)

    \(\gamma \in C(\partial \Omega, {\mathbb{R}}^{n})\) and \(g \in C(\partial \Omega, \mathbb{R})\).

  4. (A4)

    γ(x) ⋅ν(x) > 0 for all x ∈ ∂ΩD.

  5. (A5)

    \(H \in C(\overline{\Omega } \times {\mathbb{R}}^{n})\).

  6. (A6)

    H is coercive, i.e.,

    $$\displaystyle{\lim _{R\rightarrow \infty }\inf \{H(x,p)\, :\, (x,p) \in \overline{\Omega } \times {\mathbb{R}}^{n},\ \vert p\vert \geq R\} = \infty.}$$

In what follows, we assume always that (A1)–(A6) hold.

3.1 Initial-Boundary Value Problems

Given a function \(u_{0} \in C(\overline{\Omega })\), we consider the problem of evolution type

$$\displaystyle{ \left \{\begin{array}{llll} &u_{t} + H(x,D_{x}u) = 0&&\mathrm{in}\ \Omega \times (0,\infty ), \\ &\gamma (x) \cdot D_{x}u = g(x) &&\mathrm{on}\ \partial \Omega \times (0,\,\infty ),\\ \end{array} \right. }$$
(ENP)
$$\displaystyle{ \quad \ \ u(x,0) = u_{0}(x)\qquad \ \text{ for }x \in \overline{\Omega }. }$$
(ID)

Here u = u(x, t) is a function of \((x,t) \in \overline{\Omega } \times [0,\infty )\) and represents the unknown function.

When we say u is a (viscosity) solution of (ENP)–(ID), u is assumed to satisfy the initial condition (ID) in the pointwise (classical) sense.

Henceforth Q denotes the set \(\overline{\Omega } \times (0,\,\infty )\).

Theorem 3.1 (Comparison). 

Let \(u \in \mathrm{USC}(\overline{Q})\) and \(v \in \mathrm{LSC}(\overline{Q})\) be a viscosity subsolution and supersolution of (ENP) , respectively. Assume furthermore that u(x,0) ≤ v(x,0) for all \(x \in \overline{\Omega }\) . Then u ≤ v in Q.

To proceed, we concede the validity of the above theorem and will come back to its proof in Sect. 3.3.

Remark 3.1.

The above theorem guarantees that if u is a viscosity solution of (ENP)–(ID) and continuous for t = 0, then it is unique.

Theorem 3.2 (Existence). 

There exists a viscosity solution u of (ENP) (ID) in the space \(C(\overline{Q})\).

Proof.

Fix any \(\varepsilon \in (0,\,1)\). Choose a function \(u_{0,\varepsilon } \in {C}^{1}(\overline{\Omega })\) so that

$$\displaystyle{\vert u_{0,\varepsilon }(x) - u_{0}(x)\vert <\varepsilon \ \ \ \ \text{ for all }\ x \in \overline{\Omega }.}$$

Let \(\rho \in {C}^{1}({\mathbb{R}}^{n})\) be a defining function of Ω. Since

$$\displaystyle{D\rho (x) = \vert D\rho (x)\vert \nu (x)\ \ \ \text{ for }x \in \partial \Omega,}$$

we may choose a constant \(M_{\varepsilon } > 0\) so large that

$$\displaystyle{M_{\varepsilon }\gamma (x) \cdot D\rho (x) \geq \max _{\partial \Omega }(\vert g\vert + \vert \gamma \cdot \mathit{Du}_{0,\varepsilon }\vert )\ \ \ \ \text{ for all }\ x \in \partial \Omega.}$$

Next choose a function \(\zeta \in {C}^{1}(\mathbb{R})\) so that

$$\displaystyle{\left \{\begin{array}{llll} \zeta ^{\prime}(0) & = 1, && \\ - 1 \leq &\zeta (r) \leq 0&&\text{ for }\ r \leq 0, \\ 0 \leq &\zeta ^{\prime}(r) \leq 1\ \ \ &&\text{ for }\ r \leq 0. \end{array} \right.}$$

Setting

$$\displaystyle{\chi _{\varepsilon }(x) =\varepsilon \zeta (M_{\varepsilon }\rho (x)/\varepsilon ),}$$

we have

$$\displaystyle{\left \{\begin{array}{llll} &-\varepsilon \leq \chi _{\varepsilon }(x) \leq 0\ \ \ \ &&\text{ for all }\ x \in \overline{\Omega },\\ &\gamma (x) \cdot D\chi _{\varepsilon }(x) \geq \vert g(x)\vert + \vert \gamma (x) \cdot \mathit{Du } _{ 0,\varepsilon }(x)\vert \ \ \ &&\text{ for all }\ x \in \partial \Omega, \end{array} \right.}$$

and we may choose a constant \(C_{\varepsilon } > 0\) such that

$$\displaystyle{\vert D\chi _{\varepsilon }(x)\vert \leq C_{\varepsilon }\ \ \ \text{ for all }\ x \in \overline{\Omega }.}$$

Then define the functions \(f_{\varepsilon }^{\pm }\in {C}^{1}(\overline{\Omega })\) by

$$\displaystyle{f_{\varepsilon }^{\pm }(x) = u_{ 0,\varepsilon }(x) \pm (\chi _{\varepsilon }(x) + 2\varepsilon ),}$$

and observe that

$$\displaystyle{\left \{\begin{array}{llll} &u_{0}(x) \leq f_{\varepsilon }^{+}(x) \leq u_{0}(x) + 3\varepsilon \ &&\text{ for all }\ x \in \overline{\Omega }, \\ &u_{0}(x) \geq f_{\varepsilon }^{-}(x) \geq u_{0}(x) - 3\varepsilon \ &&\text{ for all }\ x \in \overline{\Omega }, \\ &\gamma (x) \cdot Df_{\varepsilon }^{+}(x) \geq g(x)\ \ &&\text{ for all }\ x \in \partial \Omega, \\ &\gamma (x) \cdot Df_{\varepsilon }^{-}(x) \leq g(x)\ \ &&\text{ for all }\ x \in \partial \Omega. \end{array} \right.}$$

Now, we choose a constant \(A_{\varepsilon } > 0\) large enough so that

$$\displaystyle{\vert H(x,Df_{\varepsilon }^{\pm }(x))\vert \leq A_{\varepsilon }\ \ \ \text{ for all }\ x \in \overline{\Omega },}$$

and set

$$\displaystyle{g_{\varepsilon }^{\pm }(x,t) = f_{\varepsilon }^{\pm }(x) \pm A_{\varepsilon }t\ \ \ \text{ for }(x,t) \in \overline{Q}.}$$

The functions \(g_{\varepsilon }^{+},\,g_{\varepsilon }^{-}\in {C}^{1}(\overline{Q})\) are a viscosity supersolution and subsolution of (ENP), respectively.

Setting

$$\displaystyle\begin{array}{rcl} {h}^{+}(x,t)& =& \,\inf \{g_{\varepsilon }^{+}(x,t)\, :\,\varepsilon \in (0,\,1)\}, {}\\ {h}^{-}(x,t)& =& \,\sup \{g_{\varepsilon }^{-}(x,t)\, :\,\varepsilon \in (0,\,1)\}, {}\\ \end{array}$$

we observe that \({h}^{+} \in \mathrm{USC}(\overline{Q})\) and \({h}^{-}\in \mathrm{LSC}(\overline{Q})\) are, respectively, a viscosity supersolution and subsolution of (ENP). Moreover we have

$$\displaystyle\begin{array}{rcl} & & \qquad u_{0}(x) = {h}^{\pm }(x,0)\ \ \ \text{ for all }\ x \in \overline{\Omega }, {}\\ & & {h}^{-}(x,t) \leq u_{ 0}(x) \leq {h}^{+}(x,t)\ \ \ \text{ for all }\ (x,t) \in \overline{Q}. {}\\ \end{array}$$

By Theorem 1.4, we find that there exists a viscosity solution u of (ENP) which satisfies

$$\displaystyle{{h}^{-}(x,t) \leq u(x,t) \leq {h}^{+}(x,t)\ \ \ \text{ for all }\ (x,t) \in \overline{Q}.}$$

Applying Theorem 3.1 to u  ∗  and u  ∗  yields

$$\displaystyle{{u}^{{\ast}}\leq u_{ {\ast}}\ \ \ \ \text{ for all }\ (x,t) \in \overline{Q},}$$

while \(u_{{\ast}}\leq {u}^{{\ast}}\) in \(\overline{Q}\) by definition, which in particular implies that \(u \in C(\overline{Q})\). The proof is complete. □ 

Theorem 3.3 (Uniform continuity). 

The viscosity solution \(u \in C(\overline{Q})\) of (ENP) (ID) is uniformly continuous in \(\overline{Q}\) . Furthermore, if \(u_{0} \in \mathrm{Lip}(\overline{\Omega })\) , then \(u \in \mathrm{Lip}(\overline{Q})\).

Lemma 3.1.

Let \(u_{0} \in \mathrm{Lip}(\overline{\Omega })\) . Then there is a constant C > 0 such that the functions u 0 (x) + Ct and u 0 (x) − Ct are, respectively, a viscosity supersolution and subsolution of (ENP) (ID).

Proof.

Let ρ and ζ be the functions which are used in the proof of Theorem 3.2. Choose the collection \(\{u_{0,\varepsilon }\}_{\varepsilon \in (0,\,1)} \subset {C}^{1}(\overline{\Omega })\) of functions so that

$$\displaystyle{\left \{\begin{array}{l} \lim _{\varepsilon \rightarrow 0}\|u_{0,\varepsilon } - u_{0}\|_{\infty,\Omega } = 0, \\ \sup _{\varepsilon \in (0,\,1)}\|\mathit{Du}_{0,\varepsilon }\|_{\infty,\Omega } < \infty.\end{array} \right.}$$

As in the proof of Theorem 3.2, we may fix a constant M > 0 so that

$$\displaystyle\begin{array}{rcl} M\gamma (x) \cdot D\rho (x)& =& M\vert D\rho (x)\vert \nu (x) \cdot \gamma (x) {}\\ & \geq & \vert g(x)\vert + \vert \gamma (x) \cdot \mathit{Du}_{0,\varepsilon }(x)\vert \ \ \ \text{ for all }\ x \in \partial \Omega. {}\\ \end{array}$$

Next set

$$\displaystyle{R =\sup _{\varepsilon \in (0,\,1)}\|\mathit{Du}_{0,\varepsilon }\|_{\infty,\Omega } + M\|D\rho \|_{\infty,\Omega },}$$

and choose C > 0 so that

$$\displaystyle{\max _{\overline{\Omega }\times \overline{B}_{ R}}\vert H\vert \leq C.}$$

Now, we put

$$\displaystyle{v_{\varepsilon }^{\pm }(x,t) = u_{ 0,\varepsilon }(x) \pm (M\varepsilon \zeta (\rho (x)/\varepsilon ) + \mathit{Ct})\ \ \ \text{ for }\ (x,t) \in \overline{Q},}$$

and note that \(v_{\varepsilon }^{+}\) and \(v_{\varepsilon }^{-}\) are a classical supersolution and subsolution of (ENP). Sending \(\varepsilon \rightarrow 0+\), we conclude by Proposition 1.9 that the functions u 0(x) + Ct and u 0(x) − Ct are a viscosity supersolution and subsolution of (ENP), respectively. □ 

Proof (Theorem 3.3). 

We first assume that \(u_{0} \in \mathrm{Lip}(\overline{\Omega })\), and show that \(u \in \mathrm{Lip}(\overline{Q})\). According to Lemma 3.1, there exists a constant C > 0 such that the function \(u_{0}(x) -\mathit{Ct}\) is a viscosity subsolution of (ENP). By Theorem 3.1, we get

$$\displaystyle{u(x,t) \geq u_{0}(x) -\mathit{Ct}\ \ \ \ \text{ for all }\ (x,t) \in \overline{Q}.}$$

Fix any t > 0, and apply Theorem 3.1 to the functions u(x, t + s) and \(u(x,s) -\mathit{Ct}\) of (x, s), both of which are viscosity solutions of (ENP), to get

$$\displaystyle{u(x,t + s) \geq u(x,s) -\mathit{Ct}\ \ \ \ \text{ for all }\ (x,s) \in \overline{Q}.}$$

Hence, if (p, q) ∈ D  +  u(x, s), then we find that as t → 0 + ,

$$\displaystyle{u(x,s) \leq u(x,s + t) + \mathit{Ct} \leq u(x,s) + \mathit{qt} + \mathit{Ct} + o(t),}$$

and consequently, q ≥ − C. Moreover, if x ∈ Ω, we have

$$\displaystyle{0 \geq q + H(x,p) \geq H(x,p) - C.}$$

Due to the coercivity of H, there exists a constant R > 0 such that

$$\displaystyle{p \in B_{R}.}$$

Therefore, we get

$$\displaystyle{q \leq -H(x,p) \leq \max _{\overline{\Omega }\times \overline{B}_{ R}}\vert H\vert.}$$

Thus, if (x, s) ∈ Ω ×(0, ) and (p, q) ∈ D  +  u(x, s), then we have

$$\displaystyle{\vert p\vert + \vert q\vert \leq M := R + C +\max _{\overline{\Omega }\times \overline{B}_{ R}}\vert H\vert.}$$

Thanks to Proposition 1.14, we conclude that u is Lipschitz continuous in Q.

Next, we show in the general case that \(u \in \mathrm{UC}(\overline{Q})\). Let \(\varepsilon \in (0,\,1)\), and choose a function \(u_{0,\varepsilon } \in \mathrm{Lip}(\overline{\Omega })\) so that

$$\displaystyle{\|u_{0,\varepsilon } - u_{0}\|_{\infty }\leq \varepsilon.}$$

Let \(u_{\varepsilon }\) be the viscosity solution of (ENP) satisfying the initial condition

$$\displaystyle{u_{\varepsilon }(x,0) = u_{0,\varepsilon }(x)\ \ \ \text{ for all }\ x \in \overline{\Omega }.}$$

As we have shown above, we know that \(u_{\varepsilon } \in \mathrm{Lip}(\overline{Q})\). Moreover, by Theorem 3.1 we have

$$\displaystyle{\|u_{\varepsilon } - u\|_{\infty,\overline{Q}} \leq \varepsilon.}$$

It is now obvious that \(u \in \mathrm{UC}(\overline{Q})\). □ 

3.2 Additive Eigenvalue Problems

Under our hypotheses (A1)–(A6), the boundary value problem

$$\displaystyle{ \left \{\begin{array}{llll} &H(x,\mathit{Du}) = 0 &&\mathrm{in}\ \Omega,\\ &\gamma (x) \cdot \mathit{Du } = g(x)&&\mathrm{on}\ \partial \Omega \end{array} \right. }$$
(SNP)

may not have a viscosity solution. For instance, the Hamiltonian H(x, p) =  | p | 2 + 1 satisfies (A5) and (A6), but, since H(x, p) > 0, (SNP) does not have any viscosity subsolution.

Instead of (SNP), we consider the additive eigenvalue problem

$$\displaystyle{ \left \{\begin{array}{llll} &H(x,Dv) = a &&\mathrm{in}\ \Omega,\\ &\gamma (x) \cdot Dv = g(x)&&\mathrm{on}\ \partial \Omega. \end{array} \right. }$$
(EVP)

This is a problem to seek for a pair \((a,v) \in \mathbb{R} \times C(\overline{\Omega })\) such that v is a viscosity solution of the stationary problem (EVP). If \((a,v) \in \mathbb{R} \times C(\overline{\Omega })\) is such a pair, then a and v are called an (additive) eigenvalue and eigenfunction of (EVP), respectively. This problem is often called the ergodic problem in the viewpoint of ergodic optimal control.

Theorem 3.4.

  1. (i)

    There exists a solution \((a,v) \in \mathbb{R} \times \mathrm{Lip}(\overline{\Omega })\) of (EVP).

  2. (ii)

    The eigenvalue of (EVP) is unique. That is, if \((a,v),\,(b,w) \in \mathbb{R} \times C(\overline{\Omega })\) are solutions of (EVP) , then a = b.

The above result has been obtained by Lions et al., Homogenization of Hamilton-Jacobi equations, unpublished.

In what follows we write c # for the unique eigenvalue a of (EVP).

Corollary 3.1.

Let \(u \in C(\overline{Q})\) be the solution of (ENP) (ID) . Then the function u(x,t) + c # t is bounded on \(\overline{Q}\).

Corollary 3.2.

We have

$$\displaystyle{{c}^{\#} =\inf \{ a \in \mathbb{R}\, :\, \mathrm{EVP}\text{ has a viscosity subsolution }v\}.}$$

Lemma 3.2.

Let \(b,c \in \mathbb{R}\) and \(v,w \in C(\overline{\Omega })\) . Assume that v (resp., w) is a viscosity supersolution (resp., subsolution) of (EVP) with a = b (resp., a = c). Then b ≤ c.

Remark 3.2.

As the following proof shows, the assertion of the above lemma is valid even if one replaces the continuity of v and w by the boundedness.

Proof.

By adding a constant to v if needed, we may assume that v ≥ w in \(\overline{\Omega }\). Since the functions v(x) − bt and w(x) − ct are a viscosity supersolution and subsolution of (ENP), by Theorem 3.1 we get

$$\displaystyle{v(x) - bt \geq w(x) - ct\ \ \ \text{ for all }\ (x,t) \in \overline{Q},}$$

from which we conclude that b ≤ c. □ 

Proof (Theorem 3.4). 

Assertion (ii) is a direct consequence of Lemma 3.2.

We prove assertion (i). Consider the boundary value problem

$$\displaystyle{ \left \{\begin{array}{llll} &\lambda v + H(x,Dv) = 0&&\mathrm{in}\ \Omega,\\ &\gamma (x) \cdot Dv = g&&\mathrm{on}\ \partial \Omega, \end{array} \right. }$$
(41)

where λ > 0 is a given constant. We will take the limit as λ → 0 later on.

We fix λ ∈ (0, 1). Let \(\rho \in {C}^{1}({\mathbb{R}}^{n})\) be a defining function of the domain Ω. Select a constant A > 0 so large that (x) ⋅(x) ≥ | g(x) | for all x ∈ ∂Ω, and then B > 0 so large that B ≥ A | ρ(x) |  +  | H(x, ± ADρ(x)) | for all \(x \in \overline{\Omega }\). Observe that the functions (x) + B ∕ λ and − (x) − B ∕ λ are a classical supersolution and subsolution of (41), respectively.

The Perron method (Theorem 1.4) guarantees that there is a viscosity solution v λ of (41) which satisfies

$$\displaystyle{\vert v_{\lambda }(x)\vert \leq A\rho (x) + B/\lambda \leq B/\lambda \ \ \ \ \text{ for all }\ x \in \overline{\Omega }.}$$

Now, since

$$\displaystyle{-\lambda v_{\lambda }(x) \leq B\ \ \ \ \text{ for all }\ x \in \overline{\Omega },}$$

v λ satisfies in the viscosity sense

$$\displaystyle{H(x,Dv_{\lambda }(x)) \leq B\ \ \ \ \text{ for all }\ x \in \Omega,}$$

which implies, together with the coercivity of H, the equi-Lipschitz continuity of \(\{v_{\lambda }\}_{\lambda \in (0,\,1)}\). Thus the collections \(\{v_{\lambda } -\inf _{\Omega }v_{\lambda }\}_{\lambda \in (0,\,1)}\) and \(\{\lambda v_{\lambda }\}_{\lambda \in (0,\,1)}\) of functions on \(\overline{\Omega }\) are relatively compact in \(C(\overline{\Omega })\). We may select a sequence \(\{\lambda _{j}\}_{j\in \mathbb{N}} \subset (0,\,1)\) such that

$$\displaystyle\begin{array}{rcl} & & \lambda _{j} \rightarrow 0, {}\\ & & v_{\lambda _{j}}(x) -\inf _{\Omega }v_{\lambda _{j}} \rightarrow v(x), {}\\ & & \lambda _{j}v_{\lambda _{j}}(x) \rightarrow w(x) {}\\ \end{array}$$

for some functions \(v,w \in C(\overline{\Omega })\) as j → , where the convergences to v and w are uniform on \(\overline{\Omega }\). Observe that for all \(x \in \overline{\Omega }\),

$$\displaystyle\begin{array}{rcl} w(x)& =& \,\lim _{j\rightarrow \infty }\lambda _{j}v_{\lambda _{j}}(x) {}\\ & =& \,\lim _{j\rightarrow \infty }\lambda _{j}\big[(v_{\lambda _{j}}(x) -\inf _{\Omega }v_{\lambda _{j}}) +\inf _{\Omega }v_{\lambda _{j}}\big] {}\\ & =& \,\lim _{j\rightarrow \infty }\lambda _{j}\inf _{\Omega }v_{\lambda _{j}}, {}\\ \end{array}$$

which shows that w is constant on \(\overline{\Omega }\). If we write this constant as a, then we see by Proposition 1.9 that v is a viscosity solution of (EVP). This completes the proof of (i). □ 

Proof (Corollary 3.1). 

Let \(v \in C(\overline{\Omega })\) be an eigenfunction of (EVP). That is, v is a viscosity solution of (EVP), with a = c #. Then, for any constant \(C \in \mathbb{R}\), the function w(x, t) : = v(x) − c # t + C is a viscosity solution of (ENP). We may choose constants C i , i = 1, 2, so that \(v(x) + C_{1} \leq u_{0}(x) \leq v(x) + C_{2}\) for all \(x \in \overline{\Omega }\). By Theorem 3.1, we see that

$$\displaystyle{v(x) - {c}^{\#}t + C_{ 1} \leq u(x,t) \leq v(x) - {c}^{\#}t + C_{ 2}\ \ \ \text{ for all }\ (x,t) \in \overline{Q},}$$

which shows that the function u(x, t) + c # t is bounded on \(\overline{Q}\). □ 

Proof (Corollary 3.2). 

It is clear that

$$\displaystyle{{c}^{\#} \geq {c}^{\star } :=\inf \{ a \in \mathbb{R}\, :\, \text{(EVP) has a viscosity subsolution }v\}.}$$

To show that \({c}^{\#} \leq {c}^{\star }\), we suppose by contradiction that c # > c  ⋆ . By the definition of \({c}^{\star }\), there is a b ∈ [c  ⋆ , c #) such that (EVP), with a = b, has a viscosity subsolution ψ. Let v be a viscosity solution of (EVP), with a = c #. Since b < c #, v is a viscosity supersolution of (EVP), with a = b. We may assume that ψ ≤ v in \(\overline{\Omega }\). Theorem 1.4 now guarantees the existence of a viscosity solution of (EVP), which contradicts Theorem 3.4, (ii) (see Remark 3.2). □ 

Example 3.1.

We consider the case where n = 1, Ω = ( − 1, 1), H(x, p) = H(p) : =  | p | and γ( ± 1) =  ± 1, respectively, and evaluate the eigenvalue c #. We set g min = min{g( − 1), g(1)}. Assume first that g min ≥ 0. In this case, the function v(x) = 0 is a classical subsolution of (SNP) and, hence, c # ≤ 0. On the other hand, since H(p) ≥ 0 for all \(p \in \mathbb{R}\), we have c # ≥ 0. Thus, c # = 0. We next assume that g min < 0. It is easily checked that if g(1) = g min, then the function v(x) = g min x is a viscosity solution of (EVP), with a =  | g min | . (Notice that

$$\displaystyle\begin{array}{rcl} -{D}^{+}v(-1)& =& (-\infty,\,-\vert g_{ \text{min}}\vert \,]\ \cup \ [\,-\vert g_{\text{min}}\vert,\,\vert g_{\text{min}}\vert \,], {}\\ -{D}^{-}v(-1)& =& [\,\vert g_{ \text{min}}\vert,\,\infty ).) {}\\ \end{array}$$

Similarly, if g( − 1) = g min, then the function v(x) =  | g min | x is a viscosity solution of (EVP), with a =  | g min | . These observations show that c # =  | g min | .

3.3 Proof of Comparison Theorem

This subsection will be devoted to the proof of Theorem 3.1.

We begin with the following two lemmas.

Lemma 3.3.

Let u be the function from Theorem 3.1 . Set P = Ω × (0, ∞). Then, for every (x,t) ∈ ∂Ω × (0, ∞), we have

$$\displaystyle{ u(x,t) =\limsup _{P\ni (y,s)\rightarrow (x,t)}u(y,s). }$$
(42)

Proof.

Fix any (x, t) ∈ ∂Ω ×(0, ). To prove (42), we argue by contradiction, and suppose that

$$\displaystyle{\limsup _{P\ni (y,s)\rightarrow (x,t)}u(y,s) < u(x,t).}$$

We may choose a constant r ∈ (0, t) so that

$$\displaystyle{ u(y,s) + r < u(x,t)\ \ \ \text{ for all }(y,s) \in P \cap (\overline{B}_{r}(x) \times [t - r,\,t + r]). }$$
(43)

Note that

$$\displaystyle{P \cap (\overline{B}_{r}(x) \times [t - r,\,t + r]) = (\Omega \cap \overline{B}_{r}(x)) \times [t - r,\,t + r].}$$

Since u is bounded on \(\overline{\Omega } \times [t - r,\,t + r]\), we may choose a constant α > 0 so that for all \((y,s) \in \overline{\Omega } \times [t - r,t + r]\),

$$\displaystyle{ u(y,s) + r -\alpha (\vert y - x{\vert }^{2} + {(s-t)}^{2}) < u(x,t)\ \ \mathrm{if}\ \ \vert y-x\vert \geq r/2\ \text{ or }\ \vert s - t\vert \geq r/2. }$$
(44)

Let ρ be a defining function of Ω. Let ζ be the function on \(\mathbb{R}\) introduced in the proof of Theorem 3.2. For \(k \in \mathbb{N}\) we define the function \(\psi \in {C}^{1}({\mathbb{R}}^{n+1})\) by

$$\displaystyle{\psi (y,s) = {k}^{-1}\zeta ({k}^{2}\rho (y)) +\alpha (\vert y - x{\vert }^{2} + {(s - t)}^{2}).}$$

Consider the function

$$\displaystyle{u(y,s) -\psi (y,s)}$$

on the set \(\left (\overline{\Omega } \cap \overline{B}_{r}(x)\right ) \times [t - r,\,t + r]\). Let \((y_{k},s_{k}) \in \left (\overline{\Omega } \cap \overline{B}_{r}(x)\right ) \times [t - r,\,t + r]\) be a maximum point of the above function. Assume that k > r  − 1.

Using (43) and (44), we observe that for all \((y,s) \in (\overline{\Omega } \cap \overline{B}_{r}(x)) \times [t - r,\,t + r]\),

$$\displaystyle{u(y,s) -\psi (y,s) < u(x,t) = u(x,t) -\psi (x,t)}$$

if either  y ∈ Ω,   | y − x | ≥ r ∕ 2,  or   | s − t | ≥ r ∕ 2. Accordingly, we have

$$\displaystyle{(y_{k},s_{k}) \in \left (\partial \Omega \cap \overline{B}_{r/2}(x)\right ) \times (t - r/2,\,t + r/2).}$$

Hence, setting

$$\displaystyle{p_{k} = kD\rho (y_{k}) + 2\alpha (y_{k} - x)\ \ \text{ and }\ \ q_{k} = 2\alpha (s_{k} - t),}$$

we have

$$\displaystyle{\min \{q_{k} + H(y_{k},p_{k}),\,\gamma (y_{k}) \cdot p_{k} - g(y_{k})\} \leq 0.}$$

If we note that

$$\displaystyle{\gamma (y_{k}) \cdot D\rho (y_{k}) \geq \min _{\partial \Omega }\gamma \cdot D\rho > 0,}$$

then, by sending k → , we get a contradiction. □ 

Lemma 3.4.

Let \(y,z \in {\mathbb{R}}^{n}\) , and assume that y ⋅ z > 0. Then there exists a quadratic function ζ in \({\mathbb{R}}^{n}\) which satisfies:

$$\displaystyle{\left \{\begin{array}{llll} &\zeta (tx) = {t}^{2}\zeta (x) &&\text{ for all }\ (x,t) \in {\mathbb{R}}^{n} \times \mathbb{R}, \\ &\zeta (x) > 0 &&\mathrm{if}\ \ x\not =0, \\ &z \cdot D\zeta (x) = 2(y \cdot z)(y \cdot x)&&\text{ for all }\ x \in {\mathbb{R}}^{n}. \end{array} \right.}$$

Proof.

We define the function ζ by

$$\displaystyle{\zeta (x) =\Big \vert x -\frac{y \cdot x} {y \cdot z}z\Big{\vert }^{2} + {(y \cdot x)}^{2}.}$$

We observe that for any \(t \in \mathbb{R}\),

$$\displaystyle\begin{array}{rcl} \zeta (x + tz)& =& \,{\left \vert x + tz -\frac{y \cdot (x + tz)} {y \cdot z} z\right \vert }^{2} + {(y \cdot (x + tz))}^{2} {}\\ & =& \,{\left \vert x -\frac{y \cdot x} {y \cdot z}z\right \vert }^{2} + {(y \cdot x)}^{2} + 2t(y \cdot x)(y \cdot z) + {t}^{2}{(y,z)}^{2}, {}\\ & & {}\\ \end{array}$$

from which we find that

$$\displaystyle{z \cdot D\zeta (x) = 2(y \cdot z)(y \cdot x).}$$

If ζ(x) = 0, then yx = 0 and

$$\displaystyle{0 =\zeta (x) =\Big \vert x -\frac{y \cdot x} {y \cdot z}z\Big{\vert }^{2} = \vert x{\vert }^{2}.}$$

Hence, we have x = 0 if ζ(x) = 0, which shows that ζ(x) > 0 if x ≠ 0. It is obvious that the function ζ is homogeneous of degree two. The function ζ has the required properties. □ 

For the proof of Theorem 3.1, we argue by contradiction: we suppose that

$$\displaystyle{\sup _{\overline{\Omega }\times [0,\infty )}(u - v) > 0,}$$

and, to conclude the proof, we will get a contradiction.

We may assume that there exist a constant δ > 0 and a finite open interval J ⊂ (0, ) such that

$$\displaystyle\begin{array}{rcl} \begin{array}{ll} &u\ \text{is a viscosity subsolution of} \\ &\left \{\begin{array}{llll} &u_{t}(x,t) + H(x,D_{x}u(x,t)) \leq -\delta &&\mathrm{in}\ \Omega \times J, \\ &\gamma (x) \cdot D_{x}u(x,t) \leq g(x) &&\mathrm{on}\ \partial \Omega \times J,\end{array} \right.\end{array}& &{}\end{array}$$
(45)
$$\displaystyle\begin{array}{rcl} \max _{\overline{\Omega }\times \overline{J}}(u - v) > 0 >\max _{\overline{\Omega }\times \partial J}(u - v),& &{}\end{array}$$
(46)

and

$$\displaystyle{ u\ \text{and}\ v\ \text{are bounded on}\ \overline{\Omega } \times \overline{J}. }$$
(47)

Proof.

We choose a T > 0 so that  \(\sup _{\overline{\Omega }\times (0,\,T)}(u - v) > 0\)  and set

$$\displaystyle{u_{\varepsilon }(x,t) = u(x,t) - \frac{\varepsilon } {T - t}\ \ \ \text{ for }\ (x,t) \in \overline{\Omega } \times [0,\,T),}$$

where \(\varepsilon > 0\) is a constant. It is then easy to check that \(u_{\varepsilon }\) is a viscosity subsolution of

$$\displaystyle{\left \{\begin{array}{llll} &u_{\varepsilon,t} + H(x,D_{x}u_{\varepsilon }(x,t)) \leq - \frac{\varepsilon }{{T}^{2}} & & \mathrm{in}\ \Omega \times (0,T), \\ &\frac{\partial u_{\varepsilon }} {\partial \gamma } (x,t) \leq g(x) &&\mathrm{on}\ \partial \Omega \times (0,\,T). \end{array} \right.}$$

Choosing \(\varepsilon > 0\) sufficiently small, we have

$$\displaystyle{\sup _{\overline{\Omega }\times [0,T)}(u_{\varepsilon } - v) > 0 >\max _{\overline{\Omega }\times \{0\}}(u_{\varepsilon } - v).}$$

If we choose α > 0 sufficiently small, then

$$\displaystyle{\max _{\overline{\Omega }\times [0,T-\alpha ]}(u_{\varepsilon } - v) > 0 >\max _{\overline{\Omega }\times \partial [0,T-\alpha ]}(u_{\varepsilon } - v).}$$

Thus, if we set J = (0, T − α) and replace u by \(u_{\varepsilon }\), then we are in the situation of (45)–(47). □ 

We may assume furthermore that \(u \in \mathrm{Lip}(\overline{\Omega } \times \overline{J})\) as follows.

We may assume that there exist a constant δ > 0 and a finite open interval J ⊂ (0, ) such that

$$\displaystyle\begin{array}{rcl} \begin{array}{ll} &u\ \text{is a viscosity subsolution of} \\ &\left \{\begin{array}{llll} &u_{t}(x,t) + H(x,D_{x}u(x,t)) \leq -\delta &&\mathrm{in}\ \Omega \times J, \\ &\gamma (x) \cdot D_{x}u(x,t) \leq g(x) &&\mathrm{on}\ \partial \Omega \times J,\end{array} \right.\end{array}& &{}\end{array}$$
(48)
$$\displaystyle\begin{array}{rcl} \max _{\overline{\Omega }\times \overline{J}}(u - v) > 0 >\max _{\overline{\Omega }\times \partial J}(u - v),& &{}\end{array}$$
(49)

and

$$\displaystyle{ u \in \mathrm{Lip}(\overline{\Omega } \times \overline{J})\ \text{and}\ v\ \text{is bounded on}\ \overline{\Omega } \times \overline{J}. }$$
(50)

Proof.

Let J be as in Reduction 1. We set J = (a, b). Let M > 0 be a bound of | u | on \(\overline{\Omega } \times [a,\,b]\).

For each \(\varepsilon > 0\) we define the sup-convolution in the t-variable

$$\displaystyle{u_{\varepsilon }(x,t) =\max _{s\in [a,b]}\Big(u(x,s) -\frac{{(t - s)}^{2}} {2\varepsilon } \Big).}$$

We note as in Sect. 1.6 that

$$\displaystyle{M \geq u_{\varepsilon }(x,t) \geq u(x,t) \geq -M\ \ \ \text{ for all }\ (x,t) \in \overline{\Omega } \times [a,\,b].}$$

Noting that

$$\displaystyle{ \frac{1} {2\varepsilon }{(t - s)}^{2} \leq 2M\ \;\Longleftrightarrow\;\ \vert t - s\vert \leq 2\sqrt{\varepsilon M} }$$
(51)

and setting \(m_{\varepsilon } = 2\sqrt{\varepsilon M}\), we find that

$$\displaystyle{u_{\varepsilon }(x,t) =\max _{a<s<b}\Big(u(x,s) -\frac{{(t - s)}^{2}} {2\varepsilon } \Big)\ \ \ \text{ for all }\ (x,t) \in \overline{\Omega } \times (a + m_{\varepsilon },\,b - m_{\varepsilon }).}$$

Let \((x,t) \in \overline{\Omega } \times (a + m_{\varepsilon },\,b - m_{\varepsilon })\). Choose an s ∈ (a, b) so that

$$\displaystyle{u_{\varepsilon }(x,t) = u(x,s) -\frac{{(t - s)}^{2}} {2\varepsilon }.}$$

Note by (51) that

$$\displaystyle{\vert t - s\vert \leq m_{\varepsilon }.}$$

Let \((p,q) \in {D}^{+}u_{\varepsilon }(x,t)\) and choose a function \(\phi \in {C}^{1}(\overline{\Omega } \times (a,\,b))\) so that (x, t) = (p, q) and \(\max (u_{\varepsilon }-\phi ) = (u_{\varepsilon }-\phi )(x,t)\). Observe as in Sect. 1.6 that

$$\displaystyle{(p,(s - t)/\varepsilon ) \in {D}^{+}u(x,s)\ \ \text{ and }\ \ \frac{(t - s)} {\varepsilon } + q = 0.}$$

Hence,

$$\displaystyle{(p,q) \in {D}^{+}u(x,s).}$$

Therefore, we have

$$\displaystyle{ \left \{\begin{array}{llll} &q + H(x,p)+\delta \leq 0 &&\mathrm{if}\ \ x \in \Omega,\\ &\min \{q + H(x, p)+\delta,\,\gamma (x) \cdot p - g(x)\} \leq 0 & &\mathrm{if } \ \ x \in \partial \Omega. \end{array} \right. }$$
(52)

Moreover, we see that

$$\displaystyle{\vert q\vert = \frac{\vert t - s\vert } {\varepsilon } \leq \frac{m_{\varepsilon }} {\varepsilon },}$$

and

$$\displaystyle{H(x,p) \leq -q \leq \frac{m_{\varepsilon }} {\varepsilon } \ \ \mathrm{if}\ x \in \Omega.}$$

Hence, by the coercivity of H, we have

$$\displaystyle{ \vert q\vert + \vert p\vert \leq R(\varepsilon )\ \ \mathrm{if}\ x \in \Omega, }$$
(53)

for some constant \(R(\varepsilon ) > 0\).

Thus, we conclude from (52) that \(u_{\varepsilon }\) is a viscosity subsolution of

$$\displaystyle{\left \{\begin{array}{llll} &u_{t} + H(x,D_{x}u) \leq -\delta &&\mathrm{in}\ \ \overline{\Omega } \times (a + m_{\varepsilon },\,b - m_{\varepsilon }), \\ &\gamma \cdot D_{x}u \leq g &&\mathrm{on}\ \ \partial \Omega \times (a + m_{\varepsilon },\,b - m_{\varepsilon }), \end{array} \right.}$$

and from (53) that \(u_{\varepsilon }\) is Lipschitz continuous in \(\Omega \times (a + m_{\varepsilon },\,b - m_{\varepsilon })\). By Lemma 3.3, we have

$$\displaystyle{u_{\varepsilon }(x,t) =\limsup _{\Omega \times (a+m_{\varepsilon },\,b-m_{\varepsilon })\ni (y,s)\rightarrow (x,t)}u_{\varepsilon }(y,s)\ \ \ \text{ for all }\ (x,t) \in \partial \Omega \times (a + m_{\varepsilon },\,b - m_{\varepsilon }).}$$

Since \(u_{\varepsilon } \in \mathrm{Lip}(\Omega \times (a + m_{\varepsilon },\,b - m_{\varepsilon }))\), the limsup operation in the above formula can be replaced by the limit operation. Hence,

$$\displaystyle{u_{\varepsilon } \in C(\overline{\Omega } \times (a + m_{\varepsilon },\,b - m_{\varepsilon })),}$$

which guarantees that \(u_{\varepsilon }\) is Lipschitz continuous in \(\overline{\Omega } \times (a + m_{\varepsilon },\,b - m_{\varepsilon })\).

Finally, if we replace u and J by \(u_{\varepsilon }\) and \((a + 2m_{\varepsilon },\,b - 2m_{\varepsilon })\), respectively, and select \(\varepsilon > 0\) small enough so that

$$\displaystyle{\max _{\overline{\Omega }\times [a+2m_{\varepsilon },b-2m_{\varepsilon }]}(u_{\varepsilon } - v) > 0 >\max _{\overline{\Omega }\times \partial [a+2m_{\varepsilon },b-2m_{\varepsilon }]}(u_{\varepsilon } - v),}$$

then conditions (48)–(50) are satisfied. □ 

We may assume that there exist a constant δ > 0 and a finite open interval J ⊂ (0, ) such that

$$\displaystyle\begin{array}{rcl} \begin{array}{ll} &u\ \text{is a viscosity subsolution of} \\ &\left \{\begin{array}{llll} &u_{t}(x,t) + H(x,D_{x}u(x,t)) \leq -\delta &&\mathrm{in}\ \Omega \times J, \\ &\gamma (x) \cdot D_{x}u(x,t) \leq g(x)-\delta &&\mathrm{on}\ \partial \Omega \times J,\end{array} \right.\end{array}& &{}\end{array}$$
(54)
$$\displaystyle\begin{array}{rcl} \begin{array}{ll} &v\ \text{is a viscosity supersolution of} \\ &\left \{\begin{array}{llll} &v_{t}(x,t) + H(x,D_{x}v(x,t)) \geq \delta &&\mathrm{in}\ \Omega \times J, \\ &\gamma (x) \cdot D_{x}v(x,t) \geq g(x)+\delta &&\mathrm{on}\ \partial \Omega \times J,\end{array} \right.\end{array}& &{}\end{array}$$
(55)
$$\displaystyle\begin{array}{rcl} \max _{\overline{\Omega }\times \overline{J}}(u - v) > 0 >\max _{\overline{\Omega }\times \partial J}(u - v),& &{}\end{array}$$
(56)

and

$$\displaystyle{ u \in \mathrm{Lip}(\overline{\Omega } \times \overline{J})\ \text{and}\ v\ \text{is bounded on}\ \overline{\Omega } \times \overline{J}. }$$
(57)

Proof.

Let u, v, J be as in Reduction 2. Set J = (a, b). Let ρ be a defining function of Ω as before. Let \(0 <\varepsilon < 1\). We set

$$\displaystyle{u_{\varepsilon }(x,t) = u(x,t) -\varepsilon \rho (x)\ \text{ and }\ v_{\varepsilon }(x,t) = v(x,t) +\varepsilon \rho (x)\ \ \text{ for }\ (x,t) \in \overline{\Omega } \times \overline{J},}$$

and

$$\displaystyle{H_{\varepsilon }(x,p) = H(x,p -\varepsilon D\rho (x)) +\varepsilon \ \ \ \text{ for }\ (x,p) \in \overline{\Omega } \times {\mathbb{R}}^{n}.}$$

Let \((x,t) \in \overline{\Omega } \times J\) and \((p,q) \in {D}^{-}v_{\varepsilon }(x,t)\). Then we have

$$\displaystyle{(p -\varepsilon D\rho (x),q) \in {D}^{-}v(x,t).}$$

Since v is a viscosity supersolution of (ENP), if x ∈ Ω, then

$$\displaystyle{q + H(x,p -\varepsilon D\rho (x)) \geq 0.}$$

If x ∈ ∂Ω, then either

$$\displaystyle{q + H(x,p -\varepsilon D\rho (x)) \geq 0,}$$

or

$$\displaystyle\begin{array}{rcl} \gamma (x) \cdot p& =& \,\gamma (x) \cdot (p -\varepsilon D\rho (x)) +\varepsilon \gamma (x) \cdot D\rho (x) {}\\ & \geq & \,g(x) +\varepsilon \gamma (x) \cdot D\rho (x) \geq g(x)+\lambda \varepsilon, {}\\ \end{array}$$

where

$$\displaystyle{\lambda =\min _{\partial \Omega }\gamma \cdot D\rho \ (\,> 0\,).}$$

Now let \((p,q) \in {D}^{+}u_{\varepsilon }(x,t)\). Note that \((p +\varepsilon D\rho (x),q) \in {D}^{+}u(x,t)\). Since \(u \in \mathrm{Lip}(\overline{\Omega } \times [a,\,b])\), we have a bound C 0 > 0 such that

$$\displaystyle{\vert q\vert \leq C_{0}.}$$

If x ∈ Ω, then

$$\displaystyle\begin{array}{rcl} q + H(x,p -\varepsilon D\rho (x))& \leq & \,q + H(x,p +\varepsilon D\rho (x)) +\omega (2\varepsilon \vert D\rho (x)\vert ) {}\\ & \leq & \,-\delta +\omega (2\varepsilon C_{1}), {}\\ \end{array}$$

where

$$\displaystyle{C_{1} =\max _{\overline{\Omega }}\vert D\rho \vert,}$$

and ω denotes the modulus of continuity of H on the set \(\overline{\Omega } \times \overline{B}_{R+2C_{1}}\), with R > 0 being chosen so that

$$\displaystyle{\min _{\overline{\Omega }\times ({\mathbb{R}}^{n}\setminus B_{ R})}H > C_{0}.}$$

(Here we have used the fact that \(H(x,p +\varepsilon D\rho (x)) \leq C_{0}\), which implies that \(\vert p +\varepsilon D\rho (x)\vert \leq R\).)

If x ∈ ∂Ω, then either

$$\displaystyle{q + H(x,p -\varepsilon D\rho (x)) \leq -\delta +\omega (2\varepsilon C_{1}),}$$

or

$$\displaystyle{\gamma (x) \cdot p \leq \gamma (x) \cdot (p +\varepsilon D\rho (x)) -\varepsilon \gamma (x) \cdot D\rho (x) \leq g(x) -\lambda \varepsilon.}$$

Thus we see that \(v_{\varepsilon }\) is a viscosity supersolution of

$$\displaystyle{\left \{\begin{array}{llll} &v_{\varepsilon,t} + H_{\varepsilon }(x,D_{x}v_{\varepsilon }) \geq \varepsilon &&\mathrm{in}\ \Omega \times J, \\ &\gamma (x) \cdot D_{x}v_{\varepsilon }(x,t) \geq g(x)+\lambda \varepsilon &&\mathrm{on}\ \partial \Omega \times J, \end{array} \right.}$$

and \(u_{\varepsilon }\) is a viscosity subsolution of

$$\displaystyle{\left \{\begin{array}{llll} &u_{\varepsilon,t} + H_{\varepsilon }(x,D_{x}u_{\varepsilon }) \leq -\delta +\omega (2C_{1}\varepsilon )+\varepsilon &&\mathrm{in}\ \Omega \times J, \\ &\gamma \cdot \mathit{Du}_{\varepsilon } \leq g(x)-\lambda \varepsilon &&\mathrm{on}\ \partial \Omega \times J, \end{array} \right.}$$

If we replace u, v, H and δ by \(u_{\varepsilon }\), \(v_{\varepsilon }\), \(H_{\varepsilon }\) and

$$\displaystyle{\min \{\varepsilon,\lambda \varepsilon,\delta -\omega (2C_{1}\varepsilon )-\varepsilon \},}$$

respectively, and choose \(\varepsilon > 0\) sufficiently small, then conditions (54)–(57) are satisfied. □ 

Final step: Let u, v, J and δ be as in Reduction 3. We choose a maximum point \((z,\tau ) \in \overline{\Omega } \times \overline{J}\) of the function u − v. Note that τ ∈ J, that is, τ ∉ ∂J.

By replacing u, if necessary, by the function

$$\displaystyle{u(x,t) -\varepsilon \vert x - z{\vert }^{2} -\varepsilon {(t-\tau )}^{2},}$$

where \(\varepsilon > 0\) is a small constant, we may assume that (z, τ) is a strict maximum point of u − v.

By making a change of variables, we may assume that z = 0 and

$$\displaystyle{\Omega \cap B_{2r} =\{ x = (x_{1},\ldots,x_{n}) \in B_{2r}\, :\, x_{n} < 0\},}$$

while we may assume as well that [τ − r, τ + r] ⊂ J.

We set \(\hat{\gamma }=\gamma (0)\) and apply Lemma 3.4, with \(y = (0,\ldots,0,1) \in {\mathbb{R}}^{n}\) and \(z =\hat{\gamma }\), to find a quadratic function ζ so that

$$\displaystyle{\left \{\begin{array}{llll} &\zeta (t\xi ) = {t}^{2}\zeta (\xi ) &&\text{ for all }\ (\xi,t) \in {\mathbb{R}}^{n} \times \mathbb{R}, \\ &\zeta (\xi ) > 0 &&\mathrm{if}\ \ \xi \not =0, \\ &\hat{\gamma } \cdot D\zeta (\xi ) = 2\hat{\gamma }_{n}\xi _{n}&&\text{ for all }\ \xi = (\xi _{1},\ldots,\xi _{n}) \in {\mathbb{R}}^{n}, \end{array} \right.}$$

where \(\hat{\gamma }_{n}\) denotes the n-th component of the n-tuple \(\hat{\gamma }\).

By replacing ζ by a constant multiple of ζ, we may assume that

$$\displaystyle\begin{array}{rcl} & & \zeta (\xi ) \geq \vert \xi {\vert }^{2}\qquad \qquad \qquad \qquad \text{ for all }\xi \in {\mathbb{R}}^{n}, {}\\ & & \vert D\zeta (\xi )\vert \leq C_{0}\vert \xi \vert \qquad \qquad \qquad \text{ for all }\xi \in {\mathbb{R}}^{n}, {}\\ & & {}\\ \end{array}$$
$$\displaystyle\begin{array}{rcl} \hat{\gamma }\cdot D\zeta (\xi )\left \{\begin{array}{@{}l@{\quad }l@{}} \geq 0\quad &\mathrm{if}\ \xi _{n} \geq 0, \\ \leq 0\quad &\mathrm{if}\ \xi _{n} \leq 0, \end{array} \right.& & {}\\ \end{array}$$

where C 0 > 0 is a constant.

Let M > 0 be a Lipschitz bound of the function u. Set

$$\displaystyle{\hat{g} = g(0),\ \ \ \mu =\hat{ g} \frac{\hat{\gamma }} {\vert \hat{\gamma }{\vert }^{2}}\ \ \text{ and }\ \ M_{1} = M + \vert \mu \vert.}$$

We may assume by replacing r by a smaller positive constant if needed that for all  x ∈ B r  ∩ ∂Ω,

$$\displaystyle{ \vert \gamma (x) -\hat{\gamma }\vert < \frac{\delta } {2(\vert \mu \vert + C_{0}M_{1})}\ \ \ \text{ and }\ \ \ \vert g(x) -\hat{ g}\vert < \frac{\delta } {2}. }$$
(58)

For α > 1 we consider the function

$$\displaystyle{\Phi (x,t,y,s) = u(x,t) - v(y,s) -\mu \cdot (x - y) -\alpha \zeta (x - y) -\alpha {(t - s)}^{2}}$$

on \(K :={ \left ((\overline{\Omega } \cap \overline{B}_{r}(0,\tau ) \times [\tau -r,\,\tau +r]\right )}^{2}\). Let \((x_{\alpha },t_{\alpha },y_{\alpha },s_{\alpha })\) be a maximum point of the function Φ. By the inequality \(\Phi (y_{\alpha },s_{\alpha },y_{\alpha },s_{\alpha }) \leq \Phi (x_{\alpha },t_{\alpha },y_{\alpha },s_{\alpha })\), we get

$$\displaystyle\begin{array}{rcl} \alpha (\vert x_{\alpha } - y_{\alpha }{\vert }^{2} + {(t_{\alpha } - s_{\alpha })}^{2})& \leq & \,\alpha (\zeta (x_{\alpha } - y_{\alpha }) + {(t_{\alpha } - s_{\alpha })}^{2}) {}\\ & \leq & \,u(x_{\alpha },t_{\alpha }) - u(y_{\alpha },s_{\alpha }) + \vert \mu \vert \vert x_{\alpha } - y_{\alpha }\vert {}\\ &\leq & \,M_{1}{(\vert x_{\alpha } - y_{\alpha }{\vert }^{2} + \vert t_{\alpha } - s_{\alpha }{\vert }^{2})}^{1/2}, {}\\ \end{array}$$

and hence

$$\displaystyle{ \alpha {(\vert x_{\alpha } - y_{\alpha }{\vert }^{2} + \vert t_{\alpha } - s_{\alpha }{\vert }^{2})}^{1/2} \leq M_{ 1}. }$$
(59)

As usual we may deduce that as α → ,

$$\displaystyle{\left \{\begin{array}{ll} &(x_{\alpha },\tau _{\alpha }),\,(y_{\alpha },s_{\alpha }) \rightarrow (0,\tau ), \\ &u(x_{\alpha },t_{\alpha }) \rightarrow u(0,\tau ), \\ &v(y_{\alpha },s_{\alpha }) \rightarrow v(0,\tau ).\end{array} \right.}$$

Let α > 1 be so large that

$$\displaystyle{(x_{\alpha },t_{\alpha }),(y_{\alpha },s_{\alpha }) \in (\overline{\Omega } \cap B_{r}) \times (\tau -r,\,\tau +r).}$$

Accordingly, we have

$$\displaystyle\begin{array}{rcl} & & (\mu +\alpha D\zeta (x_{\alpha } - y_{\alpha }),\,2\alpha (t_{\alpha } - s_{\alpha })) \in {D}^{+}u(x_{\alpha },t_{\alpha }), {}\\ & & (\mu +\alpha D\zeta (x_{\alpha } - y_{\alpha }),\,2\alpha (t_{\alpha } - s_{\alpha })) \in {D}^{-}v(y_{\alpha },s_{\alpha }). {}\\ \end{array}$$

Using (59), we have

$$\displaystyle{ \alpha \vert D\zeta (x_{\alpha } - y_{\alpha })\vert \leq C_{0}\alpha \vert x_{\alpha } - y_{\alpha }\vert \leq C_{0}M_{1}. }$$
(60)

If x α  ∈ ∂Ω, then x α, n  = 0 and \((x_{\alpha } - y_{\alpha })_{n} \geq 0\). Hence, in this case, we have

$$\displaystyle{\hat{\gamma }\cdot D\zeta (x_{\alpha } - y_{\alpha }) \geq 0,}$$

and moreover, in view of (58) and (60),

$$\displaystyle\begin{array}{rcl} \gamma (x_{\alpha }) \cdot (\mu +\alpha D\zeta (x_{\alpha } - y_{\alpha })) \geq \,& \hat{\gamma } &\cdot (\mu +\alpha D\zeta (x_{\alpha } - y_{\alpha })) {}\\ \,& -& \vert \gamma (x_{\alpha }) -\hat{\gamma }\vert (\vert \mu \vert + C_{0}M_{1}) {}\\ & >& \,g(x_{\alpha }) -\vert \hat{g} - g(x_{\alpha })\vert - \frac{\delta } {2} > g(x_{\alpha }) -\delta. {}\\ \end{array}$$

Now, by the viscosity property of u, we obtain

$$\displaystyle{2\alpha (t_{\alpha } - s_{\alpha }) + H(x_{\alpha },\mu +\alpha D\zeta (x_{\alpha } - y_{\alpha })) \leq -\delta,}$$

which we certainly have when x α  ∈ Ω.

If y α  ∈ ∂Ω, then \((x_{\alpha } - y_{\alpha })_{n} \leq 0\) and

$$\displaystyle{\hat{\gamma }\cdot D\zeta (x_{\alpha } - y_{\alpha }) \leq 0.}$$

As above, we find that if y α  ∈ ∂Ω, then

$$\displaystyle{\gamma (y_{\alpha }) \cdot (\mu +\alpha D\zeta (x_{\alpha } - y_{\alpha })) <\delta,}$$

and hence, by the viscosity property of v,

$$\displaystyle{2(t_{\alpha } - s_{\alpha }) + H(y_{\alpha },\mu +\alpha D\zeta (x_{\alpha } - y_{\alpha })) \geq \delta,}$$

which is also valid in case when y α  ∈ Ω.

Thus, we always have

$$\displaystyle{\left \{\begin{array}{ll} &2\alpha (t_{\alpha } - s_{\alpha }) + H(x_{\alpha },\mu +\alpha D\zeta (x_{\alpha } - y_{\alpha })) \leq -\delta,\\ &2(t_{\alpha } - s_{\alpha }) + H(y_{\alpha },\mu +\alpha D\zeta (x_{\alpha } - y_{\alpha })) \geq \delta. \end{array} \right.}$$

Sending α →  along a sequence, we obtain

$$\displaystyle{q + H(0,\mu +p) \leq -\delta \ \ \ \text{ and }\ \ \ q + H(0,\mu +p) \geq \delta }$$

for some \(p \in \overline{B}_{C_{0}M_{1}}\) and q ∈ [ − 2M 1, 2M 1], which is a contradiction. This completes the proof of Theorem 3.1. □ 

4 Stationary Problem: Weak KAM Aspects

In this section we discuss some aspects of weak KAM theory for Hamilton–Jacobi equations with the Neumann boundary condition. We refer to Fathi [25, 27], E [22] and Evans [24] for origins and developments of weak KAM theory.

Throughout this section we assume that (A1)–(A6) and the following (A7) hold:

  1. (A7)

    The Hamiltonian H is convex. That is, the function pH(x, p) is convex in \({\mathbb{R}}^{n}\) for any \(x \in \overline{\Omega }\).

As in Sect. 2 we consider the stationary problem

$$\displaystyle{ \left \{\begin{array}{llll} &H(x,\mathit{Du}(x)) = 0\ \ \ &&\mathrm{in}\ \Omega, \\ &\frac{\partial u} {\partial \gamma } (x) = g(x)\ \ \ &&\mathrm{on}\ \partial \Omega.\end{array} \right. }$$
(SNP)

As remarked before this boundary value problem may have no solution in general, but, due to Theorem 3.4, if we replace H by H − a with the right choice of \(a \in \mathbb{R}\), the problem (SNP) has a viscosity solution. Furthermore, if we replace H by H − a with a sufficiently large \(a \in \mathbb{R}\), the problem (SNP) has a viscosity subsolution. With a change of Hamiltonians of this kind in mind, we make the following hypothesis throughout this section:

  1. (A8)

    The problem (SNP) has a viscosity subsolution.

4.1 Aubry Sets and Representation of Solutions

We start this subsection by the following Lemma.

Lemma 4.1.

Let \(u \in \mathrm{USC}(\overline{\Omega })\) be a viscosity subsolution of (SNP) . Then \(u \in \mathrm{Lip}(\overline{\Omega })\) . Moreover, u has a Lipschitz bound which depends only on H and Ω.

Proof.

By the coercivity of H, there exists a constant M > 0 such that H(x, p) > 0 for all \((x,p) \in \overline{\Omega } \times ({\mathbb{R}}^{n} \setminus B_{M})\). Fix such a constant M > 0 and note that u is a viscosity subsolution of | Du(x) | ≤ M in Ω. Accordingly, we see by Lemma 2.2 that u ∈ Lip(Ω). Furthermore, if C > 0 is the constant from Lemma 2.1, then we have | u(x) − u(y) | ≤ CM | x − y | for all x, y ∈ Ω. (See also Appendix A.3.)

Since the function u(x), as a function of (x, t), is a viscosity subsolution of (ENP), Lemma 3.3 guarantees that u is continuous up to the boundary ∂Ω. Thus, we get | u(x) − u(y) | ≤ CM | x − y | for all \(x,y \in \overline{\Omega }\), which completes the proof. □ 

We introduce the distance-like function \(d\, :\, \overline{\Omega } \times \overline{\Omega } \rightarrow \mathbb{R}\) by

$$\displaystyle{d(x,y) =\sup \{ v(x) - v(y)\, :\, v \in \mathrm{USC}(\overline{\Omega }) \cap {\mathcal{S}}^{-}\},}$$

where \({\mathcal{S}}^{-} = {\mathcal{S}}^{-}(\overline{\Omega })\) has been defined as the set of all viscosity subsolutions of (SNP). By (A8), we have \({\mathcal{S}}^{-}\not =\varnothing \) and hence d(x, x) = 0 for all \(x \in \overline{\Omega }\). Since \(\mathrm{USC}(\overline{\Omega }) \cap {\mathcal{S}}^{-}\) is equi-Lipschitz continuous on \(\overline{\Omega }\) by Lemma 4.1, we see that the functions (x, y)↦v(x) − v(y), with \(v \in \mathrm{USC}(\overline{\Omega }) \cap {\mathcal{S}}^{-}\), are equi-Lipschitz continuous and d is Lipschitz continuous on \(\overline{\Omega } \times \overline{\Omega }\). By Proposition 1.10, the functions xd(x, y), with \(y \in \overline{\Omega }\), are viscosity subsolutions of (SNP). Hence, by the definition of d(x, z) we get

$$\displaystyle{d(x,y) - d(z,y) \leq d(x,z)\ \ \ \text{ for all }x,y,z \in \overline{\Omega }.}$$

We set

$$\displaystyle{\mathcal{F}_{y} =\{ v(x) - v(y)\, :\, v \in {\mathcal{S}}^{-}\},\ \ \text{ with }\ y \in \overline{\Omega },}$$

and observe by using Proposition 1.10 and Lemma 1.3 that \(\mathcal{F}_{y}\) satisfies (P1) and (P2), with Ω replaced by \(\overline{\Omega } \setminus \{ y\}\), of Proposition 1.13. Hence, by Proposition 1.13, the function \(d(\cdot,y) =\sup \mathcal{F}_{y}\) is a viscosity solution of (SNP) in \(\overline{\Omega } \setminus \{ y\}\).

The following proposition collects these observations.

Proposition 4.1.

We have:

  1. (i)

    d(x,x) = 0   for all   \(x \in \overline{\Omega }\).

  2. (ii)

    d(x,y) ≤ d(x,z) + d(z,y)   for all   \(x,y,z \in \overline{\Omega }\).

  3. (iii)

    \(d(\cdot,\,y) \in {\mathcal{S}}^{-}(\overline{\Omega })\)   for all   \(y \in \overline{\Omega }\).

  4. (iv)

    \(d(\cdot,\,y) \in \mathcal{S}(\overline{\Omega } \setminus \{ y\})\)   for all   \(y \in \overline{\Omega }\).

The Aubry set (or Aubry–Mather set) \(\mathcal{A}\) associated with (SNP) is defined by

$$\displaystyle{\mathcal{A} =\{ y \in \overline{\Omega }\, :\, d(\cdot,\,y) \in \mathcal{S}(\overline{\Omega })\}.}$$

Example 4.1.

Let n = 1, Ω = ( − 1, 1), H(x, p) =  | p | − f(x), f(x) = 1 − | x | , γ( ± 1) =  ± 1 and g( ± 1) = 0. The function v ∈ C 1([ − 1, 1]) given by

$$\displaystyle{v(x) = \left \{\begin{array}{@{}l@{\quad }l@{}} 1 -\frac{1} {2}{(x + 1)}^{2}\quad &\mathrm{if}\ x \leq 0, \\ \frac{1} {2}{(x - 1)}^{2} \quad &\mathrm{if}\ x \geq 0 \end{array} \right.}$$

is a classical solution of (SNP). We show that d(x, 1) = v(x) for all x ∈ [ − 1, 1]. It is enough to show that d(x, 1) ≤ v(x) for all x ∈ [ − 1, 1]. To prove this, we suppose by contradiction that max x ∈ [ − 1, 1](d(x, 1) − v(x)) > 0. We may choose a constant \(\varepsilon > 0\) so small that \(\max _{x\in [-1,1]}(d(x,1) - v(x) -\varepsilon (1 - x)) > 0\). Let \(x_{\varepsilon } \in [-1,\,1]\) be a maximum point of the function \(d(x,1) - v(x) -\varepsilon (1 - x)\). Since this function vanishes at x = 1, we have \(x_{\varepsilon } \in [-1,\,1)\). If \(x_{\varepsilon } > -1\), then we have

$$\displaystyle{0 \geq H(x_{\varepsilon },v^{\prime}(x_{\varepsilon })-\varepsilon ) = \vert v^{\prime}(x_{\varepsilon })\vert +\varepsilon -f(x_{\varepsilon }) =\varepsilon > 0,}$$

which is impossible. Here we have used the fact that v′(x) =  | x | − 1 ≤ 0 for all x ∈ [ − 1, 1]. On the other hand, if \(x_{\varepsilon } = -1\), then we have

$$\displaystyle{0 \geq \min \{ H(-1,v^{\prime}(-1)-\varepsilon ),\,-(v^{\prime}(-1)-\varepsilon )\} =\min \{\varepsilon,\,\varepsilon \}=\varepsilon > 0,}$$

which is again impossible. Thus we get a contradiction. That is, we have d(x, 1) ≤ v(x) and hence d(x, 1) = v(x) for all x ∈ [ − 1, 1]. Arguments similar to the above show moreover that

$$\displaystyle{d(x,-1) = \left \{\begin{array}{@{}l@{\quad }l@{}} \frac{1} {2}{(x + 1)}^{2} \quad &\mathrm{if}\ x \leq 0, \\ 1 -\frac{1} {2}{(x - 1)}^{2}\quad &\mathrm{if}\ x \geq 0, \end{array} \right.}$$

and

$$\displaystyle{d(x,y) = \left \{\begin{array}{@{}l@{\quad }l@{}} d(x,1) - d(y,1) \quad &\mathrm{if}\ x \leq y,\\ d(x, -1) - d(y, -1)\quad &\mathrm{if } \ x \geq y. \end{array} \right.}$$

Since two functions d(x, ± 1) are classical solutions of (SNP), we see that \(\pm 1 \in \mathcal{A}\). Noting that d(x, y) ≥ 0 and d(x, x) = 0 for all x, y ∈ [ − 1, 1], we find that for each fixed y ∈ [ − 1, 1] the function xd(x, y) has a minimum at x = y. If y ∈ ( − 1, 1), then H(y, 0) =  − f(y) < 0. Hence, we see that the interval ( − 1, 1) does not intersect \(\mathcal{A}\). Thus, we conclude that \(\mathcal{A} =\{ -1,\,1\}\).

A basic observation on \(\mathcal{A}\) is the following:

Proposition 4.2.

The Aubry set \(\mathcal{A}\) is compact.

Proof.

It is enough to show that \(\mathcal{A}\) is a closed subset of \(\overline{\Omega }\). Note that the function d is Lipschitz continuous in \(\overline{\Omega } \times \overline{\Omega }\). Therefore, if \(\{y_{k}\}_{k\in \mathbb{N}} \subset \mathcal{A}\) converges to \(y \in \overline{\Omega }\), then the sequence \(\{d(\cdot,\,y_{k})\}_{k\in \mathbb{N}}\) converges to the function d( ⋅, y) in \(C(\overline{\Omega })\). By the stability of the viscosity property under the uniform convergence, we see that \(d(\cdot,\,y) \in \mathcal{S}\). Hence, we have \(y \in \mathcal{A}\). □ 

The main assertion in this section is the following and will be proved at the end of the section.

Theorem 4.1.

Let \(u \in C(\overline{\Omega })\) be a viscosity solution of (SNP) . Then

$$\displaystyle{ u(x) =\inf \{ u(y) + d(x,y)\, :\, y \in \mathcal{A}\}\ \ \ \text{ for all }x \in \overline{\Omega }. }$$
(61)

We state the following approximation result on viscosity subsolutions of (SNP).

Theorem 4.2.

Let \(u \in C(\overline{\Omega })\) be a viscosity subsolution of (SNP) . There exists a collection \(\{{u}^{\varepsilon }\}_{\varepsilon \in (0,\,1)} \subset {C}^{1}(\overline{\Omega })\) such that for any \(\varepsilon \in (0,\,1)\) ,

$$\displaystyle{\left \{\begin{array}{llll} &H(x,{\mathit{Du}}^{\varepsilon }(x)) \leq \varepsilon &&\mathrm{in}\ \Omega, \\ &\frac{\partial {u}^{\varepsilon }} {\partial \gamma } (x) \leq g(x) &&\mathrm{on}\ \partial \Omega,\end{array} \right.}$$

and

$$\displaystyle{\|{u}^{\varepsilon } - u\|_{\infty,\Omega } <\varepsilon.}$$

A localized version of the above theorem is in [39] (see also Appendix A.4 and [8]) and the above theorem seems to be new in the global nature.

As a corollary, we get the following theorem.

Theorem 4.3.

Let \(f_{1},f_{2} \in C(\overline{\Omega })\) and g 1 ,g 2 ∈ C(∂Ω). Let \(u,v \in C(\overline{\Omega })\) be viscosity solutions of

$$\displaystyle{\left \{\begin{array}{llll} &H(x,\mathit{Du}) \leq f_{1} & & \mathrm{in}\ \Omega, \\ &\frac{\partial u} {\partial \gamma } \leq g_{1} & & \mathrm{on}\ \partial \Omega, \end{array} \right.}$$

and

$$\displaystyle{\left \{\begin{array}{llll} &H(x,Dv) \leq f_{2} & & \mathrm{in}\ \Omega, \\ &\frac{\partial v} {\partial \gamma } \leq g_{2} & & \mathrm{on}\ \partial \Omega,\end{array} \right.}$$

respectively. Let 0 < λ < 1 and set w = (1 −λ)u + λv. Then w is a viscosity subsolution of

$$\displaystyle{ \left \{\begin{array}{llll} &H(x,Dw) \leq (1-\lambda )f_{1} +\lambda f_{2} & & \mathrm{in}\ \Omega, \\ &\frac{\partial w} {\partial \gamma } \leq (1-\lambda )g_{1} +\lambda g_{2} & & \mathrm{on}\ \partial \Omega, \end{array} \right. }$$
(62)

Proof.

By Theorem 4.2, for each \(\varepsilon \in (0,\,1)\) there are functions \({u}^{\varepsilon },\,{v}^{\varepsilon } \in {C}^{1}(\overline{\Omega })\) such that

$$\displaystyle\begin{array}{rcl} \|{u}^{\varepsilon } - u\|_{\infty,\Omega } +\| {v}^{\varepsilon } - v\|_{\infty,\Omega } <\varepsilon,& & {}\\ \left \{\begin{array}{llll} &H(x,{\mathit{Du}}^{\varepsilon }(x)) \leq f_{1}(x)+\varepsilon &&\mathrm{in}\ \overline{\Omega }, \\ &\frac{\partial {u}^{\varepsilon }} {\partial \gamma } (x) \leq g_{1}(x) &&\mathrm{on}\ \partial \Omega,\end{array} \right.& & {}\\ \end{array}$$

and

$$\displaystyle{\left \{\begin{array}{llll} &H(x,D{v}^{\varepsilon }(x)) \leq f_{2}(x)+\varepsilon &&\mathrm{in}\ \overline{\Omega }, \\ &\frac{\partial {v}^{\varepsilon }} {\partial \gamma } (x) \leq g_{2}(x) &&\mathrm{on}\ \partial \Omega. \end{array} \right.}$$

If we set \({w}^{\varepsilon } = (1-\lambda ){u}^{\varepsilon } +\lambda {v}^{\varepsilon }\), then we get with use of (A7)

$$\displaystyle{\left \{\begin{array}{llll} &H(x,D{w}^{\varepsilon }(x)) \leq (1-\lambda )f_{1}(x) +\lambda f_{2}(x)+\varepsilon &&\mathrm{in}\ \overline{\Omega }, \\ &\frac{\partial {w}^{\varepsilon }} {\partial \gamma } (x) \leq (1-\lambda )g_{1}(x) +\lambda g_{2}(x) &&\mathrm{on}\ \partial \Omega. \end{array} \right.}$$

Thus, in view of the stability property (Proposition 1.9), we see in the limit as \(\varepsilon \rightarrow 0\) that w is a viscosity subsolution of (62). □ 

The following theorem is also a consequence of (A7), the convexity of H, and Theorem 4.2.

Theorem 4.4.

Let \(\mathcal{F}\subset \) USC( \(\overline{\Omega }\) ) be a nonempty collection of viscosity subsolutions of (SNP) . Assume that \(u(x) :=\inf \mathcal{F}(x) > -\infty \) for all \(x \in \overline{\Omega }\) . Then \(u \in \mathrm{Lip}(\overline{\Omega })\) and it is a viscosity subsolution of (SNP).

This theorem may be regarded as part of the theory of Barron–Jensen’s lower semicontinuous viscosity solutions. There are at least two approaches to this theory: the original one by Barron–Jensen [11] and the other due to Barles [5]. The following proof is close to Barles’ approach.

Proof.

By Lemma 4.1, the collection \(\mathcal{F}\) is equi-Lipschitz in \(\overline{\Omega }\). Hence, u is a Lipschitz continuous function in \(\overline{\Omega }\). For each \(x \in \overline{\Omega }\) there is a sequence \(\{u_{x,k}\}_{k\in \mathbb{N}} \subset \mathcal{F}\) such that \(\lim _{k\rightarrow \infty }u_{x,k}(x) = u(x).\) Fix such sequences \(\{u_{x,k}\}_{k\in \mathbb{N}}\), with \(x \in \overline{\Omega }\) and select a countable dense subset \(Y \subset \overline{\Omega }\). Observe that \(Y \times \mathbb{N}\) is a countable set and

$$\displaystyle{u(x) =\inf \{ u_{y,k}(x)\, :\, (y,k) \in Y \times \mathbb{N}\}\ \ \ \text{ for all }\ x \in \overline{\Omega }.}$$

Thus we may assume that \(\mathcal{F}\) is a sequence.

Let \(\mathcal{F} =\{ u_{k}\}_{k\in \mathbb{N}}\). Then we have

$$\displaystyle{u(x) =\lim _{k\rightarrow \infty }(u_{1} \wedge u_{2} \wedge \cdots \wedge u_{k})(x)\ \ \ \text{ for all }\ x \in \overline{\Omega }.}$$

We show that \(u_{1} \wedge u_{2} \wedge \cdots \wedge u_{k}\) is a viscosity subsolution of (SNP) for every \(k \in \mathbb{N}\)‘. It is enough to show that if v and w are viscosity subsolutions of (SNP), then so is the function v ∧ w.

Let v and w be viscosity subsolutions of (SNP). Fix any \(\varepsilon > 0\). In view of Theorem 4.2, we may select functions \(v_{\varepsilon },\,w_{\varepsilon } \in {C}^{1}(\overline{\Omega })\) so that both for \((\phi _{\varepsilon },\phi ) = (v_{\varepsilon },v)\) and \((\phi _{\varepsilon },\phi ) = (w_{\varepsilon },w)\), we have

$$\displaystyle{\left \{\begin{array}{llll} &H(x,D\phi _{\varepsilon }(x)) \leq \varepsilon &&\text{ for all }\ x \in \Omega, \\ &\frac{\partial \phi _{\varepsilon }} {\partial \gamma }(x) \leq g(x) &&\text{ for all }\ x \in \partial \Omega,\\ &\|\phi _{ \varepsilon } -\phi \|_{\infty,\Omega } <\varepsilon. \end{array} \right.}$$

Note that \((v_{\varepsilon } \wedge w_{\varepsilon })(x) = v_{\varepsilon }(x) - (v_{\varepsilon } - w_{\varepsilon })_{+}(x)\). Let \(\{\eta _{k}\}_{k\in \mathbb{N}} \subset {C}^{1}(\mathbb{R})\) be such that

$$\displaystyle{\left \{\begin{array}{llll} &\eta _{k}(r) \rightarrow r_{+} & & \text{ uniformly on} \ \mathbb{R} \ as \ k \rightarrow \infty, \\ &0 \leq \eta _{k}^{\prime}(r) \leq 1&&\text{ for all }\ r \in \mathbb{R},\,k \in \mathbb{N}. \end{array} \right.}$$

We set \(z_{\varepsilon,k} = v_{\varepsilon } -\eta _{k} \circ (v_{\varepsilon } - w_{\varepsilon })\) and observe that

$$\displaystyle{Dz_{\varepsilon,k}(x) = \left (1 -\eta _{k}^{\prime}(v_{\varepsilon }(x) - w_{\varepsilon }(x))\right )Dv_{\varepsilon }(x) +\eta _{k}^{\prime}(v_{\varepsilon }(x) - w_{\varepsilon }(x))Dw_{\varepsilon }(x).}$$

By the convexity of H, we see easily that \(z_{\varepsilon,k}\) satisfies

$$\displaystyle{\left \{\begin{array}{llll} &H(x,Dz_{\varepsilon,k}(x)) \leq \varepsilon &&\text{ for all }\ x \in \Omega, \\ &\frac{\partial z_{\varepsilon,k}} {\partial \gamma } (x) \leq g(x) &&\text{ for all }\ x \in \partial \Omega. \end{array} \right.}$$

Since v ∧ w is a uniform limit of \(z_{\varepsilon,k}\) in \(\overline{\Omega }\) as \(k \rightarrow \infty \) and \(\varepsilon \rightarrow 0\), we see that v ∧ w is a viscosity subsolution of (SNP).

By the Ascoli–Arzela theorem or Dini’s lemma, we deduce that the convergence

$$\displaystyle{u(x) =\lim _{k\rightarrow \infty }(u_{1} \wedge \cdots \wedge u_{k})(x)}$$

is uniform in \(\overline{\Omega }\). Thus we conclude that u is a viscosity subsolution of (SNP). □ 

Remark 4.1.

Theorem 4.2 has its localized version which concerns viscosity subsolutions of

$$\displaystyle{\left \{\begin{array}{llll} &H(x,\mathit{Du}(x)) \leq 0&&\mathrm{in}\ U \cap \Omega, \\ &\frac{\partial u} {\partial \gamma } (x) \leq g(x) &&\mathrm{on}\ U \cap \partial \Omega, \end{array} \right.}$$

where U is an open subset of \({\mathbb{R}}^{n}\) having nonempty intersection with Ω. More importantly, it has a version for the Neumann problem for Hamilton–Jacobi equations of evolution type, which concerns solutions of

$$\displaystyle{\left \{\begin{array}{llll} &u_{t}(x,t) + H(x,D_{x}u(x,t)) \leq 0&&\mathrm{in}\ U \cap (\Omega \times \mathbb{R}_{+}), \\ &\frac{\partial u} {\partial \gamma } (x,t) \leq g(x) &&\mathrm{on}\ U \cap (\partial \Omega \times \mathbb{R}_{+}), \end{array} \right.}$$

where U is an open subset of \({\mathbb{R}}^{n} \times \mathbb{R}_{+}\), with \(U \cap (\Omega \times \mathbb{R}_{+})\not =\varnothing \). Consequently, Theorems 4.3 and 4.4 are valid for these problems with trivial modifications. For these, see Appendix A.4.

Theorem 4.5.

We have

$$\displaystyle{{c}^{\#} =\inf \left \{\max _{ x\in \overline{\Omega }}H(x,D\psi (x))\, :\,\psi \in {C}^{1}(\overline{\Omega }),\ \partial \psi /\partial \gamma \leq g\mathrm{on}\ \ \partial \Omega \right \}.}$$

Remark 4.2.

A natural question here is if there is a function \(\psi \in {C}^{1}(\overline{\Omega })\) which attains the infimum in the above formula. See [12, 28].

Proof.

Let c  ⋆  denote the right hand side of the above minimax formula. By the definition of c  ⋆ , it is clear that for any a > c  ⋆ , there is a classical subsolution of (EVP). Hence, by Corollary 3.2, we see that c # ≤ c  ⋆ .

On the other hand, by Theorem 3.4, there is a viscosity solution v of (EVP), with a = c #. By Theorem 4.2, for any a > c # there is a classical subsolution of (EVP). That is, we have c  ⋆  ≤ c #. Thus we conclude that c # = c  ⋆ . □ 

Theorem 4.6 (Comparison). 

Let \(v,w \in C(\overline{\Omega })\) be a viscosity subsolution and supersolution of (SNP) , respectively. Assume that v ≤ w on \(\mathcal{A}\) . Then v ≤ w in  \(\overline{\Omega }\).

For the proof of the above theorem, we need the following lemma.

Lemma 4.2.

Let K be a compact subset of \(\overline{\Omega } \setminus \mathcal{A}\) . Then there exists a function \(\psi \in {C}^{1}(U \cap \overline{\Omega })\) , where U is an open neighborhood of K in \({\mathbb{R}}^{n}\) , and a positive constant δ > 0 such that

$$\displaystyle{ \left \{\begin{array}{llll} &H(x,D\psi (x)) \leq -\delta \ \ \ &&\mathrm{in}\ U \cap \Omega, \\ &\frac{\partial \psi } {\partial \gamma }(x) \leq g(x)-\delta \ \ \ &&\mathrm{on}\ U \cap \partial \Omega. \end{array} \right. }$$
(63)

We assume temporarily the validity of the above lemma and complete the proof of Theorems 4.6 and 4.1. The proof of the above lemma will be given in the sequel.

Proof (Theorem 4.6). 

By contradiction, we suppose that \(M :=\sup _{\overline{\Omega }}(v - w) > 0\). Let

$$\displaystyle{K =\{ x \in \overline{\Omega }\, :\, (v - w)(x) = M\},}$$

which is a compact subset of \(\overline{\Omega } \setminus \mathcal{A}\). According to Lemma 4.2, there are δ > 0 and \(\psi \in {C}^{1}(U \cap \overline{\Omega })\), where U is an open neighborhood of K such that ψ is a subsolution of (63).

According to Theorem 4.2, for each \(\varepsilon \in (0,\,1)\) there is a function \({v}^{\varepsilon } \in {C}^{1}(\overline{\Omega })\) such that

$$\displaystyle{\left \{\begin{array}{llll} &H(x,D{v}^{\varepsilon }(x)) \leq \varepsilon \ &&\mathrm{in}\ \ \Omega, \\ &\frac{\partial {v}^{\varepsilon }} {\partial \gamma } (x) \leq g(x) &&\mathrm{on}\ \ \partial \Omega,\end{array} \right.}$$

and

$$\displaystyle{\|{v}^{\varepsilon } - v\|_{\infty,\Omega } <\varepsilon.}$$

We fix a λ ∈ (0, 1) so that \(\delta _{\varepsilon } := -(1-\lambda )\varepsilon +\delta \lambda > 0\) and set

$$\displaystyle{u_{\varepsilon }(x) = (1-\lambda ){v}^{\varepsilon }(x) +\lambda \psi (x).}$$

This function satisfies

$$\displaystyle{\left \{\begin{array}{llll} &H(x,\mathit{Du}_{\varepsilon }(x)) \leq -\delta _{\varepsilon }\ &&\mathrm{in}\ \ U \cap \Omega, \\ &\frac{\partial u_{\varepsilon }} {\partial \gamma } (x) \leq g(x) -\delta _{\varepsilon } &&\mathrm{on}\ \ U \cap \partial \Omega. \end{array} \right.}$$

This contradicts the viscosity property of the function w if \(u_{\varepsilon } - w\) attains a maximum at a point \(z \in U \cap \overline{\Omega }\). Hence, we have

$$\displaystyle{\max _{\overline{U}\cap \overline{\Omega }}(u_{\varepsilon } - w) =\max _{\partial U\cap \overline{\Omega }}(u_{\varepsilon } - w).}$$

Sending \(\varepsilon \rightarrow 0\) and then λ → 0 yields

$$\displaystyle{\max _{\overline{U}\cap \overline{\Omega }}(v - w) =\max _{\partial U\cap \overline{\Omega }}(v - w),}$$

that is,

$$\displaystyle{M =\max _{\partial U\cap \overline{\Omega }}(v - w).}$$

This is a contradiction. □ 

Remark 4.3.

Obviously, the continuity assumption on v, w in the above lemma can be replaced by the assumption that \(v \in \mathrm{USC}(\overline{\Omega })\) and \(w \in \mathrm{LSC}(\overline{\Omega })\).

Proof (Theorem 4.1). 

We write w(x) for the right hand side of (61) in this proof. By the definition of d, we have

$$\displaystyle{u(x) - u(y) \leq d(x,y)\ \ \ \text{ for all }x,y \in \overline{\Omega },}$$

from which we see that u(x) ≤ w(x).

By the definition of w, for every \(x \in \mathcal{A}\), we have

$$\displaystyle{w(x) \leq u(x) + d(x,x) = u(x).}$$

Hence, we have w = u on \(\mathcal{A}\).

Now, by Proposition 1.10 (its version for supersolutions), we see that w is a viscosity supersolution of (SNP) while Theorem 4.4 guarantees that w is a viscosity subsolution of (SNP). We invoke here Theorem 4.6, to see that u = w in \(\overline{\Omega }\). □ 

Proof (Lemma 4.2). 

In view of Theorem 4.2, it is enough to show that there exist functions \(w \in \mathrm{Lip}(\overline{\Omega })\) and \(f \in C(\overline{\Omega })\) such that

$$\displaystyle{\left \{\begin{array}{@{}l@{\quad }l@{}} f(x) \geq 0\quad &\mathrm{in}\ \ \Omega,\\ f(x) > 0\quad &\mathrm{in}\ \ K, \\ \quad \end{array} \right.}$$

and w is a viscosity subsolution of

$$\displaystyle{\left \{\begin{array}{@{}l@{\quad }l@{}} H(x,Dw(x)) \leq -f(x)\quad &\mathrm{in}\ \ \Omega, \\ \frac{\partial w} {\partial \gamma } (x) \leq g(x) \quad &\mathrm{on}\ \ \partial \Omega.\end{array} \right.}$$

For any \(z \in \overline{\Omega } \setminus \mathcal{A}\), the function \(x\mapsto d(x,z)\) is not a viscosity supersolution of (SNP) at z while it is a viscosity subsolution of (SNP). Hence, according to Lemma 1.3, there exist a function \(\psi _{z} \in \mathrm{Lip}(\overline{\Omega })\), a neighborhood U z of z in \({\mathbb{R}}^{n}\) and a constant δ z  > 0 such that ψ z is a viscosity subsolution of (SNP) and it is moreover a viscosity subsolution of

$$\displaystyle{\left \{\begin{array}{llll} &H(x,D\psi _{z}(x)) \leq -\delta _{z}&&\mathrm{in}\ U_{z} \cap \Omega, \\ &\frac{\partial \psi _{z}} {\partial \gamma } (x) \leq g(x) -\delta _{z} &&\mathrm{on}\ U_{z} \cap \partial \Omega.\end{array} \right.}$$

We choose a function \(f_{z} \in C(\overline{\Omega })\) so that 0 < f z (x) ≤ δ for all \(x \in \overline{\Omega } \cap U_{z}\) and f z (x) = 0 for all \(x \in \overline{\Omega } \setminus U_{z}\), and note that ψ z is a viscosity subsolution of

$$\displaystyle{\left \{\begin{array}{llll} &H(x,D\psi _{z}(x)) \leq -f_{z}(x)&&\mathrm{in}\ \Omega, \\ &\frac{\partial \psi _{z}} {\partial \gamma } (x) \leq g(x) - f_{z}(x) &&\mathrm{on}\ \partial \Omega. \end{array} \right.}$$

We select a finite number of points z 1, , z k of K so that \(\{U_{z_{i}}\}_{i=1}^{k}\) covers K.

Now, we define the function \(\psi \in \mathrm{Lip}(\overline{\Omega })\) by

$$\displaystyle{\psi (x) = \frac{1} {k}\sum _{i=1}^{k}\psi _{ z_{i}}(x),}$$

and observe by Theorem 4.3 that ψ is a viscosity subsolution of

$$\displaystyle{\left \{\begin{array}{llll} &H(x,D\psi (x)) \leq -f(x)&&\mathrm{in}\ \Omega, \\ &\frac{\partial \psi } {\partial \gamma }(x) \leq g(x) - f(x) &&\mathrm{on}\ \partial \Omega, \end{array} \right.}$$

where \(f \in C(\overline{\Omega })\) is given by

$$\displaystyle{f(x) = \frac{1} {k}\sum _{i=1}^{k}f_{ z_{i}}(x).}$$

Finally, we note that inf K f > 0. □ 

4.2 Proof of Theorem 4.2

We give a proof of Theorem 4.2 in this subsection.

We begin by choosing continuous functions on \({\mathbb{R}}^{n}\) which extend the functions g, γ and ν. We denote them again by the same symbols g, γ and ν.

The following proposition guarantees the existence of test functions which are convenient to prove Theorem 4.2.

Theorem 4.7.

Let \(\varepsilon > 0\) and M > 0. Then there exist a constant Λ > 0 and moreover, for each R > 0, a neighborhood U of ∂Ω, a function \(\chi \in {C}^{1}((\Omega \cup U) \times {\mathbb{R}}^{n})\) and a constant δ > 0 such that for all \((x,\xi ) \in (\Omega \cup U) \times {\mathbb{R}}^{n}\) ,

$$\displaystyle{M\vert \xi \vert \leq \chi (x,\xi ) \leq \Lambda (\vert \xi \vert + 1),}$$

and for all (x,ξ) ∈ U × B R ,

$$\displaystyle{\gamma (x)\cdot D_{\xi }\chi (x,\xi )\left \{\begin{array}{@{}l@{\quad }l@{}} \leq g(x) + 2\varepsilon \quad &\mathrm{if}\ \ \nu (x)\cdot \xi \leq \delta, \\ \geq g(x) + \frac{\varepsilon } {2}\quad &\mathrm{if}\ \ \nu (x)\cdot \xi \geq -\delta. \end{array} \right.}$$

It should be noted that the constant Λ in the above statement does not depend on R while U, χ and δ do.

We begin the proof with two Lemmas.

We fix r > 1 and set

$$\displaystyle{\mathbb{R}_{r}^{2n} =\{ (y,z) \in {\mathbb{R}}^{n} \times {\mathbb{R}}^{n}\, :\, y \cdot z \geq {r}^{-1},\,\max \{\vert y\vert,\vert z\vert \}\leq r\}.}$$

We define the function \(\zeta \in {C}^{\infty }(\mathbb{R}_{r}^{2n} \times {\mathbb{R}}^{n})\) by

$$\displaystyle{\zeta (y,z,\xi ) =\Big \vert \xi - \frac{y\cdot \xi } {y \cdot z}z\Big{\vert }^{2} + {(y\cdot \xi )}^{2}.}$$

Lemma 4.3.

The function ζ has the properties:

$$\displaystyle{\left \{\begin{array}{llll} &\zeta (y,z,t\xi ) = {t}^{2}\zeta (y,z,\xi )\ \ &&\text{ for all }(y,z,\xi,t) \in \mathbb{R}_{r}^{2n} \times {\mathbb{R}}^{n} \times \mathbb{R}, \\ &\zeta (y,z,\xi ) > 0\ \ &&\text{ for all }(y,z,\xi ) \in \mathbb{R}_{r}^{2n} \times ({\mathbb{R}}^{n} \setminus \{ 0\}), \\ &z \cdot D_{\xi }\zeta (y,z,\xi ) = 2(y \cdot z)(y\cdot \xi )\ \ &&\text{ for all }(y,z,\xi ) \in \mathbb{R}_{r}^{2n} \times {\mathbb{R}}^{n}.\end{array} \right.}$$

This is a version of Lemma 3.4, the proof of which is easily adapted to the present case.

We define the function \(\phi \,:\, \mathbb{R}_{r}^{2n} \times {\mathbb{R}}^{n} \rightarrow \mathbb{R}\) by

$$\displaystyle{\phi (y,z,\xi ) ={ \left (\zeta (y,z,\xi ) + 1\right )}^{1/2}.}$$

Lemma 4.4.

There exists a constant Λ > 1, which depends only on r, such that for all \((y,z,\xi ) \in \mathbb{R}_{r}^{2n} \times {\mathbb{R}}^{n}\) ,

$$\displaystyle{\left \{\begin{array}{ll} &z \cdot D_{\xi }\phi (y,z,\xi ) =\phi {(y,z,\xi )}^{-1}(y \cdot z)(y\cdot \xi ), \\ &\max \{{\Lambda }^{-1}\vert \xi \vert,\,1\} \leq \phi (y,z,\xi ) \leq \Lambda (\vert \xi \vert + 1), \\ &\max \{\vert D_{y}\phi (y,z,\xi )\vert,\vert D_{z}\phi (y,z,\xi )\vert \}\leq \Lambda, \\ &\vert D_{\xi }\phi (y,z,\xi )\vert \leq \Lambda.\end{array} \right.}$$

Proof.

It is clear by the definition of ϕ that

$$\displaystyle{\phi (y,z,\xi ) \geq 1.}$$

We may choose a constant C > 1 so that for all \((y,z,\xi ) \in \mathbb{R}_{r}^{2n} \times {S}^{n-1}\),

$$\displaystyle{\max \{\zeta (y,z,\xi ),\zeta {(y,z,\xi )}^{-1},\vert D_{ y}\zeta (y,z,\xi )\vert,\vert D_{z}\zeta (y,z,\xi )\vert,\vert D_{\xi }\zeta (y,z,\xi )\vert \}\leq C,}$$

where \({S}^{n-1} :=\{ x \in {\mathbb{R}}^{n}\, :\, \vert x\vert = 1\}\). By the homogeneity of the function ζ(y, z, ξ) in ξ, we have

$$\displaystyle\begin{array}{rcl} \max \{\zeta (y,z,\xi ),\vert D_{y}\zeta (y,z,\xi )\vert,\vert D_{z}\zeta (y,z,\xi )\vert \}& \leq & C\vert \xi {\vert }^{2}, \\ \vert D_{\xi }\zeta (y,z,\xi )\vert & \leq & C\vert \xi \vert, \\ \zeta (y,z,\xi )& \geq & {C}^{-1}\vert \xi {\vert }^{2}{}\end{array}$$
(64)

for all \((y,z,\xi ) \in \mathbb{R}_{r}^{2n} \times {\mathbb{R}}^{n}\). From this it follows that

$$\displaystyle{{C}^{-1/2}\vert \xi \vert \leq \phi (y,z,\xi ) \leq {C}^{1/2}(\vert \xi \vert + 1).}$$

By a direct computation, we get

$$\displaystyle{D_{x}\phi (y,z,\xi ) = \frac{D_{x}\zeta (y,z,\xi )} {2\phi (y,z,\xi )} \ \ \ \text{ for }x = y,z,\xi.}$$

Hence, using (64), we get

$$\displaystyle{\vert D_{y}\phi (y,z,\xi )\vert \leq \frac{C\vert \xi {\vert }^{2}} {2\phi (y,z,\xi )} \leq {C}^{3/2}\vert \xi \vert.}$$

In the same way, we get

$$\displaystyle{\vert D_{z}\phi (y,z,\xi )\vert \leq {C}^{3/2}\vert \xi \vert.}$$

Also, we get

$$\displaystyle{\vert D_{\xi }\phi (y,z,\xi )\vert \leq \frac{C\vert \xi {\vert }^{2}} {2\phi (y,z,\xi )} \leq {C}^{3/2}\vert \xi \vert.}$$

We observe that

$$\displaystyle{z \cdot D_{\xi }\phi (y,z,\xi ) = \frac{z \cdot D_{\xi }\zeta (y,z,\xi )} {2\phi (y,z,\xi )} = \frac{(y \cdot z)(y\cdot \xi )} {\phi (y,z,\xi )}.}$$

By setting Λ = C 3 ∕ 2, we conclude the proof. □ 

Let α > 0. For any \(W \subset {\mathbb{R}}^{n}\) we denote by W α the α–neighborhood of W, that is,

$$\displaystyle{{W}^{\alpha } =\{ x \in {\mathbb{R}}^{n}\, :\, \mathrm{dist}(x,W) <\alpha \}.}$$

For each δ ∈ (0, 1) we select \(\nu _{\delta } \in {C}^{1}({\Omega }^{1}, {\mathbb{R}}^{n})\), \(\gamma _{\delta } \in {C}^{1}({\Omega }^{1}, {\mathbb{R}}^{n})\) and \(g_{\delta } \in {C}^{1}({\Omega }^{1}, \mathbb{R})\) so that for all x ∈ Ω 1,

$$\displaystyle{ \max \{\vert \nu _{\delta }(x) -\nu (x)\vert,\,\vert \gamma _{\delta }(x) -\gamma (x)\vert,\,\vert g_{\delta }(x) - g(x)\vert \} <\delta. }$$
(65)

(Just to be sure, note that \({\Omega }^{1} =\{ x \in {\mathbb{R}}^{n}\, :\, \mathrm{dist}(x,\Omega ) < 1\}\).)

By assumption, we have

$$\displaystyle{\nu (x) \cdot \gamma (x) > 0\ \ \ \text{ for all }x \in \partial \Omega.}$$

Hence, we may fix δ 0 ∈ (0, 1) so that

$$\displaystyle{\inf \{\nu _{\delta }(x) \cdot \gamma _{\delta }(x)\, :\, x \in {(\partial \Omega )}^{\delta _{0}},\ \delta \in (0,\,\delta _{0})\} > 0.}$$

We choose a constant r > 1 so that if δ ∈ (0, δ 0), then

$$\displaystyle{ \left \{\begin{array}{ll} &\min \{\nu _{\delta }(x) \cdot \gamma _{\delta }(x),\,\vert \gamma _{\delta }(x)\vert \}\geq {r}^{-1}, \\ &\max \{\vert \nu _{\delta }(x)\vert,\,\vert \gamma _{\delta }(x)\vert \}\leq r, \\ &\vert g_{\delta }(x)\vert + 1 < r. \end{array} \right. }$$
(66)

for all \(x \in {(\partial \Omega )}^{\delta _{0}}\). In particular, we have

$$\displaystyle{ (\nu _{\delta }(x),\,\gamma _{\delta }(x)) \in \mathbb{R}_{r}^{2n}\ \ \ \text{ for all }\ x \in {(\partial \Omega )}^{\delta _{0} }\ \text{ and }\ \delta \in (0,\,\delta _{0}). }$$
(67)

To proceed, we fix any \(\varepsilon \in (0,\,1)\), M > 0 and R > 0. For each δ ∈ (0, δ 0) we define the function \(\psi _{\delta } \in {C}^{1}({(\partial \Omega )}^{\delta _{0}} \times {\mathbb{R}}^{n})\) by

$$\displaystyle{\psi _{\delta }(x,\xi ) = (g_{\delta }(x)+\varepsilon ) \frac{\gamma _{\delta }(x)\cdot \xi } {\vert \gamma _{\delta }(x){\vert }^{2}},}$$

choose a cut-off function \(\eta _{\delta } \in C_{0}^{1}({\mathbb{R}}^{n})\) so that

$$\displaystyle{\left \{\begin{array}{ll} &\mathrm{supp}\eta _{\delta } \subset {(\partial \Omega )}^{\delta }, \\ &0 \leq \eta _{\delta }(x) \leq 1\ \ \ \text{ for all }x \in {\mathbb{R}}^{n}, \\ &\eta _{\delta }(x) = 1\ \ \ \text{ for all }x \in {(\partial \Omega )}^{\delta /2}, \end{array} \right.}$$

and define the function \(\chi _{\delta } \in {C}^{1}({\Omega }^{\delta _{0}})\) by

$$\displaystyle{\chi _{\delta }(x,\xi ) = M\langle \xi \rangle (1 -\eta _{\delta }(x)) +\eta _{\delta }(x)\left [\psi _{\delta }(x,\xi ) + ({r}^{2} + M)\Lambda \phi _{\delta }(x,\xi )\right ],}$$

where Λ and ϕ are the constant and function from Lemma 4.4,  \(\langle \xi \rangle := {(\vert \xi {\vert }^{2} + 1)}^{1/2}\)  and  \(\phi _{\delta }(x,\xi ) :=\phi (\nu _{\delta }(x),\gamma _{\delta }(x),\xi )\). Since \(\mathrm{supp}\eta _{\delta } \subset {(\partial \Omega )}^{\delta _{0}}\) for all δ ∈ (0, δ 0), in view of (67) we see that χ δ is well-defined.

Proof (Theorem 4.7). 

Let δ 0 ∈ (0, 1) and \(\psi _{\delta },\,\phi _{\delta },\,\chi _{\delta } \in {C}^{1}({\Omega }^{\delta _{0}} \times {\mathbb{R}}^{n})\) be as above.

Let δ ∈ (0, δ 0), which will be fixed later on. It is obvious that for all \((x,\xi ) \in {(\Omega )}^{\delta _{0}} \times {\mathbb{R}}^{n}\),

$$\displaystyle{ \left \{\begin{array}{ll} &\gamma _{\delta }(x) \cdot D_{\xi }\psi _{\delta }(x,\xi ) = g_{\delta }(x)+\varepsilon, \\ &\vert \psi _{\delta }(x,\xi )\vert \leq {r}^{2}\vert \xi \vert.\end{array} \right. }$$
(68)

For any \((x,\xi ) \in {(\partial \Omega )}^{\delta } \times {\mathbb{R}}^{n}\), using (66), (68) and Lemma 4.4, we get

$$\displaystyle{\psi _{\delta }(x,\xi ) + ({r}^{2} + M)\Lambda \phi _{\delta }(x,\xi ) \geq -{r}^{2}\vert \xi \vert + ({r}^{2} + M)\vert \xi \vert \geq M\vert \xi \vert,}$$

and

$$\displaystyle\begin{array}{rcl} \psi _{\delta }(x,\xi ) + ({r}^{2} + M)\Lambda \phi _{\delta }(x,\xi )& \leq & \,{r}^{2}\vert \xi \vert + ({r}^{2} + M){\Lambda }^{2}(\vert \xi \vert + 1) {}\\ & \leq & \,(2{r}^{2} + M){\Lambda }^{2}(\vert \xi \vert + 1). {}\\ \end{array}$$

Thus, we have

$$\displaystyle{ M\vert \xi \vert \leq \chi _{\delta }(x,\xi ) \leq (2{r}^{2} + M){\Lambda }^{2}(\vert \xi \vert + 1)\ \ \text{ for all }\ (x,\xi ) \in {\Omega }^{\delta } \times {\mathbb{R}}^{n}. }$$
(69)

Now, note that if \((x,\xi ) \in {(\partial \Omega )}^{\delta /2} \times {\mathbb{R}}^{n}\), then

$$\displaystyle{\chi _{\delta }(x,\xi ) =\psi _{\delta }(x,\xi ) + ({r}^{2} + M)\Lambda \phi _{\delta }(x,\xi ).}$$

Hence, by Lemma 4.4 and (68), we get

$$\displaystyle{\gamma _{\delta }(x) \cdot D_{\xi }\chi _{\delta }(x,\xi ) = g_{\delta }(x) +\varepsilon +({r}^{2} + M)\Lambda \frac{(\nu _{\delta }(x) \cdot \gamma _{\delta }(x))(\nu _{\delta }(x)\cdot \xi )} {\phi _{\delta }(x,\xi )} }$$

for all \((x,\xi ) \in {(\partial \Omega )}^{\delta /2} \times {\mathbb{R}}^{n}\).

Next, let \((x,\xi ) \in {\Omega }^{\delta } \times {\mathbb{R}}^{n}\). Since

$$\displaystyle{D_{\xi }\chi _{\delta }(x,\xi ) = M(1 -\eta _{\delta }(x))D\langle \xi \rangle +\eta _{\delta }(x)\left [D_{\xi }\psi _{\delta }(x,\xi ) + ({r}^{2} + M)\Lambda D_{\xi }\phi _{\delta }(x,\xi )\right ],}$$

using Lemma 4.4, we get

$$\displaystyle\begin{array}{rcl} \vert D_{\xi }\chi _{\delta }(x,\xi )\vert & \leq & \,\max \left \{M\vert D\langle \xi \rangle \vert,\, \frac{\vert g_{\delta }(x) +\varepsilon \vert } {\vert \gamma _{\delta }(x)\vert } + ({r}^{2} + M)\Lambda \vert D_{\xi }\phi _{\delta }(x,\xi )\vert \right \} \\ &\leq & \,\max \{M,\,{r}^{2} + ({r}^{2} + M){\Lambda }^{2}\} = (2{r}^{2} + M){\Lambda }^{2}.{}\end{array}$$
(70)

Let \((x,\xi ) \in {(\partial \Omega )}^{\delta /2} \times B_{R}\). Note by (65) and (70) that

$$\displaystyle{\left \vert (\gamma _{\delta }(x) -\gamma (x)) \cdot D_{\xi }\chi _{\delta }(x,\xi )\right \vert \leq \delta (2{r}^{2} + M){\Lambda }^{2}.}$$

Note also that if ν(x) ⋅ξ ≤ δ, then

$$\displaystyle\begin{array}{rcl} & & ({r}^{2} + M)\Lambda \frac{(\nu _{\delta }(x) \cdot \gamma _{\delta }(x))(\nu _{\delta }(x)\cdot \xi )} {\phi _{\delta }(x,\xi )} {}\\ & & \quad \leq ({r}^{2} + M)\Lambda \frac{(\nu _{\delta }(x) \cdot \gamma _{\delta }(x))(\nu (x)\cdot \xi )} {\phi _{\delta }(x,\xi )} + ({r}^{2} + M)\Lambda {r}^{2}R\delta {}\\ & & \quad \leq ({r}^{2} + M)\Lambda {r}^{2}\delta (1 + R). {}\\ \end{array}$$

Hence, if ν(x) ⋅ξ ≤ δ, then

$$\displaystyle\begin{array}{rcl} & \gamma & (x) \cdot D_{\xi }\chi _{\delta }(x,\xi ) \leq \gamma _{\delta }(x) \cdot D_{\xi }\chi _{\delta }(x,\xi ) +\delta (2{r}^{2} + M){\Lambda }^{2} {}\\ & \leq & \delta (2{r}^{2} + M){\Lambda }^{2} + g_{\delta }(x) +\varepsilon +({r}^{2} + M)\Lambda \frac{(\nu _{\delta }(x) \cdot \gamma _{\delta }(x))(\nu _{\delta }(x)\cdot \xi )} {\phi _{\delta }(x,\xi )} {}\\ & \leq & g(x) +\varepsilon +\delta \left [1 + (2{r}^{2} + M){\Lambda }^{2}{r}^{2} + ({r}^{2} + M)\Lambda {r}^{2}(1 + R)\right ]. {}\\ \end{array}$$

Similarly, we see that if ν(x) ⋅ξ ≥ − δ, then

$$\displaystyle{ \gamma (x) \cdot D_{\xi }\chi _{\delta }(x,\xi ) \geq g(x) +\varepsilon -\delta \left [1 + (2{r}^{2} + M){\Lambda }^{2}{r}^{2} + ({r}^{2} + M)\Lambda {r}^{2}(1 + R)\right ]. }$$

If we select δ ∈ (0, δ 0) so that

$$\displaystyle{\delta \left [1 + (2{r}^{2} + M){\Lambda }^{2}{r}^{2} + ({r}^{2} + M)\Lambda {r}^{2}(1 + R)\right ] \leq \frac{\varepsilon } {2},}$$

then we have for all (x, ξ) ∈ (∂Ω)δ ∕ 2 ×B R ,

$$\displaystyle{\gamma (x)\cdot D_{\xi }\chi _{\delta }(x,\xi )\left \{\begin{array}{@{}l@{\quad }l@{}} \leq g(x) + 2\varepsilon \quad &\mathrm{if}\ \ \nu (x)\cdot \xi \leq \delta, \\ \geq g(x) + \frac{\varepsilon } {2}\quad &\mathrm{if}\ \ \nu (x)\cdot \xi \geq -\delta. \end{array} \right.}$$

Thus, the function χ = χ δ has the required properties, with (∂Ω)δ ∕ 2 and (2r 2 + M)Λ 2 in place of U and Λ, respectively. □ 

We are ready to prove the following theorem.

Theorem 4.8.

Let \(\varepsilon > 0\) and \(u \in \mathrm{Lip}(\overline{\Omega })\) be a viscosity subsolution of (SNP) . Then there exist a neighborhood U of ∂Ω and a function \(u_{\varepsilon } \in {C}^{1}(\Omega \cup U)\) such that

$$\displaystyle{ \left \{\begin{array}{llll} &H(x,\mathit{Du}_{\varepsilon }(x)) \leq \varepsilon &&\text{ for all }\ x \in \Omega \cup U, \\ &\gamma (x) \cdot \mathit{Du}_{\varepsilon }(x) \leq g(x)+\varepsilon &&\text{ for all }\ x \in U, \\ &\|u_{\varepsilon } - u\|_{\infty,\Omega } \leq \varepsilon. && \end{array} \right. }$$
(71)

Proof.

Fix any \(\varepsilon > 0\) and a constant M > 1 so that M − 1 is a Lipschitz bound of the function u. With these constants \(\varepsilon\) and M, let Λ > 0 be the constant from Theorem 4.7. Set R = M + 2Λ, and let U, χ and δ be as in Theorem 4.7.

Let α > 0. We define the sup-convolution u α  ∈ C(ΩU) by

$$\displaystyle{u_{\alpha }(x) =\max _{y\in \overline{\Omega }}(u(y) -\alpha \chi (x,(y - x)/\alpha )).}$$

Let x ∈ ΩU, p ∈ D  +  u α (x) and \(y \in \overline{\Omega }\) be a maximum point in the definition of u α , that is,

$$\displaystyle{ u_{\alpha }(x) = u(y) -\alpha \chi (x,(y - x)/\alpha ). }$$
(72)

It is easily seen that

$$\displaystyle{ \left \{\begin{array}{ll} &D_{\xi }\chi (x,(y - x)/\alpha ) \in {D}^{+}u(y), \\ &p = D_{\xi }\chi (x,(y - x)/\alpha ) -\alpha D_{x}\chi (x,(y - x)/\alpha ). \end{array} \right. }$$
(73)

Fix an α 0 ∈ (0, 1) so that

$$\displaystyle{\overline{{(\partial \Omega )}^{\alpha _{0}^{2} }}\,\, \subset U.}$$

Here, of course, \(\overline{V }\) denotes the closure of V. For α ∈ (0, α 0) we set \(U_{\alpha } = {(\partial \Omega )}^{{\alpha }^{2} }\) and \(V _{\alpha } = \Omega \cup U_{\alpha } = {\Omega }^{{\alpha }^{2} }\). Note that \(\chi \in {C}^{1}(\overline{V }_{\alpha } \times {\mathbb{R}}^{n})\). We set \(W_{\alpha } =\{ (x,y) \in V _{\alpha } \times \overline{\Omega }\, :\, (<InternalRef RefID="Equ121">72</InternalRef>)\ \text{ holds}\}\).

Now, we fix any α ∈ (0, α 0). Let (x, y) ∈ W α . We may choose a point \(z \in \overline{\Omega }\) so that | x − z |  < α 2. Note that

$$\displaystyle{u(y) -\alpha \chi (x,(y - x)/\alpha ) = u_{\alpha }(x) \geq u(z) -\alpha \chi (x,(z - x)/\alpha ).}$$

Hence,

$$\displaystyle{\alpha \chi (x,(y - x)/\alpha ) \leq (M - 1)\vert z - y\vert +\alpha \chi (x,(z - x)/\alpha ).}$$

Now, since M | ξ | ≤ χ(x, ξ) ≤ Λ( | ξ |  + 1)) for all \((x,\xi ) \in V _{\alpha } \times {\mathbb{R}}^{n}\) and | x − z | ≤ α 2 < α, we get

$$\displaystyle\begin{array}{rcl} M\vert x - y\vert & \leq & \,(M - 1)(\vert x - y\vert {+\alpha }^{2}) +\alpha \Lambda (\vert z - x\vert /\alpha + 1) {}\\ & \leq & \,(M - 1)\vert x - y\vert +\alpha (M + 2\Lambda ). {}\\ \end{array}$$

Consequently,

$$\displaystyle{ \vert y - x\vert \leq \alpha (M + 2\Lambda ) = R\alpha \ \ \ \text{ for all }\ (x,y) \in W_{\alpha }. }$$
(74)

Next, we choose a constant C > 0 so that

$$\displaystyle{\vert D_{x}\chi (x,\xi )\vert + \vert D_{\xi }\chi (x,\xi )\vert \leq C\ \ \ \text{ for all }\ (x,\xi ) \in V _{\alpha _{0}} \times B_{R}.}$$

Let (x, y) ∈ W α and \(z \in B_{R\alpha }(x) \cap V _{\alpha _{0}}\). Assume moreover that x ∈ U. In view of (74) and the choice of χ and δ, we have

$$\displaystyle{\gamma (x)\cdot D_{\xi }\chi (x,(y-x)/\alpha )\left \{\begin{array}{@{}l@{\quad }l@{}} \leq g(x) + 2\varepsilon \ \ \ \quad &\mathrm{if}\ \ \nu (x) \cdot (y - x) \leq \alpha \delta, \\ \geq g(x) + \frac{\varepsilon } {2}\quad &\mathrm{if}\ \ \nu (x) \cdot (y - x) \geq -\alpha \delta. \end{array} \right.}$$

We observe that

$$\displaystyle{\nu (x)\cdot (y-x)\left \{\begin{array}{@{}l@{\quad }l@{}} \leq \frac{\alpha \delta } {2} +\omega _{\nu }(R\alpha )R\alpha \quad &\mathrm{if}\ \ \nu (z) \cdot (y - x) \leq \frac{\alpha \delta } {2}, \\ \geq \frac{\alpha \delta } {2} -\omega _{\nu }(R\alpha )R\alpha \quad &\mathrm{if}\ \ \nu (z) \cdot (y - x) \geq -\frac{\alpha \delta } {2}, \end{array} \right.}$$

where ω ν denotes the modulus of continuity of the function ν on \(V _{\alpha _{0}}\). Observe as well that

$$\displaystyle\begin{array}{rcl} \left \vert \gamma (z) \cdot D_{\xi }\chi (x,(y - x)/\alpha ) -\gamma (x) \cdot D_{\xi }\chi (x,(y - x)/\alpha )\right \vert & \leq & \,C\omega _{\gamma }(R\alpha ), {}\\ \vert g(z) - g(x)\vert & \leq & \,\omega _{g}(R\alpha ), {}\\ \end{array}$$

where ω γ and ω g denote the moduli of continuity of the functions γ and g on the set \(V _{\alpha _{0}}\), respectively.

We may choose an α 1 ∈ (0, α 0) so that

$$\displaystyle{\omega _{\nu }(R\alpha _{1})R < \frac{\delta } {2}\ \ \text{ and }\ \ C\omega _{\gamma }(R\alpha _{1}) +\omega _{g}(R\alpha _{1}) < \frac{\varepsilon } {4},}$$

and conclude from the above observations that for all (x, y) ∈ W α and \(z_{i} \in B_{R\alpha }(x) \cap V _{\alpha _{0}}\), with i = 1, 2, 3, if  x ∈ U  and  α < α 1, then

$$\displaystyle{ \gamma (z_{1})\cdot D_{\xi }\chi (x,(y-x)/\alpha )\left \{\begin{array}{@{}l@{\quad }l@{}} \leq g(z_{2}) + 3\varepsilon \quad &\mathrm{if}\ \ \nu (z_{3}) \cdot (y - x) \leq \alpha \delta /2, \\ \geq g(z_{2}) + \frac{\varepsilon } {4}\quad &\mathrm{if}\ \ \nu (z_{3}) \cdot (y - x) \geq -\alpha \delta /2. \end{array} \right. }$$
(75)

We may assume, by reselecting α 1 > 0 small enough if necessary, that

$$\displaystyle{ {(\partial \Omega )}^{R\alpha _{1} } \subset U. }$$
(76)

In what follows we assume that α ∈ (0, α 1). Let (x, y) ∈ W α and \(p \in {D}^{+}u_{\alpha }(x)\). By (73) and (74), we have

$$\displaystyle{ \max \{\vert p\vert,\,\vert D_{\xi }\chi (x,(y - x)/\alpha )\vert \}\leq C(1+\alpha ). }$$
(77)

Let ω H denote the modulus of continuity of H on \(V _{\alpha _{0}} \times B_{C(1+\alpha _{0})}\).

We now assume that y ∈ ∂Ω. By (74) and (76), we have x ∈ U. Let ρ be a defining function of Ω. We may assume that   |  | ≤ 1 in \(V _{\alpha _{0}}\)  and  \(\rho _{0} :=\inf _{U_{\alpha _{ 0}}}\vert D\rho \vert > 0\). Observe that

$$\displaystyle{{\alpha }^{2} >\rho (x) =\rho (x) -\rho (y) = D\rho (z) \cdot (x - y) = \vert D\rho (z)\vert \nu (z) \cdot (x - y)}$$

for some point z on the line segment [x, y]. Hence, we get

$$\displaystyle{\nu (z) \cdot (x - y) \leq \rho _{0}^{-1}{\alpha }^{2}.}$$

If α ≤ ρ 0 δ ∕ 2, then

$$\displaystyle{\nu (z) \cdot (y - x) \geq -\alpha \delta /2.}$$

Hence, noting that | z − x | ≤ | x − y |  < , by (75), we get

$$\displaystyle{\gamma (y) \cdot D_{\xi }\chi (x,(y - x)/\alpha ) \geq g(y) + \frac{\varepsilon } {4},}$$

and, by the viscosity property of u,

$$\displaystyle{0 \geq H(y,D_{\xi }\chi (x,(y - x)/\alpha )) \geq H(x,p) -\omega _{H}((R + C)\alpha ).}$$

Thus, if \(\omega _{H}((R + C)\alpha ) <\varepsilon\) and α ≤ ρ 0 δ ∕ 2, then we have

$$\displaystyle{H(x,p) \leq \varepsilon.}$$

On the other hand, if y ∈ Ω, then, by the viscosity property of u, we have

$$\displaystyle{0 \geq H(y,D_{\xi }\chi (x,(y - x)/\alpha )).}$$

Therefore, if \(\omega _{H}((R + C)\alpha ) <\varepsilon\), then

$$\displaystyle{H(x,p) \leq \varepsilon.}$$

We may henceforth assume by selecting α 1 > 0 small enough that

$$\displaystyle{\omega _{H}((R + C)\alpha _{1}) <\varepsilon \ \ \ \text{ and }\ \ \ \alpha _{1} \leq \rho _{0}\delta /2,}$$

and we conclude that u α is a viscosity subsolution of

$$\displaystyle{ H(x,\mathit{Du}_{\alpha }(x)) \leq \varepsilon \ \ \ \mathrm{in}\ V _{\alpha }. }$$
(78)

As above, let (x, y) ∈ W α and \(p \in {D}^{+}u_{\alpha }(x)\). We assume that x ∈ U α . Then

$$\displaystyle{{-\alpha }^{2} <\rho (x) \leq \rho (x) -\rho (y) \leq D\rho (z) \cdot (x - y)}$$

for some z ∈ [x, y], which yields

$$\displaystyle{\nu (z) \cdot (y - x) < \vert D\rho (z){\vert {}^{-1}\alpha }^{2} \leq \rho _{ 0}^{-1}{\alpha }^{2}.}$$

Hence, if α ≤ ρ 0 δ ∕ 2, then

$$\displaystyle{\nu (z) \cdot (y - x) \leq \frac{\delta \alpha } {2},}$$

and, by (75), we get

$$\displaystyle{\gamma (x) \cdot D_{\xi }\chi (x,(y - x)/\alpha ) \leq g(x) + 3\varepsilon.}$$

Furthermore,

$$\displaystyle\begin{array}{rcl} \gamma (x) \cdot p& \leq & \,\gamma (x) \cdot D_{\xi }\chi (x,(y - x)/\alpha ) +\alpha C\|\gamma \|_{\infty,U_{\alpha _{ 0}}} {}\\ & \leq & \,g(x) + 3\varepsilon +\alpha C\|\gamma \|_{\infty,U_{\alpha _{ 0}}}. {}\\ \end{array}$$

We may assume again by selecting α 1 > 0 small enough that

$$\displaystyle{\alpha _{1}C\|\gamma \|_{\infty,U_{\alpha _{ 0}}} <\varepsilon.}$$

Thus, u α is a viscosity subsolution of

$$\displaystyle{ \gamma (x) \cdot \mathit{Du}_{\alpha }(x) \leq g(x) + 4\varepsilon \ \ \ \mathrm{in}\ \ U_{\alpha }. }$$
(79)

Let (x, y) ∈ W α and observe by using (74) that if \(x \in \overline{\Omega }\), then

$$\displaystyle{\vert u(x) - u_{\alpha }(x)\vert \leq \vert u(x) - u(y)\vert +\alpha \vert \chi (x,(y - x)/\alpha )\vert \leq \alpha (MR + C).}$$

We fix α ∈ (0, α 1) so that \(\alpha _{1}(MR + C) <\varepsilon\), and conclude that u α is a viscosity subsolution of (78) and (79) and satisfies

$$\displaystyle{\|u_{\alpha } - u\|_{\infty,\Omega } \leq \varepsilon.}$$

The final step is to mollify the function u α . Let {k λ } λ > 0 be a collection of standard mollification kernels.

We note by (77) or (78) that u α is Lipschitz continuous on any compact subset of \(V _{\alpha }\). Fix any λ ∈ (0, α 2 ∕ 4). We note that the closure of V α ∕ 2 + B λ is a compact subset of V α . Let M 1 > 0 be a Lipschitz bound of the function u α on V α ∕ 2 + B λ .

We set

$$\displaystyle{{u}^{\lambda }(x) = u_{\alpha } {\ast} k_{\lambda }(x)\ \ \ \text{ for }\ x \in V _{\alpha /2}.}$$

In view of Rademacher’s theorem (see Appendix A.6), we have

$$\displaystyle\begin{array}{rcl} & & H(x,\mathit{Du}_{\alpha }(x)) \leq \varepsilon \ \ \qquad \qquad \qquad \text{ for a.e. }\ x \in V _{\alpha }, {}\\ & & \gamma (x) \cdot \mathit{Du}_{\alpha }(x) \leq g(x) + 4\varepsilon \ \ \quad \qquad \text{ for a.e. }\ x \in U_{\alpha }. {}\\ \end{array}$$

Here Du α denotes the distributional derivative of u α , and we have

$$\displaystyle{{\mathit{Du}}^{\lambda } = k_{\lambda } {\ast}\mathit{Du}_{\alpha }\ \ \ \mathrm{in}\ \ V _{\alpha /2}.}$$

By Jensen’s inequality, we get

$$\displaystyle\begin{array}{rcl} H(x,{\mathit{Du}}^{\lambda }(x))& \leq & \,\int _{B_{\lambda }}H(x,\mathit{Du}_{\alpha }(x - y))k_{\lambda }(y)\,\mathrm{d}y {}\\ & \leq & \,\int _{B_{\lambda }}H(x - y,\mathit{Du}_{\alpha }(x - y))k_{\lambda }(y)\,\mathrm{d}y +\omega _{H}(\lambda ) {}\\ & \leq & \,\varepsilon +\omega _{H}(\lambda ), {}\\ \end{array}$$

where ω H is the modulus of continuity of H on the set \(V _{\alpha } \times B_{M_{1}}\). Similarly, we get

$$\displaystyle{\gamma (x) \cdot {\mathit{Du}}^{\lambda }(x) \leq g(x) + 4\varepsilon +\omega _{g}(\lambda ) + M_{1}\omega _{\gamma }(\lambda ),}$$

where ω g and ω γ are the moduli of continuity of the functions g and γ on V α , respectively. If we choose λ > 0 small enough, then (71) holds with \({u}^{\lambda } \in {C}^{1}(V _{\alpha /2})\), U α ∕ 2 and \(5\varepsilon\) in place of \(u_{\varepsilon }\), \(U\) and \(\varepsilon\), respectively. The proof is complete. □ 

Proof (Theorem 4.2). 

Let \(\varepsilon > 0\) and \(u \in \mathrm{Lip}(\overline{\Omega })\) be a viscosity subsolution of (SNP). Let ρ be a defining function of Ω. We may assume that

$$\displaystyle{D\rho (x) \cdot \gamma (x) \geq 1\ \ \ \text{ for all }\ x \in \partial \Omega.}$$

For δ > 0 we set

$$\displaystyle{{u}^{\delta }(x) = u(x) -\delta \rho (x)\ \ \ \text{ for }x \in \overline{\Omega }.}$$

It is easily seen that if δ > 0 is small enough, then u δ is a viscosity subsolution of

$$\displaystyle{\left \{\begin{array}{llll} &H(x,{\mathit{Du}}^{\delta }(x)) \leq \varepsilon &&\mathrm{in}\ \ \Omega, \\ &\gamma (x) \cdot {\mathit{Du}}^{\delta }(x) \leq g(x)-\delta &&\mathrm{on}\ \ \partial \Omega, \end{array} \right.}$$

and the following inequality holds:

$$\displaystyle{\|{u}^{\delta } - u\|_{\infty,\Omega } \leq \varepsilon.}$$

Then, Theorem 4.8, with \(\min \{\varepsilon,\delta \}\), u δ, \(H-\varepsilon\) and g − δ in place of \(\varepsilon\), u, H and g, respectively, ensures that there are a neighborhood U of ∂Ω and a function \(u_{\varepsilon } \in {C}^{1}(\Omega \cup U)\) such that

$$\displaystyle{\left \{\begin{array}{llll} &H(x,\mathit{Du}_{\varepsilon }(x)) \leq 2\varepsilon &&\mathrm{in}\ \ \Omega \cup U, \\ &\gamma (x) \cdot \mathit{Du}_{\varepsilon }(x) \leq g(x)&&\mathrm{in}\ \ U, \\ &\|u_{\varepsilon } - u\|_{\infty,\Omega } \leq 2\varepsilon, \end{array} \right.}$$

which concludes the proof. □ 

5 Optimal Control Problem Associated with (ENP)–(ID)

In this section we introduce an optimal control problem associated with the initial-boundary value problem (ENP)–(ID),

5.1 Skorokhod Problem

In this section, following [39, 44], we study the Skorokhod problem. We recall that \(\mathbb{R}_{+}\) denotes the interval \((0,\,\infty )\), so that \(\overline{\mathbb{R}}_{+} = [0,\,\infty )\). We denote by \(L_{\text{loc}}^{1}(\overline{\mathbb{R}}_{+},\, {\mathbb{R}}^{k})\) (resp., \(\mathrm{AC}_{ \text{loc}}(\overline{\mathbb{R}}_{+},\, {\mathbb{R}}^{k})\)) the space of functions \(v : \overline{\mathbb{R}}_{+} \rightarrow {\mathbb{R}}^{k}\) which are integrable (resp., absolutely continuous) on any bounded interval \(J \subset \overline{\mathbb{R}}_{+}\).

Given \(x \in \overline{\Omega }\) and \(v \in L_{\text{loc}}^{1}(\overline{\mathbb{R}}_{+}, {\mathbb{R}}^{n})\), the Skorokhod problem is to seek for a pair of functions, \((\eta,\,l) \in \mathrm{AC}_{\text{loc}}(\overline{\mathbb{R}}_{+}, {\mathbb{R}}^{n}) \times L_{\text{loc}}^{1}(\overline{\mathbb{R}}_{+},\, \mathbb{R})\), such that

$$\displaystyle{ \left \{\begin{array}{llll} &\eta (0) = x, \\ &\eta (t) \in \overline{\Omega }\ \ &&\text{ for all }t \in \overline{\mathbb{R}}_{+}, \\ &\dot{\eta }(t) + l(t)\gamma (\eta (t)) = v(t)\ \ &&\text{ for a.e. }t \in \overline{\mathbb{R}}_{+}, \\ &l(t) \geq 0\ \ &&\text{ for a.e. }t \in \overline{\mathbb{R}}_{+}, \\ &l(t) = 0\ \mathrm{if}\ \eta (t) \in \Omega \ \ &&\text{ for a.e. }t \in \overline{\mathbb{R}}_{+}. \end{array} \right. }$$
(80)

Regarding the solvability of the Skorokhod problem, our main claim is the following.

Theorem 5.1.

Let \(v \in L_{\mathrm{loc}}^{1}(\overline{\mathbb{R}}_{+},\, {\mathbb{R}}^{n})\) and \(x \in \overline{\Omega }\) . Then there exits a pair \((\eta,\,l) \in \mathrm{AC}_{\mathrm{loc}}(\overline{\mathbb{R}}_{+},\, {\mathbb{R}}^{n}) \times L_{\mathrm{loc}}^{1}(\overline{\mathbb{R}}_{+},\, \mathbb{R})\) such that (80) holds.

We refer to [44] and references therein for more general viewpoints (especially, for applications to stochastic differential equations with reflection) on the Skorokhod problem.

A natural question arises whether uniqueness of the solution (η, l) holds or not in the above theorem. On this issue we just give the following counterexample and do not discuss it further.

Example 5.1.

Let n = 2 and \(\Omega =\{ x = (x_{1},x_{2}) \in {\mathbb{R}}^{2}\, :\, x_{1} > 0\}\). (For simplicity of presentation, we consider the case where Ω is unbounded.) Define \(\gamma \in C(\partial \Omega,\, {\mathbb{R}}^{2})\) and \(v \in {L}^{\infty }(\overline{\mathbb{R}}_{+},\, {\mathbb{R}}^{2})\) by

$$\displaystyle{\gamma (0,x_{2}) = (-1,\,-3\vert x_{2}{\vert }^{-1/3}x_{ 2})\ \ \ \text{ and }\ \ \ v(t) = (-1,\,0).}$$

Set

$$\displaystyle{{\eta }^{\pm }(t) = (0,\,\pm {t}^{3})\ \ \ \text{ for all }t \geq 0.}$$

Then the pairs (η  + , 1) and (η  − , 1) are both solutions of (80), with η  ± (0) = (0, 0).

We first establish the following assertion.

Theorem 5.2.

Let \(v \in {L}^{\infty }(\overline{\mathbb{R}}_{+},\, {\mathbb{R}}^{n})\) and \(x \in \overline{\Omega }\) . Then there exits a pair \((\eta,\,l) \in \mathrm{Lip}(\overline{\mathbb{R}}_{+},\, {\mathbb{R}}^{n}) \times {L}^{\infty }(\overline{\mathbb{R}}_{+},\, \mathbb{R})\) such that (80) holds.

Proof.

We may assume that γ is defined and continuous on \({\mathbb{R}}^{n}\). Let \(\rho \in {C}^{1}({\mathbb{R}}^{n})\) be a defining function of Ω. We may assume that liminf | x | →  ρ(x) > 0 and that is bounded on \({\mathbb{R}}^{n}\). We may select a constant δ > 0 so that for all \(x \in {\mathbb{R}}^{n}\),

$$\displaystyle{\gamma (x) \cdot D\rho (x) \geq \delta \vert D\rho (x)\vert \ \ \ \text{ and }\ \ \ \vert D\rho (x)\vert \geq \delta \ \ \ \mathrm{if}\ \ 0 \leq \rho (x) \leq \delta.}$$

We set q(x) = (ρ(x) ∨ 0) ∧ δ for \(x \in {\mathbb{R}}^{n}\) and observe that q(x) = 0 for all \(x \in \overline{\Omega }\) and q(x) > 0 for all \(x \in {\mathbb{R}}^{n} \setminus \overline{\Omega }\).

Fix \(\varepsilon > 0\) and \(x \in \overline{\Omega }\). We consider the initial value problem for the ODE

$$\displaystyle{ \dot{\xi }(t) + \frac{1} {\varepsilon } q(\xi (t))\gamma (\xi (t)) = v(t)\ \text{ for a.e. }t \in \mathbb{R}_{+},\quad \xi (0) = x. }$$
(81)

By the standard ODE theory, there is a solution \(\xi \in \mathrm{Lip}(\overline{\mathbb{R}}_{+})\) of (81). Fix such a solution \(\xi \in \mathrm{Lip}(\overline{\mathbb{R}}_{+},\, {\mathbb{R}}^{n})\) in what follows.

Note that \((\mathrm{d}\,q \circ \xi /\mathrm{d}t)(t) = D\rho (\xi (t)) \cdot \dot{\xi } (t)\) a.e. in the set \(\{t \in \mathbb{R}_{+}\, :\,\rho \circ \xi (t) \in (0,\,\delta )\}\). Moreover, noting that \(q\circ \xi \in \mathrm{Lip}(\overline{\mathbb{R}}_{+},\, \mathbb{R})\) and hence it is differentiable a.e., we deduce that (d q ∘ ξ ∕ dt)(t) = 0 a.e. in the set \(\{t \in \mathbb{R}_{+}\, :\,\rho \circ \xi (t) \in \{ 0,\,\delta \}\}\).

Let m ≥ 2. We multiply the ODE of (81) by mq(ξ(t))m − 1 (ξ(t)), to get

$$\displaystyle{ \frac{\mathrm{d}} {\mathrm{d}t}q{(\xi (t))}^{m} + \frac{m} {\varepsilon } q{(\xi (t))}^{m}Dq(\xi (t)) \cdot \gamma (\xi (t)) = mq{(\xi (t))}^{m-1}Dq(\xi (t)) \cdot v(t)}$$

a.e. in the set \(\{t \in \mathbb{R}_{+}\, :\,\rho \circ \xi (t) \in (0,\,\delta )\}\). For any \(T \in \mathbb{R}_{+}\), integration over E T : = { t ∈ [0, T] : ρ ∘ ξ(t) ∈ (0, δ)} yields

$$\displaystyle\begin{array}{rcl} & & q{(\xi (T))}^{m} - q{(\xi (0))}^{m} + \frac{m} {\varepsilon } \int _{E_{T}}q{(\xi (s))}^{m}\gamma (\xi (s)) \cdot D\rho (\xi (s))\mathrm{d}s {}\\ & & \quad = m\int _{E_{T}}q{(\xi (s))}^{m-1}D\rho (\xi (s)) \cdot v(s)\mathrm{d}s. {}\\ \end{array}$$

Here we note

$$\displaystyle{\int _{E_{T}}q{(\xi (s))}^{m}\gamma (\xi (s)) \cdot D\rho (\xi (s))\mathrm{d}s \geq \delta \int _{ E_{T}}q{(\xi (s))}^{m}\vert D\rho (\xi (s))\vert \mathrm{d}s,}$$

and

$$\displaystyle\begin{array}{rcl} & & \int _{E_{T}}q{(\xi (s))}^{m-1}D\rho (\xi (s)) \cdot v(s)\mathrm{d}s {}\\ & & \quad \leq {\left (\int _{E_{T}}q{(\xi (s))}^{m}\vert D\rho (\xi (s)\vert \mathrm{d}s\right )}^{1-\frac{1} {m} }{\left (\int _{E_{ T}}\vert v(s){\vert }^{m}\vert D\rho (\xi (s))\vert \mathrm{d}s\right )}^{ \frac{1} {m} }. {}\\ \end{array}$$

Combining these, we get

$$\displaystyle\begin{array}{rcl} & & q{(\xi (T))}^{m} + \frac{m\delta } {\varepsilon } \int _{E_{T}}q{(\xi (s))}^{m}\vert D\rho (\xi (s))\vert \mathrm{d}s {}\\ & & \quad \leq m{\left (\int _{E_{T}}q{(\xi (s))}^{m}\vert D\rho (\xi (s)\vert \mathrm{d}s\right )}^{1-\frac{1} {m} }{\left (\int _{E_{ T}}\vert v(s){\vert }^{m}\vert D\rho (\xi (s))\vert \mathrm{d}s\right )}^{ \frac{1} {m} }. {}\\ \end{array}$$

Hence,

$$\displaystyle{\frac{\delta } {\varepsilon }{\left (\int _{E_{T}}q{(\xi (s))}^{m}\vert D\rho (\xi (s))\vert \mathrm{d}s\right )}^{ \frac{1} {m} } \leq {\left (\int _{E_{ T}}\vert v(s){\vert }^{m}\vert D\rho (\xi (s))\vert \mathrm{d}s\right )}^{ \frac{1} {m} }}$$

and

$$\displaystyle{q{(\xi (T))}^{m} \leq {\left (\frac{\varepsilon } {\delta }\right )}^{m-1}m\int _{ E_{T}}\vert v(s){\vert }^{m}\vert D\rho (\xi (s))\vert \mathrm{d}s.}$$

Thus, setting \(C_{0} =\| D\rho \|_{{L}^{\infty }({\mathbb{R}}^{n})}\), we find that for any \(T \in \mathbb{R}_{+}\),

$$\displaystyle{ q{(\xi (t))}^{m} \leq {\left (\frac{\varepsilon } {\delta }\right )}^{m-1}mC_{ 0}T\|v\|_{{L}^{\infty }(0,T)}^{m}\ \ \text{ for all }t \in [0,\,T]. }$$
(82)

We henceforth write \(\xi _{\varepsilon }\) for ξ, in order to indicate the dependence on \(\varepsilon\) of ξ, and observe from (82) that for any T > 0,

$$\displaystyle{ \lim _{\varepsilon \rightarrow 0+}\max _{t\in [0,\,T]}\mathrm{dist}(\xi _{\varepsilon }(t),\,\Omega ) = 0. }$$
(83)

Also, (82) ensures that for any T > 0,

$$\displaystyle{ \frac{\delta } {\varepsilon }\|q \circ \xi _{\varepsilon }\|_{{L}^{\infty }(0,T)} \leq {\left (\frac{\delta mC_{0}T} {\varepsilon } \right )}^{ \frac{1} {m} }\|v\|_{{L}^{\infty }(0,T)}. }$$

Sending m → , we find that \((\delta /\varepsilon )\|q \circ \xi _{\varepsilon }\|_{{L}^{\infty }(0,\,T)} \leq \| v\|_{{L}^{\infty }(0,\,T)}\), and moreover

$$\displaystyle{ \frac{\delta } {\varepsilon }\|q \circ \xi _{\varepsilon }\|_{{L}^{\infty }(\mathbb{R}_{+})} \leq \| v\|_{{L}^{\infty }(\mathbb{R}_{+})}. }$$
(84)

We set \(l_{\varepsilon } = (1/\varepsilon )q \circ \xi _{\varepsilon }\). Thanks to (84), we may choose a sequence \(\varepsilon _{j} \rightarrow 0+\) (see Lemma E.1) so that \(l_{\varepsilon _{j}} \rightarrow l\) weakly-star in \({L}^{\infty }(\mathbb{R}_{+})\) as j →  for a function \(l \in {L}^{\infty }(\mathbb{R}_{+})\). It is clear that l(s) ≥ 0 for a.e. \(s \in \mathbb{R}_{+}\).

ODE (81) together with (84) guarantees that \(\{\dot{\xi }_{\varepsilon }\}_{\varepsilon >0}\) is bounded in \({L}^{\infty }(\mathbb{R}_{+})\). Hence, we may assume as well that \(\xi _{\varepsilon _{j}}\) converges locally uniformly on \(\overline{\mathbb{R}}_{+}\) to a function \(\eta \in \mathrm{Lip}(\mathbb{R}_{+})\) as j → . It is then obvious that η(0) = x and the pair (η, l) satisfies

$$\displaystyle{\eta (t) +\int _{ 0}^{t}\big(l(s)\gamma (\eta (s)) - v(s)\big)\mathrm{d}s = x\ \ \text{ for all }t \in \mathbb{R}_{ +},}$$

from which we get

$$\displaystyle{\dot{\eta }(t) + l(t)\gamma (\eta (t)) = v(t)\ \ \text{ for a.e. }t \in \mathbb{R}_{+}.}$$

It follows from (83) that \(\eta (t) \in \overline{\Omega }\) for t ≥ 0.

In order to show that the pair (η, l) is a solution of (80), we need only to prove that for a.e. \(t \in \mathbb{R}_{+}\), l(t) = 0 if η(t) ∈ Ω. Set A = { t ≥ 0 : η(t) ∈ Ω}. It is clear that A is an open subset of [0, ). We can choose a sequence \(\{I_{k}\}_{k\in \mathbb{N}}\) of closed finite intervals of A such that \(A =\bigcup _{k\in \mathbb{N}}I_{k}\). Note that for each \(k \in \mathbb{N}\), the set η(I k ) is a compact subset of Ω and the convergence of \(\{\xi _{\varepsilon _{j}}\}\) to η is uniform on I k . Hence, for any fixed \(k \in \mathbb{N}\), we may choose \(J \in \mathbb{N}\) so that \(\xi _{\varepsilon _{j}}(t) \in \Omega \) for all t ∈ I k and j ≥ J. From this, we have \(q(\xi _{\varepsilon _{j}}(t)) = 0\) for t ∈ I k and j ≥ J. Moreover, in view of the weak-star convergence of \(\{l_{\varepsilon _{j}}\}\), we find that for any \(k \in \mathbb{N}\),

$$\displaystyle{\int _{I_{k}}l(t)\mathrm{d}t =\lim _{j\rightarrow \infty }\int _{I_{k}}\frac{1} {\varepsilon _{j}}q(\xi _{j}(t))\mathrm{d}t = 0,}$$

which yields l(t) = 0 for a.e. t ∈ I k . Since \(A =\bigcup _{k\in \mathbb{N}}I_{k}\), we see that l(t) = 0 a.e. in A. The proof is now complete. □ 

For \(x \in \overline{\Omega }\), let SP(x) denote the set of all triples

$$\displaystyle{(\eta,v,l) \in \mathrm{AC}_{\text{loc}}(\overline{\mathbb{R}}_{+}, {\mathbb{R}}^{n}) \times L_{ \text{loc}}^{1}(\overline{\mathbb{R}}_{ +}, {\mathbb{R}}^{n}) \times L_{ \text{loc}}^{1}(\overline{\mathbb{R}}_{ +})}$$

which satisfies (80). We set \(\mathrm{SP} =\bigcup _{x\in \overline{\Omega }}\mathrm{SP}(x)\).

We remark that for any \(x,y \in \overline{\Omega }\) and \(T \in \mathbb{R}_{+}\), there exists a triple (η, v, l) ∈ SP(x) such that η(T) = y. Indeed, given \(x,y \in \overline{\Omega }\) and \(T \in \mathbb{R}_{+}\), we choose a curve \(\eta \in \mathrm{Lip}([0,\,T],\overline{\Omega })\) (see Lemma 2.1) so that \(\eta (0) = x\), η(T) = y and \(\eta (t) \in \overline{\Omega }\) for all t ∈ [0, T]. We extend the domain of definition of η to \(\overline{\mathbb{R}}_{+}\) by setting η(t) = y for t > T. If we set v(t) = \(\dot{\eta}\)(t) and l(t) = 0 for t ≥ 0, we have (η, v, l) ∈ SP(x), which has the property, η(T) = y.

We note also that problem (80) has the following semi-group property: for any \((x,t) \in \overline{\Omega } \times \mathbb{R}_{+}\) and \((\eta _{1},\,v_{1},\,l_{1}),\,(\eta _{2},\,v_{2},\,l_{2}) \in \mathrm{SP}\), if η 1(0) = x and η 2(0) = η 1(t) hold and if (η, v, l) is defined on \(\overline{\mathbb{R}}_{+}\) by

$$\displaystyle{(\eta (s),\,v(s),\,l(s)) = \left \{\begin{array}{llll} &(\eta _{1}(s),\,v_{1}(s),\,l_{1}(s))\quad &&\text{ for }s \in [0,\,t), \\ &(\eta _{2}(s - t),\,v_{2}(s - t),\,l_{2}(s - t))\quad &&\text{ for }s \in [t,\,\infty ), \end{array} \right.}$$

then (η, v, l) ∈ SP(x).

The following proposition concerns a stability property of sequences of points in SP.

Proposition 5.1.

Let \(\{(\eta _{k},v_{k},l_{k})\}_{k\in \mathbb{N}} \subset \mathrm{SP}\) . Let \(x \in \overline{\Omega }\) and \((w,v,l) \in L_{\mathrm{loc}}(\overline{\mathbb{R}}_{+}, {\mathbb{R}}^{2n+1})\) . Assume that as k →∞,

$$\displaystyle\begin{array}{rcl} & & \eta _{k}(0) \rightarrow x, {}\\ & & (\dot{\eta }_{k},v_{k},l_{k}) \rightarrow (w,v,l)\ \ \ \text{ weakly in }\ {L}^{1}([0,\,T], {\mathbb{R}}^{2n+1}) {}\\ \end{array}$$

for every \(T \in \mathbb{R}_{+}\) . Set

$$\displaystyle{\eta (s) = x +\int _{ 0}^{s}w(r)\mathrm{d}r\ \ \ \text{ for }s \geq 0.}$$

Then (η,v,l) ∈SP (x).

Proof.

For all t > 0 and \(k \in \mathbb{N}\), we have

$$\displaystyle{\eta _{k}(t) =\eta _{k}(0) +\int _{ 0}^{t}\dot{\eta }_{ k}(s)\mathrm{d}s =\eta _{k}(0) +\int _{ 0}^{t}\left (v_{ k}(s) - l_{k}(s)\gamma (\eta _{k}(s))\right )\mathrm{d}s.}$$

First, we observe that as k → ,

$$\displaystyle{\eta _{k}(t) \rightarrow \eta (t)\ \ \ \text{ locally uniformly on }\ \overline{\mathbb{R}}_{+},}$$

and then we get in the limit as k → ,

$$\displaystyle{\eta (t) = x +\int _{ 0}^{t}\left (v(s) - l(s)\gamma (\eta (s))\right )\mathrm{d}s\ \ \ \text{ for all }\ t > 0.}$$

This shows that \(\eta \in \mathrm{AC}_{\text{loc}}(\overline{\mathbb{R}}_{+}, {\mathbb{R}}^{n})\) and

$$\displaystyle{\dot{\eta }(s) + l(s)\gamma (\eta (s)) = v(s)\ \ \ \text{ for a.e. }\ s \in \mathbb{R}_{+}.}$$

It is clear that η(0) = x, \(\eta (s) \in \overline{\Omega }\) for all \(s \in \mathbb{R}_{+}\) and l(s) ≥ 0 for a.e. \(s \in \mathbb{R}_{+}\).

To show that (η, v, l) ∈ SP(x), it remains to prove that for a.e. \(t \in \mathbb{R}_{+}\), l(t) = 0 if η(t) ∈ Ω. As in the last part of the proof of Theorem 5.2, we set A = { t ≥ 0 : η(t) ∈ Ω} and choose a sequence \(\{I_{j}\}_{j\in \mathbb{N}}\) of closed finite intervals of A such that \(A =\bigcup _{j\in \mathbb{N}}I_{j}\). Fix any \(j \in \mathbb{N}\) and choose \(K \in \mathbb{N}\) so that η k (t) ∈ Ω for all t ∈ I j and k ≥ K. From this, we have l k (t) = 0 for a.e. t ∈ I j and k ≥ K. Moreover, in view of the weak convergence of \(\{l_{k}\}\), we find that

$$\displaystyle{\int _{I_{j}}l(t)\mathrm{d}t =\lim _{k\rightarrow \infty }\int _{I_{j}}l_{k}(t)\mathrm{d}t = 0,}$$

which yields l(t) = 0 for a.e. t ∈ I j . Since j is arbitrary, we see that l(t) = 0 a.e. in \(A =\bigcup _{j\in \mathbb{N}}I_{j}\). □ 

Proposition 5.2.

There is a constant C > 0, depending only on Ω and γ, such that for all (η, v, l) ∈SP ,

$$\displaystyle{\vert \dot{\eta }(s)\vert \vee l(s) \leq C\vert v(s)\vert \ \ \text{ for a.e. }s \geq 0.}$$

An immediate consequence of the above proposition is that for (η, v, l) ∈ SP, if \(v \in {L}^{p}(\mathbb{R}_{+},\, {\mathbb{R}}^{n})\) (resp., \(v \in L_{\text{loc}}^{p}(\overline{\mathbb{R}}_{+},\, {\mathbb{R}}^{n})\)), with 1 ≤ p ≤ , then \((\dot{\eta },\,l) \in {L}^{p}(\mathbb{R}_{+},\,{R}^{n+1})\) (resp., \((\dot{\eta },\,l) \in L_{\text{loc}}^{p}(\overline{\mathbb{R}}_{+},\, {\mathbb{R}}^{n+1})\)).

Proof.

Thanks to hypothesis (A4), there is a constant δ 0 > 0 such that ν(x) ⋅γ(x) ≥ δ 0 for x ∈ ∂Ω. Let \(\rho \in {C}^{1}({\mathbb{R}}^{n})\) be a defining function of Ω.

Let \(s \in \mathbb{R}_{+}\) be such that η(s) ∈ ∂Ω, η is differentiable at s, l(s) ≥ 0 and \(\dot{\eta }(s) + l(s)\gamma (\eta (s)) = v(s)\). Observe that the function ρ ∘ η attains a maximum at s. Hence,

$$\displaystyle\begin{array}{rcl} 0& =& \, \frac{\mathrm{d}} {\mathrm{d}s}\rho (\eta (s)) = D\rho (\eta (s)) \cdot \dot{\eta } (s) = \vert D\rho (\eta (s))\vert \nu (\eta (s)) \cdot \dot{\eta } (s) {}\\ & =& \,\vert D\rho (\eta (s))\vert \nu (\eta (s)) \cdot \big (v(s) - l(s)\gamma (\eta (s))\big) {}\\ & \leq & \,\vert D\rho (\eta (s))\vert \big(\nu (\eta (s)) \cdot v(s) - l(s)\delta _{0}\big). {}\\ \end{array}$$

Thus, we get

$$\displaystyle{l(s) \leq \delta _{0}^{-1}\nu (\eta (s)) \cdot v(s) \leq \delta _{ 0}^{-1}\vert v(s)\vert }$$

and

$$\displaystyle\begin{array}{rcl} \vert \dot{\eta }(s)\vert & =& \,\vert v(s) - l(s)\gamma (\eta (s))\vert \leq \vert v(s)\vert + l(s)\vert \|\gamma \|_{\infty,\partial \Omega } {}\\ & \leq & \,(1 +\delta _{ 0}^{-1}\|\gamma \|_{ \infty,\partial \Omega })\vert v(s)\vert, {}\\ \end{array}$$

which completes the proof. □ 

5.2 Value Function I

We define the function \(L \in \mathrm{LSC}(\overline{\Omega } \times {\mathbb{R}}^{n},\,(-\infty,\infty ])\), called the Lagrangian of H, by

$$\displaystyle{L(x,\xi ) =\sup _{p\in {\mathbb{R}}^{n}}\big(\xi \cdot p - H(x,p)\big).}$$

For each x the function ξL(x, ξ) is the convex conjugate of the function pH(x, p). See Appendix A.2 for properties of conjugate convex functions.

We consider the optimal control with the dynamics given by (80), the running cost (L, g) and the pay-off u 0, and its value function V on \(\overline{Q}\), where \(Q = \overline{\Omega } \times \mathbb{R}_{+}\), is given by

$$\displaystyle\begin{array}{rcl} V (x,t)& =& \inf \Big\{\int _{0}^{t}\big(L(\eta (s),-v(s)) + g(\eta (s))l(s)\big)\mathrm{d}s \\ & & \quad + u_{0}(\eta (t))\, :\, (\eta,v,l) \in \mathrm{SP}(x)\Big\}\quad \text{ for }(x,t) \in Q,{}\end{array}$$
(85)

and V (x, 0) = u 0(x) for all \(x \in \overline{\Omega }\).

For t > 0 and \((\eta,v,l) \in \mathrm{SP} =\bigcup _{x\in \overline{\Omega }}\mathrm{SP}(x)\), we write

$$\displaystyle{\mathcal{L}(t,\eta,v,l) =\int _{ 0}^{t}\big(L(\eta (s),-v(s)) + g(\eta (s))l(s)\big)\mathrm{d}s}$$

for notational simplicity, and then formula (85) reads

$$\displaystyle{V (x,t) =\inf \big\{ \mathcal{L}(t,\eta,v,l) + u_{0}(\eta (t))\, :\, (\eta,v,l) \in \mathrm{SP}(x)\big\}.}$$

Under our hypotheses, the Lagrangian L may take the value and, on the other hand, if we set \(C =\min _{x\in \overline{\Omega }}(-H(x,0))\), then we have

$$\displaystyle{L(x,\xi ) \geq C\ \ \ \text{ for all }\ (x,\,\xi ) \in \overline{\Omega } \times {\mathbb{R}}^{n}.}$$

Thus, it is reasonable to interpret

$$\displaystyle{\int _{0}^{t}L(\eta (s),-v(s))\mathrm{d}s = \infty }$$

if the function: sL(η(s), − v(s)) is not integrable, which we adopt here.

It is easily checked as in the proof of Proposition 1.3 that the value function V satisfies the dynamic programming principle: given a point (x, t) ∈ Q and a nonanticipating mapping τ : SP(x) → [0, t], we have

$$\displaystyle\begin{array}{rcl} & & V (x,t) =\inf \big\{ \mathcal{L}(\tau (\alpha ),\alpha ) + V (\eta (\tau (\alpha )),\,t -\tau (\alpha ))\, :\,\alpha = (\eta,v,l) \in \mathrm{SP}(x)\big\}.{}\end{array}$$
(86)

Here a mapping τ : SP(x) → [0, t] is called nonanticipating if τ(α) = τ(β) whenever α(s) = β(s) a.e. in the interval [0, τ(α)].

We here digress to recall the state-constraint problem, whose Bellman equation is given by the Hamilton–Jacobi equation

$$\displaystyle{u_{t}(x,t) + H(x,D_{x}u(x,t)) = 0\ \ \ \mathrm{in}\ \ \Omega \times \mathbb{R}_{+},}$$

and to make a comparison between (ENP) and the state-constraint problem. For \(x \in \overline{\Omega }\) let SC(x) denote the collection of all \(\eta \in \mathrm{AC}_{\text{loc}}(\overline{\mathbb{R}}_{+}, {\mathbb{R}}^{n})\) such that η(0) = x and \(\eta (s) \in \overline{\Omega }\) for all \(s \in \overline{\mathbb{R}}_{+}\). The value function \(\hat{V }\, :\, \overline{\Omega } \times \mathbb{R}_{+} \rightarrow \mathbb{R}\) of the state-constraint problem is given by

$$\displaystyle{\hat{V }(x,t) =\inf \Big\{\int _{ 0}^{t}L(\eta (s),-\dot{\eta }(s))\mathrm{d}s + u_{ 0}(\eta (t))\, :\,\eta \in \mathrm{SC}(x)\Big\}.}$$

Observe that if η ∈ SC(x), with \(x \in \overline{\Omega }\), then \((\eta,\dot{\eta },0) \in \mathrm{SP}(x)\). Hence, we have

$$\displaystyle\begin{array}{rcl} \hat{V }(x,t)& =& \,\inf \big\{\mathcal{L}(t,\eta,\dot{\eta },0) + u_{0}(\eta (t))\, :\,\eta \in \mathrm{SC}(x)\big\} {}\\ & \geq & \,V (x,t)\ \ \ \text{ for all }\ (x,t) \in \overline{\Omega } \times \mathbb{R}_{+}. {}\\ \end{array}$$

Heuristically it is obvious that if g(x) ≈ , then

$$\displaystyle{V (x,t) \approx \hat{ V }(x,t).}$$

In terms of PDE the above state-constraint problem is formulated as follows: the value function \(\hat{V }\) is a unique viscosity solution of

$$\displaystyle{\left \{\begin{array}{lll} &u_{t}(x,t) + H(x,D_{x}u(x,t)) \leq 0&\mathrm{in}\ \ \Omega \times \mathbb{R}_{+}, \\ &u_{t}(x,t) + H(x,D_{x}u(x,t)) \geq 0&\mathrm{in}\ \ \overline{\Omega } \times \mathbb{R}_{+}. \end{array} \right.}$$

See [48] for a proof of this result in this generality. We refer to [17, 55] for state-constraint problems. The corresponding additive eigenvalue problem is to find \((a,v) \in \mathbb{R} \times C(\overline{\Omega })\) such that v is a viscosity solution of

$$\displaystyle{ \left \{\begin{array}{lll} &H(x,Dv(x)) \leq a&\mathrm{in}\ \ \Omega,\\ &H(x, Dv(x)) \geq a &\mathrm{in } \ \ \overline{\Omega }. \end{array} \right. }$$
(87)

We refer to [17, 40, 48] for this eigenvalue problem.

Example 5.2.

We recall (see [48]) that the additive eigenvalue \(\hat{c}\) for (87) is given by

$$\displaystyle{\hat{c} =\inf \{ a \in \mathbb{R}\, :\, (<InternalRef RefID="Equ148">87</InternalRef>)\text{ has a viscosity subsolution }v\},}$$

For a comparison between the Neumann problem and the state-constraint problem, we go back to the situation of Example 3.1. Then it is easy to see that \(\hat{c} = 0\). Thus, we have \({c}^{\#} =\hat{ c} = 0\) if and only if min{g( − 1), g(1)} ≥ 0.

We here continue the above example with some more generality. Let c # and \(\hat{c}\) denote, as above, the eigenvalues of (EVP) and (87), respectively. It is easily seen that if \(\psi \in C(\bar{\Omega })\) is a subsolution of (EVP) with a = c #, then it is also a subsolution of (87) with a = c #, which ensures that \(\hat{c} \leq {c}^{\#}\).

Next, note that the subsolutions of (87) with \(a =\hat{ c}\) are equi-Lipschitz continuous on \(\bar{\Omega }\). That is, there exists a constant M > 0 such that for any subsolution ψ of (87) with \(a =\hat{ c}\), | ψ(x) − ψ(y) | ≤ M | x − y | for all \(x,y \in \bar{ \Omega }\). Let ψ be any subsolution of (87) with \(a =\hat{ c}\), y ∈ ∂Ω and p ∈ D  +  ψ(y). Choose a \(\phi \in {C}^{1}(\bar{\Omega })\) so that (y) = p and ψ − ϕ has a maximum at y. If t > 0 is sufficiently small, then we have y − (y) ∈ Ω and, moreover, ψ(y − (y)) − ψ(y) ≤ ϕ(y − (y)) − ϕ(y). By the last inequality, we deduce that γ(y) ⋅p ≤ M | γ(y) | . Accordingly, we have γ(y) ⋅p ≤ M | γ(y) | for all p ∈ D  +  ψ(y). Thus, we see that if g(x) ≥ M | γ(x) | for all x ∈ ∂Ω, then any subsolution ψ of (87) with \(a =\hat{ c}\) is a subsolution of (EVP) with \(a =\hat{ c}\). This shows that if g(x) ≥ M | γ(x) | for all x ∈ ∂Ω, then \({c}^{\#} \leq \hat{ c}\). As we have already seen above, we have \(\hat{c} \leq {c}^{\#}\), and, therefore, \({c}^{\#} =\hat{ c}\), provided that g(x) ≥ M | γ(x) | for all x ∈ ∂Ω.

Now, assume that \({c}^{\#} =\hat{ c}\) and let \(a = {c}^{\#} =\hat{ c}\). It is easily seen that

$$\displaystyle{\{\psi \,:\,\psi \text{ is a subsolution of (EVP)}\} \subset \{\psi \,:\,\psi \text{ is a subsolution of (87)}\},}$$

which guarantees that d N  ≤ d S on \({\overline{\Omega }}^{2}\), where \(d_{N}(\cdot,y) =\sup \mathcal{F}_{y}^{N}\), \(d_{S}(\cdot,y) =\sup \mathcal{F}_{y}^{S}\), and

$$\displaystyle{\mathcal{F}_{y}^{N}(\text{resp.},\mathcal{F}_{ y}^{S}) =\{\psi -\psi (y)\, :\,\psi \text{ is a subsolution of (EVP)}\ (\,\text{resp., (87)}\,)\}.}$$

Let \(\mathcal{A}_{N}\) and \(\mathcal{A}_{S}\) denote the Aubry sets associated with (EVP) and (87), respectively. That is,

$$\displaystyle\begin{array}{rcl} \mathcal{A}_{N}& =& \{y \in \overline{\Omega }\, :\, d_{N}(\cdot,y)\text{ is a solution of (EVP)}\}, {}\\ \mathcal{A}_{S}& =& \{y \in \overline{\Omega }\, :\, d_{S}(\cdot,y)\text{ is a solution of (87)}\}. {}\\ \end{array}$$

The above inequality and the fact that d N (y, y) = d S (y, y) = 0 for all \(y \in \overline{\Omega }\) imply that \(D_{x}^{-}d_{N}(x,y)\vert _{x=y} \subset D_{x}^{-}d_{S}(x,y)\vert _{x=y}\). From this inclusion, we easily deduce that \(\mathcal{A}_{S} \subset \mathcal{A}_{N}\).

Thus the following proposition holds.

Proposition 5.3.

With the above notation, we have:

  1. (i)

    \(\hat{c} \leq {c}^{\#}\).

  2. (ii)

    If M > 0 is a Lipschitz bound of the subsolutions of (87)with \(a =\hat{ c}\) and g(x) ≥ M|γ(x)| for all x ∈ ∂Ω, then \(\hat{c} = {c}^{\#}\).

  3. (iii)

    If \(\hat{c} = {c}^{\#}\) , then \(d_{N} \leq d_{S}\) on \(\bar{{\Omega }}^{2}\) and \(\mathcal{A}_{S} \subset \mathcal{A}_{N}\) .

5.3 Basic Lemmas

In this subsection we present a proof of the sequential lower semicontinuity of the functional \((\eta,v,l)\mapsto \mathcal{L}(T,\eta,v,l)\) (see Theorem 5.3 below). We will prove an existence result (Theorem 5.6) for the variational problem involving the functional \(\mathcal{L}\) in Sect. 5.4. These results are variations of Tonelli’s theorem in variational problems. For a detailed description of the theory of one-dimensional variational problems, with a central focus on Tonelli’s theorem, we refer to [14].

Lemma 5.1.

For each A > 0 there exists a constant C A ≥ 0 such that

$$\displaystyle{L(x,\,\xi ) \geq A\vert \xi \vert - C_{A}\ \ \ \text{ for all }(x,\xi ) \in \overline{\Omega } \times {\mathbb{R}}^{n}.}$$

Proof.

Fix any A > 0 and observe that

$$\displaystyle\begin{array}{rcl} L(x,\xi )& \geq & \,\max _{p\in \overline{B}_{ A}}(\xi \cdot p - H(x,p)) {}\\ & \geq & \,A\vert \xi \vert +\min _{p\in \overline{B}_{ A}}(-H(x,p))\ \ \ \text{ for all }(x,\xi ) \in \overline{\Omega } \times {\mathbb{R}}^{n}. {}\\ \end{array}$$

Hence, setting \(C_{A} \geq \max _{\overline{\Omega }\times \overline{B}_{ A}}\vert H\vert \), we get

$$\displaystyle{L(x,\xi ) \geq A\vert \xi \vert - C_{A}\ \ \ \text{ for all }(x,\xi ) \in \overline{\Omega } \times {\mathbb{R}}^{n}.}$$

 □ 

Lemma 5.2.

There exist constants δ > 0 and C 0 > 0 such that

$$\displaystyle{L(x,\xi ) \leq C_{0}\ \ \ \text{ for all }(x,\xi ) \in \overline{\Omega } \times B_{\delta }.}$$

Proof.

By the continuity of H, there exists a constant M > 0 such that H(x, 0) ≤ M for all \(x \in \overline{\Omega }\). Also, by the coercivity of H, there exists a constant R > 0 such that H(x, p) > M + 1 for all \((x,p) \in \overline{\Omega } \times \partial B_{R}\). We set δ = R  − 1. Let \((x,\xi ) \in \overline{\Omega } \times B_{\delta }\). Let \(q \in \overline{B}_{R}\) be the minimum point of the function f(p) : = H(x, p) − ξp on \(\overline{B}_{R}\). Noting that f(0) = H(x, 0) ≤ M and f(p) >  − δR + M + 1 = M for all p ∈ ∂B R , we see that q ∈ B R and hence \(\xi \in D_{p}^{-}H(x,q)\), where \(D_{p}^{-}H(x,q)\) denotes the subdifferential at q of the function pH(x, p). Thanks to the convexity of H, this implies (see Theorem B.2) that L(x, ξ) = ξq − H(x, q). Consequently, we get

$$\displaystyle{L(x,\xi ) \leq \delta R +\max _{\overline{\Omega }\times \overline{B}_{ R}}\vert H\vert.}$$

Thus we have the desired inequality with \(C_{0} =\delta R +\max _{\overline{\Omega }\times \overline{B}_{ R}}\vert H\vert \). □ 

For later convenience, we formulate the following lemma, whose proof is left to the reader.

Lemma 5.3.

For each \(i \in \mathbb{N}\) define the function L i on \(\overline{\Omega } \times {\mathbb{R}}^{n}\) by

$$\displaystyle{L_{i}(x,\xi ) =\max _{p\in \overline{B}_{ i}}(\xi \cdot p - H(x,p)).}$$

Then \(L_{i} \in \mathrm{UC}(\overline{\Omega } \times {\mathbb{R}}^{n})\) ,

$$\displaystyle{L_{i}(x,\xi ) \leq L_{i+1}(x,\xi ) \leq L(x,\xi )\ \ \ \text{ for all }\ (x,\xi ) \in \overline{\Omega } \times {\mathbb{R}}^{n}\ \text{ and }\ i \in \mathbb{N},}$$

and for all \((x,\xi ) \in \overline{\Omega } \times {\mathbb{R}}^{n}\) ,

$$\displaystyle{L_{i}(x,\xi ) \rightarrow L(x,\xi )\ \ \ \text{ as }\ i \rightarrow \infty.}$$

The following lemma is a consequence of the Dunford–Pettis theorem.

Lemma 5.4.

Let J = [a, b], with −∞ < a < b < ∞. Let \(\{f_{j}\}_{j\in \mathbb{N}} \subset {L}^{1}(J, {\mathbb{R}}^{m})\) be uniformly integrable in J. That is, for each \(\varepsilon > 0\) , there exists δ > 0 such that for any measurable E ⊂ J and \(j \in \mathbb{N}\) , we have

$$\displaystyle{\int _{E}\vert f_{j}(t)\vert \mathrm{d}t <\varepsilon \ \ \ \mathrm{if}\ \ \vert E\vert <\delta,}$$

where |E| denotes the Lebesgue measure of E. Then \(\{f_{j}\}\) has a subsequence which converges weakly in \({L}^{1}(J, {\mathbb{R}}^{m})\).

See Appendix A.5 for a proof of the above lemma.

Lemma 5.5.

Let J = [0, T] with \(T \in \mathbb{R}_{+}\) ,\((\eta,v) \in {L}^{\infty }(J, {\mathbb{R}}^{n}) \times {L}^{1}(J, {\mathbb{R}}^{n})\) ,\(i \in \mathbb{N}\) and \(\varepsilon > 0\) . Let \(L_{i} \in \mathrm{UC}(\overline{\Omega } \times {\mathbb{R}}^{n})\) be the function defined in Lemma 5.3 . Assume that \(\eta (s) \in \overline{\Omega }\) for all s ∈ J. Then there exists a function \(q \in {L}^{\infty }(J, {\mathbb{R}}^{n})\) such that for a.e. s ∈ J,

$$\displaystyle{q(s) \in \overline{B}_{i}\ \ \ \text{ and }\ \ \ H(\eta (s),q(s)) + L_{i}(\eta (s),-v(s)) \leq -v(s) \cdot q(s) +\varepsilon.}$$

Proof.

Note that for each \((x,\xi ) \in \overline{\Omega } \times {\mathbb{R}}^{n}\) there is a point \(q = q(x,\xi ) \in \overline{B}_{i}\) such that L i (x, ξ) = ξq − H(x, q). By the continuity of the functions H and L i , there exists a constant r = r(x, ξ) > 0 such that

$$\displaystyle{L_{i}(y,z) + H(y,q) \leq z \cdot q +\varepsilon \ \ \ \text{ for all }\ (y,z) \in (\overline{\Omega } \cap B_{r}(x)) \times B_{r}(\xi ).}$$

Hence, as \(\overline{\Omega } \times {\mathbb{R}}^{n}\) is σ-compact, we may choose a sequence \(\{(x_{k},\xi _{k},q_{k},r_{k})\}_{k\in \mathbb{N}} \subset \overline{\Omega } \times {\mathbb{R}}^{n} \times \overline{B}_{i} \times \mathbb{R}_{+}\) such that

$$\displaystyle{\overline{\Omega } \times {\mathbb{R}}^{n} \subset \bigcup _{ k\in \mathbb{N}}B_{r_{k}}(x_{k}) \times B_{r_{k}}(\xi _{k})}$$

and for all \(k \in \mathbb{N}\),

$$\displaystyle{L_{i}(y,z) + H(y,q_{k}) \leq z \cdot q_{k} +\varepsilon \ \ \ \text{ for all }\ (y,z) \in B_{r_{k}}(x_{k}) \times B_{r_{k}}(\xi _{k}).}$$

Now we set \(U_{k} = (\overline{\Omega } \cap B_{r_{k}}(x_{k})) \times B_{r_{k}}(\xi _{k})\) for \(k \in \mathbb{N}\) and define the function \(P\, :\, \overline{\Omega } \times {\mathbb{R}}^{n} \rightarrow {\mathbb{R}}^{n}\) by

$$\displaystyle{P(x,\xi ) = q_{k}\ \ \ \text{ for all }\ (x,\xi ) \in U_{k} \setminus \bigcup _{j<k}U_{j}\ \text{ and all }\ k \in \mathbb{N}.}$$

It is clear that P is Borel measurable in \(\overline{\Omega } \times {\mathbb{R}}^{n}\). Moreover we have \(P(x,\xi ) \in \overline{B}_{i}\) for all \((x,\xi ) \in \overline{\Omega } \times {\mathbb{R}}^{n}\) and

$$\displaystyle{ L_{i}(x,\xi ) + H(x,P(x,\xi )) \leq \xi \cdot P(x,\xi ) +\varepsilon \ \ \ \text{ for all }(x,\xi ) \in \overline{\Omega } \times {\mathbb{R}}^{n}. }$$
(88)

We define the function \(q \in {L}^{\infty }(J,\, {\mathbb{R}}^{n})\) by setting q(s) = P(η(s),  − v(s)). From (88), we see that \(q(s) \in \overline{B}_{i}\) and

$$\displaystyle{L_{i}(\eta (s),-v(s)) + H(\eta (s),q(s)) \leq -v(s) \cdot q(s) +\varepsilon \ \ \ \text{ for a.e. }s \in J.}$$

 □ 

Lemma 5.6.

Let J = [0, T] with \(T \in \mathbb{R}_{+}\) ,\(\varepsilon > 0\) ,\(i \in \mathbb{N}\) ,\(q \in {L}^{\infty }(J, {\mathbb{R}}^{n})\) and \(\eta \in C(J, {\mathbb{R}}^{n})\) such that \(\eta (s) \in \overline{\Omega }\) for all s ∈ J. Assume that \(\|q\|_{{L}^{\infty }(J)} < i\) . Let L i be the function defined in Lemma 5.3 . Then there exists a function \(v \in {L}^{\infty }([0,\,T], {\mathbb{R}}^{n})\) such that

$$\displaystyle{ H(\eta (s),q(s)) + L_{i}(\eta (s),-v(s)) < -v(s) \cdot q(s) +\varepsilon \ \ \ \ \text{ for a.e. }s \in [0,\,T]. }$$
(89)

Before going into the proof we remark that for any \(x \in \overline{\Omega }\) the function L i (x,  ⋅) is the convex conjugate of the function \(\tilde{H}(x,\cdot )\) given by \(\tilde{H}(x,p) = H(x,p)\) if \(p \in \overline{B}_{i}\) and \(\tilde{H}(x,p) = \infty \) otherwise.

Proof.

The same construction as in the proof of Lemma 5.5, with the roles of H and L i being exchanged, yields a measurable function \(v\, :\, [0,\,T] \rightarrow {\mathbb{R}}^{n}\) for which (89) holds. Set \(C =\max _{\overline{\Omega }\times \overline{B}_{ i}}\vert H\vert \) and observe that

$$\displaystyle{L_{i}(x,\xi ) \geq i\vert \xi \vert - C\ \ \ \text{ for all }(x,\xi ) \in \overline{\Omega } \times {\mathbb{R}}^{n}.}$$

We combine this with (89), to get

$$\displaystyle{\varepsilon +\|q\|_{{L}^{\infty }(J)}\vert v(s)\vert > i\vert v(s)\vert - 2C\ \ \ \text{ for a.e. }s \in J.}$$

Hence,

$$\displaystyle{\|v\|_{{L}^{\infty }(J)} \leq \frac{\varepsilon +2C} {i -\| q\|_{{L}^{\infty }(J)}}.}$$

 □ 

The following proposition concerns the lower semicontinuity of the functional

$$\displaystyle{(\eta,v)\mapsto \int _{0}^{T}L(\eta (s),-v(s))\mathrm{d}s.}$$

Theorem 5.3.

Let J = [0, T] with \(T \in \mathbb{R}_{+}\) ,\(\{(\eta _{k},v_{k})\}_{k\in \mathbb{N}} \subset {L}^{\infty }(J, {\mathbb{R}}^{n}) \times {L}^{1}(J, {\mathbb{R}}^{n})\) and \((\eta,v) \in {L}^{\infty }(J, {\mathbb{R}}^{n}) \times {L}^{1}(J, {\mathbb{R}}^{n})\) . Assume that \(\eta _{k}(s) \in \overline{\Omega }\) for all \((s,k) \in J \times \mathbb{N}\) and that as k →∞,

$$\displaystyle\begin{array}{rcl} & & \eta _{k}(s) \rightarrow \eta (s)\quad \qquad \text{ uniformly for }\ s \in J, {}\\ & & v_{k} \rightarrow v\qquad \qquad \qquad \text{ weakly in }\ {L}^{1}(J, {\mathbb{R}}^{n}). {}\\ \end{array}$$

Let ψ be a function in \({L}^{\infty }(J, \mathbb{R})\) such that ψ(s) ≥ 0 for a.e. s ∈ J. Then

$$\displaystyle{ \int _{J}\psi (s)L(\eta (s),-v(s))\mathrm{d}s \leq \liminf _{k\rightarrow \infty }\int _{J}\psi (s)L(\eta _{k}(s),-v_{k}(s))\mathrm{d}s. }$$
(90)

Proof.

Fix any \(i \in \mathbb{N}\). Due to Lemma 5.5, there is a function \(q \in {L}^{\infty }(J, {\mathbb{R}}^{n})\) such that \(q(s) \in \overline{B}_{i}\) and

$$\displaystyle{ H(\eta (s),q(s)) + L_{i}(\eta (s),-v(s)) < -v(s) \cdot q(s) + \frac{1} {i} \ \ \ \text{ for a.e. }\ s \in J. }$$
(91)

Note that for all \(k \in \mathbb{N}\),

$$\displaystyle\begin{array}{rcl} \int _{J}\psi (s)L(\eta _{k}(s),-v_{k}(s))\mathrm{d}s& \geq & \,\int _{J}\psi (s)L_{i}(\eta _{k}(s),-v_{k}(s))\mathrm{d}s {}\\ & \geq & \,\int _{J}\psi (s)[-v_{k}(s) \cdot q(s) - H(\eta _{k}(s),q(s))]\mathrm{d}s, {}\\ \end{array}$$

and

$$\displaystyle\begin{array}{rcl} & & \lim _{k\rightarrow \infty }\int _{J}\psi (s)[-v_{k}(s) \cdot q(s) - H(\eta _{k}(s),q(s))]\mathrm{d}s {}\\ & & \quad =\int _{J}\psi (s)[-v(s) \cdot q(s) - H(\eta (s),q(s))]\mathrm{d}s. {}\\ \end{array}$$

Hence, using (91), we get

$$\displaystyle\begin{array}{rcl} \liminf _{k\rightarrow \infty }\int _{J}\psi (s)L(\eta _{k}(s),-v_{k}(s))\mathrm{d}s& \geq & \int _{J}\psi (s)[-v(s) \cdot q(s) - H(\eta (s),q(s))]\mathrm{d}s {}\\ & \geq & \int _{J}\psi (s)[L_{i}(\eta (s),-v(s)) - 1/i]\mathrm{d}s. {}\\ \end{array}$$

By the monotone convergence theorem, we conclude that (90) holds. □ 

Corollary 5.1.

Under the hypotheses of the above theorem, let \(\{f_{k}\} \subset {L}^{1}(J, \mathbb{R})\) be a sequence of functions converging weakly in \({L}^{1}(J, \mathbb{R})\) to f. Assume furthermore that for all \(k \in \mathbb{N}\) ,

$$\displaystyle{L(\eta _{k}(s),-v_{k}(s)) \leq f_{k}(s)\ \ \ \text{ for a.e. }\ s \in J.}$$

Then

$$\displaystyle{L(\eta (s),-v(s)) \leq f(s)\ \ \ \text{ for a.e. }\ s \in J.}$$

Proof.

Set E = { s ∈ J : L(η(s), − v(s)) > f(s)}. By Theorem 5.3, we deduce that

$$\displaystyle\begin{array}{rcl} 0& \geq & \liminf _{k\rightarrow \infty }\int _{J}\mathbf{1}_{E}(s)[L(\eta _{k}(s),-v_{k}(s)) - f_{k}(s)]\mathrm{d}s {}\\ & \geq & \int _{J}\mathbf{1}_{E}(s)[L(\eta (s),-v(s)) - f(s)]\mathrm{d}s {}\\ & =& \int _{J}[L(\eta (s),-v(s)) - f(s)]_{+}\mathrm{d}s, {}\\ \end{array}$$

where [⋯ ] +  denotes the positive part of [⋯ ]. Thus we see that  L(η(s), − v(s)) ≤ f(s)  for a.e.  s ∈ J. □ 

Lemma 5.7.

Let J = [0, T], with \(T \in \mathbb{R}_{+}\) , and \(q \in C(\overline{\Omega } \times J, {\mathbb{R}}^{n})\) . Let \(x \in \overline{\Omega }\) . Then there exists a triple (η,v,l) ∈SP (x) such that

$$\displaystyle{H(\eta (s),\,q(\eta (s),s)) + L(\eta (s),\,-v(s)) = -v(s) \cdot q(\eta (s),s)\ \ \ \text{ for a.e. }s \in J.}$$

Proof.

Fix \(k \in \mathbb{N}\). Set δ = T ∕ k and s j  = (j − 1)δ for j = 1, 2, , k + 1. We define inductively a sequence \(\{(x_{j},\eta _{j},v_{j},l_{j})\}_{j=1}^{k} \subset \overline{\Omega } \times \mathrm{SP}\). We set x 1 = x and choose a \(\xi _{1} \in {\mathbb{R}}^{n}\) so that

$$\displaystyle{H(x_{1},q(x_{1},0)) + L(x_{1},-\xi _{1}) \leq -\xi _{1} \cdot q(x_{1},0) + 1/k.}$$

Set v 1(s) = ξ 1 for s ≥ 0 and choose a pair \((\eta _{1},l_{1}) \in \mathrm{Lip}(\overline{\mathbb{R}}_{+},\,\overline{\Omega }) \times {L}^{\infty }(\mathbb{R}_{+},\, \mathbb{R})\) so that \((\eta _{1},v_{1},l_{1}) \in \mathrm{SP}(x_{1})\). In fact, Theorem 5.2 guarantees the existence of such a pair.

We argue by induction and now suppose that k ≥ 2 and we are given \((x_{i},\eta _{i},v_{i},l_{i})\) for all i = 1, , j − 1 and some 2 ≤ j ≤ k. Then set x j  = η j − 1(δ), choose a \(\xi _{j} \in {\mathbb{R}}^{n}\) so that

$$\displaystyle{ H(x_{j},q(x_{j},s_{j})) + L(x_{j},-\xi _{j}) \leq -\xi _{j} \cdot q(x_{j},s_{j}) + 1/k, }$$
(92)

set v j (s) = ξ j for s ≥ 0, and select a pair \((\eta _{j},l_{j}) \in \mathrm{Lip}(\overline{\mathbb{R}}_{+},\overline{\Omega }) \times {L}^{\infty }(\mathbb{R}_{+}, \mathbb{R})\) so that \((\eta _{j},v_{j},l_{j}) \in \mathrm{SP}(x_{j})\). Thus, by induction, we can select a sequence \(\{(x_{j},\eta _{j},v_{j},l_{j})\}_{j=1}^{k}\) \(\subset \overline{\Omega } \times \mathrm{SP}\) such that x 1 = η 1(0), \(x_{j} =\eta _{j-1}(\delta ) =\eta _{j}(0)\) for j = 2, , k and for each j = 1, 2, , k, (92) holds with ξ j  = v j (s) for all s ≥ 0. We set \(\alpha _{j} = (\eta _{j},v_{j},l_{j})\) for j = 1, , k.

Note that the choice of \(x_{j},\,\eta _{j},\,v_{j},\,l_{j}\), with j = 1, , k, depends on k, which is not explicit in our notation. We define \(\bar{\alpha }_{k} = (\bar{\eta }_{k},\bar{v}_{k},\bar{l}_{k}) \in \mathrm{SP}(x)\) by setting

$$\displaystyle{\bar{\alpha }_{k}(s) =\alpha _{j}(s - s_{j})\ \ \ \text{ for }\ s \in [s_{j},\,s_{j+1})\ \text{ and }\ j = 1,\ldots,k.}$$

and

$$\displaystyle{\bar{\alpha }_{k}(s) = (\eta _{k}(\delta ),0,0)\ \ \ \text{ for }\ s \geq s_{k+1} = T.}$$

Also, we define \(\bar{x}_{k},\,\bar{q}_{k} \in {L}^{\infty }(J, {\mathbb{R}}^{n})\) by

$$\displaystyle{\bar{x}_{k}(s) = x_{j}\ \ \text{ and }\ \ \bar{q}_{k}(s) = q(x_{j},s_{j})\ \ \ \text{ for }\ s \in [s_{j},\,s_{j+1})\ \text{ and }\ j = 1,\ldots,k.}$$

Now we observe by (92) that for all j = 1, , k,

$$\displaystyle{L(x_{j},-\xi _{j}) \leq \vert \xi _{j}\vert R +\max _{\overline{\Omega }\times \overline{B}_{ R}}\vert H\vert + 1,}$$

where R > 0 is such a constant that \(R \geq \max _{\overline{\Omega }\times J}\vert q\vert \). Combining this estimate with Lemma 5.1, we see that there is a constant C 1 > 0, independent of k, such that

$$\displaystyle{\max _{s\geq 0}\vert \bar{v}_{k}(s)\vert =\max _{1\leq j\leq k}\vert \xi _{j}\vert \leq C_{1}.}$$

By Proposition 5.2, we find a constant C 2 > 0, independent of k, such that \(\|\dot{\bar{\eta }}_{k}\|_{{L}^{\infty }(\mathbb{R}_{+})} \vee \|\bar{ l}_{k}\|_{{L}^{\infty }(\mathbb{R}_{+})} \leq C_{2}\).

We may invoke standard compactness theorems, to find a triple \((\eta,\,v,\,l) \in \mathrm{Lip}(J, {\mathbb{R}}^{n}) \times {L}^{\infty }(J, {\mathbb{R}}^{n+1})\) and a subsequence of \(\{(\bar{\eta }_{k},\bar{v}_{k},\bar{l}_{k})\}_{k\in \mathbb{N}}\), which will be denoted again by the same symbol, so that for every 0 < S < , as k → ,

$$\displaystyle{\bar{\eta }_{k} \rightarrow \eta \ \ \ \text{ uniformly on }\ [0,\,S],}$$
$$\displaystyle{(\dot{\bar{\eta }}_{k},\bar{v}_{k},\bar{l}_{k}) \rightarrow (\dot{\eta },v,l)\ \ \ \text{ weakly-star in }\ {L}^{\infty }([0,\,S], {\mathbb{R}}^{2n+1}).}$$

By Proposition 5.1, we see that (η, v, l) ∈ SP(x). It follows as well that \(\bar{x}_{k}(s) \rightarrow \eta (s)\) and \(\bar{q}_{k}(s) \rightarrow q(\eta (s),s)\) uniformly for s ∈ J as k → .

Now, the inequalities (92), 1 ≤ j ≤ k, can be rewritten as

$$\displaystyle{L(\bar{x}_{k}(s),-\bar{v}_{k}(s)) \leq -\bar{v}_{k}(s) \cdot \bar{ q}_{k}(s) - H(\bar{x}_{k}(s),\bar{q}_{k}(s)) + 1/k\ \ \ \text{ for all }\ s \in [0,\,T).}$$

It is obvious to see that the sequence of functions

$$\displaystyle{-\bar{v}_{k}(s) \cdot q_{k}(s) + 1/k - H(\bar{x}_{k}(s),\bar{q}_{k}(s))}$$

on J converges weakly-star in \({L}^{\infty }(J, \mathbb{R})\) to the function

$$\displaystyle{-v(s) \cdot q(\eta (s),s) - H(\eta (s),q(\eta (s),s)).}$$

Hence, by Corollary 5.1, we conclude that

$$\displaystyle{H(\eta (s),q(\eta (s),s)) + L(\eta (s),-v(s)) \leq -v(s) \cdot q(\eta (s),s)\ \ \ \text{ for a.e. }\ s \in J,}$$

which implies the desired equality. □ 

Theorem 5.4.

Let J = [0, T], with \(T \in \mathbb{R}_{+}\) , and \(\{(\eta _{k},\,v_{k},\,l_{k})\}_{k\in \mathbb{N}} \subset \mathrm{SP}\) . Assume that there is a constant C > 0, independent of \(k \in \mathbb{N}\) , such that

$$\displaystyle{\mathcal{L}(T,\eta _{k},v_{k},l_{k}) \leq C\ \ \mbox{ for all }\ k \in \mathbb{N}.}$$

Then there exists a triple (η, v, l) ∈SP such that

$$\displaystyle{\mathcal{L}(T,\eta,v,l) \leq \liminf _{k\rightarrow \infty }\mathcal{L}(T,\eta _{k},v_{k},l_{k}).}$$

Moreover, there is a subsequence \(\{(\eta _{k_{j}},\,v_{k_{j}},\,l_{k_{j}})\}_{j\in \mathbb{N}}\) of \(\{(\eta _{k},\,v_{k},\,l_{k})\}\) such that as j →∞,

$$\displaystyle\begin{array}{rcl} \eta _{k_{j}}(s) \rightarrow \eta (s)\ \ \ \text{ uniformly on }\ J,& & {}\\ (\dot{\eta }_{k_{j}},\,v_{k_{j}},\,l_{k_{j}}) \rightarrow (\dot{\eta },\,v,\,l)\ \ \ \text{ weakly in }\ {L}^{1}(J, {\mathbb{R}}^{2n+1}).& & {}\\ \end{array}$$

Proof.

We may assume without loss of generality that \(\eta _{k}(t) =\eta _{k}(T)\), v k (t) = 0 and l k (t) = 0 for all t ≥ T and all \(k \in \mathbb{N}\).

According to Proposition 5.2, there is a constant C 0 > 0 such that for any (η, v, l) ∈ SP, \(\vert \dot{\eta }(t)\vert \vee \vert l(t)\vert \leq C_{0}\vert v(t)\vert \) for a.e. t ≥ 0. Note by Lemma 5.1 that for each A > 0 there is a constant C A  > 0 such that L(x, ξ) ≥ A | ξ | − C A for all \((x,\xi ) \in \overline{\Omega } \times {\mathbb{R}}^{n}\). From this lower bound of L, it is obvious that for all \((x,\xi,r) \in \partial \Omega \times {\mathbb{R}}^{n} \times \overline{\mathbb{R}}_{+}\), if r ≤ C 0 | ξ | , then

$$\displaystyle{ L(x,\,\xi ) + g(x)r \geq \left (A - C_{0}\max _{\partial \Omega }\vert g\vert \right )\vert \xi \vert - C_{A}, }$$
(93)

which ensures that there is a constant C 1 > 0 such that for (η, v, l) ∈ SP,

$$\displaystyle{ L(\eta (s),-v(s)) + g(\eta (s))l(s) + C_{1} \geq 0\ \ \mbox{ for a.e. }s \geq 0. }$$
(94)

Set

$$\displaystyle{\Lambda =\liminf _{k\rightarrow \infty }\mathcal{L}(T,\eta _{k},v_{k},l_{k}),}$$

and note by (94) that − C 1 T ≤ Λ ≤ C. We may choose a subsequence \(\{(\eta _{k_{j}},v_{k_{j}},l_{k_{j}})\}_{j\in \mathbb{N}}\) of \(\{(\eta _{k},v_{k},l_{k})\}\) so that

$$\displaystyle{\Lambda =\lim _{j\rightarrow \infty }\mathcal{L}(T,\eta _{k_{j}},v_{k_{j}},l_{k_{j}}).}$$

Using (94), we obtain for any measurable E ⊂ [0, T],

$$\displaystyle\begin{array}{rcl} & & \int _{E}\big(L(\eta _{k}(s),-v_{k}(s)) + g(\eta _{k}(s))l_{k}(s) + C_{1}\big)\,\mathrm{d}s {}\\ & & \quad \leq \int _{0}^{T}\big(L(\eta _{ k}(s),-v_{k}(s)) + g(\eta _{k}(s))l_{k}(s) + C_{1}\big)\,\mathrm{d}s \leq C + C_{1}T. {}\\ \end{array}$$

This together with (93) yields

$$\displaystyle{\left (A - C_{0}\max _{\partial \Omega }\vert g\vert \right )\int _{E}\vert v_{k}(s)\vert \,\mathrm{d}s \leq C_{A}\vert E\vert + C + C_{1}T\ \ \mbox{ for all }A > 0.}$$

This shows that the sequence \(\{v_{k}\}\) is uniformly integrable on [0, T]. Since \(\vert \dot{\eta }_{k}(s)\vert \vee \vert l_{k}(s)\vert \leq C_{0}\vert v_{k}(s)\vert \) for a.e. s ≥ 0 and v k (s) = 0 for all s > T, we see easily that the sequence \(\{(\dot{\eta }_{k},v_{k},l_{k})\}\) is uniformly integrable on \(\overline{\mathbb{R}}_{+}\).

Due to Lemma 5.4, we may assume by reselecting the subsequence \(\{(\eta _{k_{j}},v_{k_{j}},l_{k_{j}})\}\) if necessary that as j → ,

$$\displaystyle{(\dot{\eta }_{k_{j}},v_{k_{j}},l_{k_{j}}) \rightarrow (w,v,l)\ \ \ \text{ weakly in }\ {L}^{1}([0,\,S], {\mathbb{R}}^{2n+1})}$$

for every S > 0 and some \((w,v,l) \in L_{\text{loc}}^{1}(\overline{\mathbb{R}}_{+},\, {\mathbb{R}}^{2n+1})\). We may also assume that \(\eta _{k_{j}}(0) \rightarrow x\) as j →  for some \(x \in \overline{\Omega }\). By Proposition 5.1, if we set \(\eta (s) = x +\int _{ 0}^{s}w(r)\mathrm{d}r\) for s ≥ 0, then (η, v, l) ∈ SP(x) and, as j → ,

$$\displaystyle{\eta _{k_{j}}(s) \rightarrow \eta (s)\ \ \ \text{ locally uniformly on }\ \overline{\mathbb{R}}_{+}.}$$

We apply Theorem 5.3, with the function ψ(s) ≡ 1, to find that

$$\displaystyle{\int _{J}L(\eta (s),-v(s))\mathrm{d}s \leq \liminf _{j\rightarrow \infty }\int _{J}L(\eta _{k_{j}}(s),-v_{k_{j}}(s))\mathrm{d}s.}$$

Consequently, we have

$$\displaystyle{\mathcal{L}(T,\eta,v,l) \leq \liminf _{j\rightarrow \infty }\mathcal{L}(T,\eta _{k_{j}},v_{k_{j}},l_{k_{j}}) = \Lambda,}$$

which completes the proof. □ 

5.4 Value Function II

Theorem 5.5.

Let \(u \in \mathrm{UC}(\overline{\Omega } \times \overline{\mathbb{R}}_{+})\) be the viscosity solution of (ENP) (ID) . Then V = u in \(\overline{\Omega } \times \mathbb{R}_{+}\).

This is a version of classical observations on the value functions in optimal control, and, in this regard, we refer for instance to [43, 45]. The above theorem has been established in [39]. The above theorem gives a variational formula for the unique solution of (ENP)–(ID). This variational formula is sometimes called the Lax–Oleinik formula.

For the proof of Theorem 5.5, we need the following three lemmas.

Lemma 5.8.

Let \(U \subset {\mathbb{R}}^{n}\) be an open set and J = [a, b] a finite subinterval of \(\overline{\mathbb{R}}_{+}\) . Let \(\psi \in {C}^{1}((\overline{U} \cap \overline{\Omega }) \times J)\) and assume that

$$\displaystyle{ \psi _{t}(x,t) + H(x,D_{x}\psi (x,t)) \leq 0\text{ for all }\ (x,t) \in (U \cap \Omega ) \times J, }$$
(95)
$$\displaystyle{ \frac{\partial \psi } {\partial \gamma }(x,t) \leq g(x)\text{ for all }\ (x,t) \in (U \cap \partial \Omega ) \times J, }$$
(96)
$$\displaystyle{ \psi (x,t) \leq V (x,t)\text{ for all }\ (x,t) \in (\partial U \cap \overline{\Omega }) \times J, }$$
(97)
$$\displaystyle{ \psi (x,a) \leq V (x,a)\text{ for all }\ x \in \overline{U} \cap \overline{\Omega }. }$$
(98)

Then ψ ≤ V in \((U \cap \overline{\Omega }) \times J\).

We note that the following inclusion holds: \(\partial (U \cap \overline{\Omega }) \subset [\partial U \cap \overline{\Omega }] \cup (U \cap \partial \Omega )\).

Proof.

Let \((x,t) \in (U \cap \overline{\Omega }) \times J\). Define the mapping τ : SP(x) → [0, t − a] by

$$\displaystyle{\tau (\eta,v,l) =\inf \{ s \geq 0\, :\,\eta (s)\not\in U\} \wedge (t - a).}$$

It is clear that τ is nonanticipating. Let α = (η, v, l) ∈ SP(x), and observe that η(s) ∈ U for all s ∈ [0, τ(α)) and that η(τ(α)) ∈ ∂U if τ(α) < t − a. In particular, we find from (97) and (98) that

$$\displaystyle{ \psi (\eta (\tau (\alpha )),t -\tau (\alpha )) \leq V (\eta (\tau (\alpha )),t -\tau (\alpha )). }$$
(99)

Fix any α = (η, v, l) ∈ SP(x). Note that

$$\displaystyle\begin{array}{rcl} & & \psi (\eta (\tau (\alpha )),t -\tau (\alpha )) -\psi (x,t) {}\\ & & \quad =\int _{ 0}^{\tau (\alpha )} \frac{\mathrm{d}} {\mathrm{d}s}\psi (\eta (s),t - s)\mathrm{d}s {}\\ & & \quad =\int _{ 0}^{\tau (\alpha )}\big(D_{ x}\psi (\eta (s),t - s) \cdot \dot{\eta } (s) -\psi _{t}(\eta (s),t - s)\big)\mathrm{d}s {}\\ & & \quad =\int _{ 0}^{\tau (\alpha )}\big(D_{ x}\psi (\eta (s),t - s) \cdot (v(s) - l(s)\gamma (\eta (s))) -\psi _{t}(\eta (s),t - s)\big)\mathrm{d}s. {}\\ \end{array}$$

Now, using (95), (96) and (99), we get

$$\displaystyle\begin{array}{rcl} & & \psi (x,t) - V (\eta (\tau (\alpha )),t -\tau (\alpha )) {}\\ & & \quad \leq \int _{0}^{\tau (\alpha )}\big(-D_{ x}\psi (\eta (s),t - s) \cdot v(s) + l(s)D_{x}\psi (\eta (s)) \cdot \gamma (\eta (s)) {}\\ & & \qquad +\psi _{t}(\eta (s),t - s)\big)\mathrm{d}s {}\\ & & \quad \leq \int _{0}^{\tau (\alpha )}\big(H(\eta (s),D_{ x}\psi (\eta (s),t - s)) + L(\eta (s),-v(s)) + l(s)g(\eta (s)) {}\\ & & \qquad +\psi _{t}(\eta (s),t - s)\big)\mathrm{d}s {}\\ & & \leq \mathcal{L}(\tau (\alpha ),\eta,v,l), {}\\ \end{array}$$

which immediately shows that

$$\displaystyle{\psi (x,t) \leq \inf \left (\mathcal{L}(\tau (\alpha ),\eta,v,l) + V (\eta (\tau (\alpha )),t -\tau (\alpha ))\right ),}$$

where the infimum is taken over all α = (η, v, l) ∈ SP(x). Thus, by (86), we get  ψ(x, t) ≤ V (x, t). □ 

Lemma 5.9.

For any \(\varepsilon > 0\) there is a constant \(C_{\varepsilon } > 0\) such that \(V (x,t) \geq u_{0}(x) -\varepsilon -C_{\varepsilon }t\) for \((x,t) \in \overline{Q}\).

Proof.

Fix any \(\varepsilon > 0\). According to the proof of Theorem 3.2, there are a function \(f \in {C}^{1}(\overline{\Omega })\) and a constant C > 0 such that if we set ψ(x, t) = f(x) − Ct for \((x,t) \in \overline{Q}\), then ψ is a classical subsolution of (ENP) and \(u_{0}(x) \geq f(x) \geq u_{0}(x)-\varepsilon\) for all \(x \in \overline{\Omega }\).

We apply Lemma 5.8, with \(U = {\mathbb{R}}^{n}\), a = 0, arbitrary b > 0, to obtain

$$\displaystyle{V (x,t) \geq \psi (x,t) \geq -\varepsilon + u_{0}(x) -\mathit{Ct}\ \ \ \text{ for all }\ (x,t) \in Q,}$$

which completes the proof. □ 

Lemma 5.10.

There is a constant C > 0 such that V (x,t) ≤ u 0 (x) + Ct for (x,t) ∈ Q.

Proof.

Let (x, t) ∈ Q. Set η(s) = x, v(s) = 0 and l(s) = 0 for s ≥ 0. Then (η, v, l) ∈ SP(x). Hence, we have

$$\displaystyle{V (x,t) \leq u_{0}(x) +\int _{ 0}^{t}L(x,0)\mathrm{d}s = u_{ 0}(x) + tL(x,0) \leq u_{0}(x) - t\min _{p\in {\mathbb{R}}^{n}}H(x,p).}$$

Setting \(C = -\min _{\overline{\Omega }\times {\mathbb{R}}^{n}}H\), we get V (x, t) ≤ u 0(x) + Ct. □ 

Proof (Theorem 5.5). 

By Lemmas 5.9 and 5.10, there is a constant C > 0 and for each \(\varepsilon > 0\) a constant \(C_{\varepsilon } > 0\) such that

$$\displaystyle{-\varepsilon - C_{\varepsilon }t \leq V (x,t) - u_{0}(x) \leq \mathit{Ct}\ \ \text{ for all }(x,t) \in Q.}$$

This shows that V is locally bounded on \(\overline{Q}\) and that

$$\displaystyle{\lim _{t\rightarrow 0+}V (x,t) = u_{0}(x)\ \ \text{ uniformly for }x \in \overline{\Omega }.}$$

In particular, we have \(V _{{\ast}}(x,0) = {V }^{{\ast}}(x,0) = u_{0}(x)\) for all \(x \in \overline{\Omega }\).

We next prove that V is a subsolution of (ENP). Let \((\hat{x},\hat{t}) \in Q\) and \(\phi \in {C}^{1}(\overline{Q})\). Assume that V  ∗  − ϕ attains a strict maximum at \((\hat{x},\hat{t})\). We want to show that if \(\hat{x} \in \Omega \), then

$$\displaystyle{\phi _{t}(\hat{x},\hat{t}) + H(\hat{x},D_{x}\phi (\hat{x},\hat{t})) \leq 0,}$$

and if \(\hat{x} \in \partial \Omega \), then either

$$\displaystyle{\phi _{t}(\hat{x},\hat{t}) + H(\hat{x},D_{x}\phi (\hat{x},\hat{t})) \leq 0\ \ \text{ or }\ \ \gamma (\hat{x}) \cdot D_{x}\phi (\hat{x},\hat{t}) \leq g(\hat{x}).}$$

We argue by contradiction and thus suppose that

$$\displaystyle{\phi _{t}(\hat{x},\hat{t}) + H(\hat{x},D_{x}\phi (\hat{x},\hat{t})) > 0}$$

and furthermore

$$\displaystyle{\gamma (\hat{x}) \cdot D_{x}\phi (\hat{x},\hat{t}) > g(\hat{x})\ \ \ \mathrm{if}\ \ \hat{x} \in \partial \Omega.}$$

By continuity, we may choose a constant \(r \in (0,\hat{t}\,)\) so that

$$\displaystyle{ \phi _{t}(x,t) + H(x,D_{x}\phi (x,t)) > 0\ \ \ \text{ for all }\ (x,t) \in (\overline{B}_{r}(\hat{x}) \cap \overline{\Omega }) \times \hat{ J}, }$$
(100)

where \(\hat{J} = [\hat{t} - r,\,\hat{t} + r]\), and

$$\displaystyle{ \gamma (x) \cdot D_{x}\phi (x,t) > g(x)\ \ \ \text{ for all }\ (x,t) \in (\overline{B}_{r}(\hat{x}) \cap \partial \Omega ) \times \hat{ J}. }$$
(101)

(Of course, if \(\hat{x} \in \Omega \), we can choose r so that \(\overline{B}_{r}(\hat{x}) \cap \partial \Omega = \varnothing \).)

We may assume that \(({V }^{{\ast}}-\phi )(\hat{x},\hat{t}) = 0\). Set

$$\displaystyle{B = \left (\big(\partial B_{r}(\hat{x}) \cap \overline{\Omega }\big) \times \hat{ J}\right ) \cup \left (\big(B_{r}(\hat{x}) \cap \overline{\Omega }\big) \times \{\hat{ t} - r\}\right ),}$$

and \(m = -\max _{B}({V }^{{\ast}}-\phi )\). Note that m > 0 and V (x, t) ≤ ϕ(x, t) − m for (x, t) ∈ B.

We set \(\varepsilon = r/2\). In view of the definition of V  ∗ , we may choose a point \((\bar{x},\bar{t}) \in \overline{\Omega } \cap B_{\varepsilon }(\hat{x}) \times (\hat{t}-\varepsilon,\,\hat{t}+\varepsilon )\) so that \((V -\phi )(\bar{x},\bar{t}) > -m\). Set \(a =\bar{ t} -\hat{ t} + r\), and note that \(a >\varepsilon\) and \(\mathrm{dist}(\bar{x},\partial B_{r}(\hat{x})) >\varepsilon\). For each \(\alpha = (\eta,v,l) \in \mathrm{SP}(\bar{x})\) we set

$$\displaystyle{S(\alpha ) =\{ s \geq 0\, :\,\eta (s) \in \partial B_{r}(\hat{x})\}\ \ \text{ and }\ \ \tau = a \wedge \inf S(\alpha ).}$$

Clearly, the mapping \(\tau \,:\, \mathrm{SP}(\bar{x}) \rightarrow [0,a]\) is nonanticipating. Observe also that if τ(α) < a, then \(\eta (\tau (\alpha )) \in \partial B_{r}(\hat{x})\) or, otherwise, \(\bar{t} -\tau (\alpha ) =\bar{ t} - a =\hat{ t} - r\). That is, we have

$$\displaystyle{ (\eta (\tau (\alpha )),\bar{t} -\tau (\alpha )) \in B\ \ \ \text{ for all }\ \alpha = (\eta,v,l) \in \mathrm{SP}(\bar{x}). }$$
(102)

Note as well that \((\eta (s),\bar{t} - s) \in \overline{B}_{r}(\hat{x}) \times \hat{ J}\) for all s ∈ [0, τ(α)].

We apply Lemma 5.7, with J = [0, a] and the function \(q(x,s) = D\phi (x,\bar{t} - s)\), to find a triple \(\alpha = (\eta,v,l) \in \mathrm{SP}(\bar{x})\) such that for a.e. s ∈ [0, a],

$$\displaystyle{ H(\eta (s),D_{x}\phi (\eta (s),\bar{t} - s)) + L(\eta (s),-v(s)) \leq -v(s) \cdot D_{x}\phi (\eta (s),\bar{t} - s) }$$
(103)

For this α, we write τ = τ(α) for simplicity of notation. Using (102), by the dynamic programming principle, we have

$$\displaystyle\begin{array}{rcl} \phi (\bar{x},\bar{t})& <& \,V (\bar{x},\bar{t}) + m {}\\ & \leq & \,\mathcal{L}(\tau,\eta,v,l) + V (\tau,\bar{t}-\tau ) + m {}\\ & \leq & \,\mathcal{L}(\tau,\eta,v,l) +\phi (\eta (\tau ),\bar{t}-\tau ). {}\\ \end{array}$$

Hence, we obtain

$$\displaystyle\begin{array}{rcl} 0& <& \int _{0}^{\tau }\big(L(\eta (s),-v(s)) + g(\eta (s))l(s) + \frac{\mathrm{d}} {\mathrm{d}s}\phi (\eta (s),\bar{t} - s)\big)\mathrm{d}s {}\\ & \leq & \int _{0}^{\tau }\big(L(\eta (s),-v(s)) + g(\eta (s))l(s) + D_{ x}\phi (\eta (s),\bar{t} - s) \cdot \dot{\eta } (s) -\phi _{t}(\eta (s),\bar{t} - s)\big)\mathrm{d}s {}\\ & \leq & \int _{0}^{\tau }\big(L(\eta (s),-v(s)) + g(\eta (s))l(s) {}\\ & & \quad + D_{x}\phi (\eta (s),\bar{t} - s) \cdot (v(s) - l(s)\gamma (\eta (s)) -\phi _{t}(\eta (s),\bar{t} - s)\big)\mathrm{d}s. {}\\ \end{array}$$

Now, using (103), (100) and (101), we get

$$\displaystyle\begin{array}{rcl} 0& <& \int _{0}^{\tau }\big(-H(\eta (s),D_{ x}\phi (\eta (s),\bar{t} - s)) + g(\eta (s))l(s) {}\\ & & \quad - l(s)D_{x}\phi (\eta (s),\bar{t} - s) \cdot \gamma (\eta (s)) -\phi _{t}(\eta (s),\bar{t} - s)\big)\mathrm{d}s {}\\ & <& \int _{0}^{\tau }l(s)\big(g(\eta (s)) -\gamma (\eta (s)) \cdot D_{ x}\phi (\eta (s),\bar{t} - s)\big)\mathrm{d}s \leq 0, {}\\ \end{array}$$

which is a contradiction. We thus conclude that V is a viscosity subsolution of (ENP).

Now, we turn to the proof of the supersolution property of V. Let \(\phi \in {C}^{1}(\overline{Q})\) and \((\hat{x},\hat{t}) \in \overline{\Omega } \times \mathbb{R}_{+}\). Assume that V  ∗  − ϕ attains a strict minimum at \((\hat{x},\hat{t})\). As usual, we assume furthermore that \(\min _{\overline{Q}}(V _{{\ast}}-\phi ) = 0\).

We need to show that if \(\hat{x} \in \Omega \), then

$$\displaystyle{\phi _{t}(\hat{x},\hat{t}) + H(\hat{x},D_{x}\phi (\hat{x},\hat{t})) \geq 0,}$$

and if \(\hat{x} \in \partial \Omega \), then

$$\displaystyle{\phi _{t}(\hat{x},\hat{t}) + H(\hat{x},D_{x}\phi (\hat{x},\hat{t})) \geq 0\ \ \text{ or }\ \ \gamma (\hat{x}) \cdot D_{x}\phi (\hat{x},\hat{t}) \geq g(\hat{x}).}$$

We argue by contradiction and hence suppose that this were not the case. That is, we suppose that

$$\displaystyle{\phi _{t}(\hat{x},\hat{t}) + H(\hat{x},D_{x}\phi (\hat{x},\hat{t})) < 0,}$$

and moreover

$$\displaystyle{\gamma (\hat{x}) \cdot D_{x}\phi (\hat{x},\hat{t}) < g(\hat{x})\ \ \ \mathrm{if}\ \ \hat{x} \in \partial \Omega.}$$

We may choose a constant \(r \in (0,\,\hat{t})\) so that

$$\displaystyle{ \phi _{t}(x,t) + H(x,D_{x}\phi (x,t)) < 0\ \ \ \text{ for all }\ (x,t) \in (B_{r}(\hat{x}) \cap \overline{\Omega }) \times \hat{ J}, }$$

where \(\hat{J} = [\hat{t} - r,\,\hat{t} + r]\), and

$$\displaystyle{ \gamma (x) \cdot D_{x}\phi (x,t) < g(x)\ \ \ \text{ for all }\ (x,t) \in (B_{r}(\hat{x}) \cap \partial \Omega ) \times \hat{ J}. }$$
(104)

We set

$$\displaystyle{R = \left ((\partial B_{r}(\hat{x}) \cap \overline{\Omega }) \times \hat{ J}\right ) \cup \left ((B_{r}(\hat{x}) \cap \overline{\Omega }) \times \{\hat{ t} - r\}\right )\ \text{ and }\ m =\min _{R}(V _{{\ast}}-\phi ),}$$

and define the function \(\psi \in {C}^{1}((\overline{B}_{r}(\hat{x}) \cap \overline{\Omega }) \times \hat{ J})\) by ψ(x, t) = ϕ(x, t) + m. Note that m > 0,  \(\inf _{(B_{ r}(\hat{x})\cap \overline{\Omega })\times \hat{J}}(V _{{\ast}}-\psi ) = -m < 0\)  and  V (x, t) ≥ ψ(x, t)  for all (x, t) ∈ R. Observe moreover that

$$\displaystyle\begin{array}{rcl} & & \psi _{t}(x,t) + H(x,D_{x}\psi (x,t)) < 0\qquad \qquad \text{ for all }\ (x,t) \in (B_{r}(\hat{x}) \cap \Omega ) \times \hat{ J} {}\\ & & \frac{\partial \psi } {\partial \gamma }(x,t) < g(x)\quad \qquad \qquad \qquad \quad \qquad \text{ for all }\ (x,t) \in (B_{r}(\hat{x}) \cap \partial \Omega ) \times \hat{ J}. {}\\ \end{array}$$

We invoke Lemma 5.8, to find that ψ ≤ V in \((B_{r}(\hat{x}) \cap \overline{\Omega }) \times \hat{ J}\). This means that \(\inf _{ (B_{r}(\hat{x})\cap \overline{\Omega })\times \hat{J}}(V _{{\ast}}-\psi ) \geq 0\). This contradiction shows that V is a viscosity supersolution of (ENP).

We apply Theorem 3.1 to V  ∗ , u and V  ∗ , to obtain \({V }^{{\ast}}\leq u \leq V _{{\ast}}\) in \(\overline{Q}\), from which we conclude that u = V in \(\overline{Q}\). □ 

Our control problem always has an optimal “control” in SP:

Theorem 5.6.

Let \((x,t) \in \overline{\Omega } \times \mathbb{R}_{+}\) . Then there exists a triple (η, v, l) ∈SP (x) such that

$$\displaystyle{V (x,t) = \mathcal{L}(t,\eta,v,l) + u_{0}(\eta (t)).}$$

If, in addition, \(V \in \mathrm{Lip}(\overline{\Omega } \times J, \mathbb{R})\) , with J being an interval of [0, t], then the triple (η, v, l), restricted to \(\tilde{J}_{t} :=\{ s \in [0,\,t]\, :\, t - s \in J\}\) , belongs to \(\mathrm{Lip}(\tilde{J}_{t}, {\mathbb{R}}^{n}) \times {L}^{\infty }(\tilde{J}_{t}, {\mathbb{R}}^{n+1})\).

Proof.

We may choose a sequence \(\{(\eta _{k},\,v_{k},\,l_{k})\} \subset \mathrm{SP}(x)\) such that

$$\displaystyle{V (x,t) =\lim _{k\rightarrow \infty }\mathcal{L}(t,\eta _{k},v_{k},l_{k}) + u_{0}(\eta _{k}(t)).}$$

In view of Theorem 5.4, we may assume by replacing the sequence \(\{(\eta _{k},\,v_{k},\,l_{k})\}\) by a subsequence if needed that for some \((\eta,v,l) \in \mathrm{SP}(x)\), η k (s) → η(s) uniformly on [0, t] as k →  and

$$\displaystyle{ \mathcal{L}(t,\eta,v,l) \leq \liminf _{k\rightarrow \infty }\mathcal{L}(t,\eta _{k},v_{k},l_{k}). }$$

It is then easy to see that

$$\displaystyle{ V (x,t) = \mathcal{L}(r,\eta,v,l) + u_{0}(\eta (t)). }$$
(105)

Note by (105) that for all r ∈ (0, t),

$$\displaystyle{V (x,t) \geq \mathcal{L}(r,\eta,v,l) + V (\eta (r),t - r),}$$

which yields together with the dynamic programming principle

$$\displaystyle{ V (x,t) = \mathcal{L}(r,\eta,v,l) + V (\eta (r),t - r) }$$
(106)

for all r ∈ (0, t).

Now, we assume that \(V \in \mathrm{Lip}(\overline{\Omega } \times J)\), where J ⊂ [0, t] is an interval. Observe by (106) that for a.e. \(r \in \tilde{ J}_{t}\),

$$\displaystyle\begin{array}{rcl} L(\eta (r),-v(r)) + l(r)g(\eta (r))& =& \,\lim _{\varepsilon \rightarrow 0}\frac{V (\eta (r),t - r) - V (\eta (r+\varepsilon ),t - r-\varepsilon )} {\varepsilon } {}\\ & \leq & \,M{(\vert \dot{\eta }(r){\vert }^{2} + 1)}^{1/2} \leq M(\vert \dot{\eta }(r)\vert + 1), {}\\ \end{array}$$

where M > 0 is a Lipschitz bound of the function V on \(\overline{\Omega } \times J\). Let C > 0 be the constant from Proposition 5.2, so that \(\vert \dot{\eta }(s)\vert \vee l(s) \leq C\vert v(s)\vert \) for a.e. s ≥ 0. By Lemma 5.1, for each A > 0, we may choose a constant C A  > 0 so that L(y, ξ) ≥ A | ξ | − C A for \((y,\,\xi ) \in \overline{\Omega } \times {\mathbb{R}}^{n}\). Accordingly, for any A > 0, we get

$$\displaystyle\begin{array}{rcl} A\vert v(r)\vert & \leq & \,L(\eta (r),-v(r)) + C_{A} \leq -l(r)g(\eta (r)) + M(\vert \dot{\eta }(r)\vert + 1) + C_{A} {}\\ & \leq & \,C(\|g\|_{\infty,\partial \Omega } + M)\vert v(r)\vert + M + C_{A}\ \ \ \text{ for a.e. }\ r \in \tilde{ J}_{t}. {}\\ \end{array}$$

This implies that \(v \in {L}^{\infty }(\tilde{J}_{t}, {\mathbb{R}}^{n})\) and moreover that \(\eta \in \mathrm{Lip}(\tilde{J}_{t}, {\mathbb{R}}^{n})\) and \(l \in {L}^{\infty }(\tilde{J}_{t}, \mathbb{R})\). The proof is complete. □ 

Corollary 5.2.

Let \(u \in \mathrm{Lip}(\overline{\Omega })\) be a viscosity solution of (SNP) and \(x \in \overline{\Omega }\) . Then there exists a (η,v,l) ∈SP (x) such that for all t > 0,

$$\displaystyle{ u(x) - u(\eta (t)) = \mathcal{L}(t,\eta,v,l). }$$
(107)

Proof.

Note that the function u(x), as a function of (x, t), is a viscosity solution of (ENP). In view of Theorem 5.6, we may choose a sequence \(\{(\eta _{j},v_{j},l_{j})\}_{j\in \mathbb{N}}\) so that η 1(0) = x, η j + 1(0) = η j (1) for all \(j \in \mathbb{N}\) and

$$\displaystyle{u(\eta _{j}(0)) - u(\eta _{j}(1)) = \mathcal{L}(1,\eta _{j},v_{j},l_{j})\ \ \ \text{ for all }\ j \in \mathbb{N}.}$$

We define (η, v, l) ∈ SP(x) by

$$\displaystyle{(\eta (s),v(s),l(s)) = (\eta _{j}(s - j + 1),v_{j}(s - j + 1),l_{j}(s - j + 1))}$$

for all s ∈ [j − 1, j) and \(j \in \mathbb{N}\). By using the dynamic programming principle, we see that (107) holds for all t > 0. □ 

5.5 Distance-Like Function d

We assume throughout this subsection that (A8) holds, and discuss a few aspects of weak KAM theory related to (SNP).

Proposition 5.4.

We have the variational formula for the function d introduced in Sect.  4.1:for all   \(x,y \in \overline{\Omega }\) ,

$$\displaystyle{ d(x,y) =\inf \big\{ \mathcal{L}(t,\eta,v,l)\, :\, t > 0,\ (\eta,v,l) \in \mathrm{SP}(x)\ \ \text{ such that }\ \ \eta (t) = y\big\}. }$$
(108)

We use the following lemma for the proof of the above proposition.

Lemma 5.11.

Let \(u_{0} \in C(\overline{\Omega })\) and \(u \in \mathrm{UC}(\overline{Q})\) be the viscosity solution of (ENP) (ID) . Set

$$\displaystyle{v(x,t) =\inf _{r>0}u(x,t + r)\ \ \ \text{ for }\ x \in \overline{Q}.}$$

Then \(v \in \mathrm{UC}(\overline{Q})\) and it is a viscosity solution of (ENP) . Moreover, for each t > 0, the function v(⋅,t) is a viscosity subsolution of (SNP).

Proof.

By assumption (A8), there is a viscosity subsolution ψ of (SNP). Note that the function (x, t)↦ψ(x) is a viscosity subsolution of (ENP) as well.

We may assume by adding a constant to ψ if needed that ψ ≤ u 0 in \(\overline{\Omega }\). By Theorem 3.1, we have  u(x, t) ≥ ψ(x) >  −  for all \((x,t) \in \overline{Q}\). Since \(u \in \mathrm{UC}(\overline{Q})\), we see immediately that \(v \in \mathrm{UC}(\overline{Q})\). Applying a version for (ENP) of Theorem 4.4, which can be proved based on Theorem D.2, to the collection of viscosity solutions (x, t)↦u(x, t + r), with r > 0, of (ENP), we find that v is a viscosity subsolution of (ENP). Also, by Proposition 1.10 (its version for supersolutions), we see that v is a viscosity supersolution of (ENP). Thus, the function v is a viscosity solution of (ENP).

Next, note that for each \(x \in \overline{\Omega }\), the function v(x,  ⋅) is nondecreasing in \(\mathbb{R}_{+}\). Let \((\hat{x},\hat{t}) \in Q\) and \(\phi \in {C}^{1}(\overline{\Omega })\). Assume that the function \(\overline{\Omega } \ni x\mapsto v(x,\hat{t}) -\phi (x)\) attains a strict maximum at \(\hat{x}\). Let α > 0 and consider the function

$$\displaystyle{v(x,t) -\phi (x) -\alpha {(t -\hat{ t})}^{2}\ \ \ \mathrm{on}\ \ \overline{\Omega } \times [0,\,\hat{t} + 1].}$$

Let \((x_{\alpha },t_{\alpha })\) be a maximum point of this function. It is easily seen that \((x_{\alpha },t_{\alpha }) \rightarrow (\hat{x},\hat{t})\) as α → . For sufficiently large α, we have t α  > 0 and either

$$\displaystyle{x_{\alpha } \in \partial \Omega \ \ \text{ and }\ \ \gamma (x_{\alpha }) \cdot D\phi (x_{\alpha }) \leq g(x_{\alpha }),}$$

or

$$\displaystyle{2\alpha (t_{\alpha } -\hat{ t}) + H(x_{\alpha },D\phi (x_{\alpha })) \leq 0.}$$

By the monotonicity of v(x, t) in t, we see easily that \(2\alpha (t_{\alpha } -\hat{ t}) \geq 0\). Hence, sending α → , we conclude that the function \(v(\cdot,\hat{t})\) is a viscosity subsolution of (SNP). □ 

Proof (Proposition 5.4). 

We write W(x, y) for the right hand side of (108).

Fix any \(y \in \overline{\Omega }\). For each \(k \in \mathbb{N}\) let \(u_{k} \in \mathrm{Lip}(\overline{Q})\) be the unique viscosity solution of (ENP)–(ID), with u 0 defined by u 0(x) = k | x − y | . By Theorem 5.5, we have the formula:

$$\displaystyle{u_{k}(x,t) =\inf \big\{ \mathcal{L}(t,\eta,v,l) + k\vert \eta (t) - y\vert \, :\, (\eta,v,l) \in \mathrm{SP}(x)\big\}.}$$

It is then easy to see that

$$\displaystyle{ \inf _{t>0}u_{k}(x,t) \leq W(x,y)\ \ \ \text{ for all }(x,k) \in \overline{\Omega } \times \mathbb{N}. }$$
(109)

Since \(d(\cdot,y) \in \mathrm{Lip}(\overline{\Omega })\), if k is sufficiently large, say k ≥ K, we have  d( ⋅, y) ≤ k | x − y |  for all \(x \in \overline{\Omega }\). Noting that the function \((x,t)\mapsto d(x,y)\) is a viscosity subsolution of (ENP) and applying Theorem 3.1, we get  d(x, y) ≤ u k (x, t)  for all (x, t) ∈ Q if k ≥ K. Combining this and (109), we find that d(x, y) ≤ W(x, y) for all \(x \in \overline{\Omega }\).

Next, we give an upper bound on W. According to Lemma 2.1, there exist a constant C 1 > 0 and a function \(\tau \,:\, \overline{\Omega } \rightarrow \overline{\mathbb{R}}_{+}\) such that τ(x) ≤ C 1 | x − y | for all \(x \in \overline{\Omega }\) and, for each \(x \in \overline{\Omega }\), there is a curve η x  ∈ Lip([0, τ(x)]) having the properties: η x (0) = x, η x (τ(x)) = y, \(\eta _{x}(s) \in \overline{\Omega }\) for all s ∈ [0, τ(x)] and \(\vert \dot{\eta }_{x}(s)\vert \leq 1\) for a.e. s ∈ [0, τ(x)]. We fix such a function τ and a collection {η x } of curves. Thanks to Lemma 5.2, we may choose constants δ > 0 and C 0 > 0 such that

$$\displaystyle{L(x,\xi ) \leq C_{0}\ \ \ \text{ for all }\ (x,\xi ) \in \overline{\Omega } \times \overline{B}_{\delta }.}$$

Fix any \(x \in \overline{\Omega } \setminus \{ y\}\) and define (η, v, l) ∈ SP(x) by setting η(s) = η x (δs) for s ∈ [0, τ(x) ∕ δ], η(s) = y for s > τ(x) ∕ δ and \((v(s),\ l(s)) = (\dot{\eta }(s),0)\) for \(s \in \overline{\mathbb{R}}_{+}\). Observe that

$$\displaystyle\begin{array}{rcl} \mathcal{L}(\tau (x)/\delta,\eta,v,l)& =& \int _{0}^{\tau (x)/\delta }L(\eta _{ x}(\delta s),\,\delta \dot{\eta }_{x}(\delta s))\mathrm{d}s {}\\ & =& {\delta }^{-1}\int _{ 0}^{\tau (x)}L(\eta _{ x}(s),-\delta \dot{\eta }_{x}(s))\mathrm{d}s {}\\ & \leq & {\delta }^{-1}C_{ 0}\tau (x) {\leq \delta }^{-1}C_{ 0}C_{1}\vert x - y\vert, {}\\ \end{array}$$

which yields

$$\displaystyle{ W(x,y) {\leq \delta }^{-1}C_{ 0}C_{1}\vert x - y\vert. }$$
(110)

We define the function \(w\, :\, \overline{Q} \rightarrow \mathbb{R}\) by

$$\displaystyle\begin{array}{rcl} w(x,t) =\inf \big\{ \mathcal{L}(r,\eta,v,l)\, :\, r > t,\ (\eta,v,l) \in \mathrm{SP}(x)\ \ \text{ such that }\ \ \eta (r) = y\big\}.& & {}\\ \end{array}$$

It is clear by the above definition that

$$\displaystyle{ W(x,y) =\inf _{t>0}w(x,t)\ \ \ \text{ for all }x \in \overline{\Omega }. }$$
(111)

Also, the dynamic programming principle yields

$$\displaystyle\begin{array}{rcl} w(x,t) =\inf \big\{ \mathcal{L}(t,\eta,v,l) + W(\eta (t),y)\, :\, (\eta,v,l) \in \mathrm{SP}(x)\big\}.& & {}\\ \end{array}$$

(We leave it to the reader to prove this identity.) In view of (110), we fix a \(k \in \mathbb{N}\) so that \({\delta }^{-1}C_{0}C_{1} \leq k\) and note that for all (x, t) ∈ Q,

$$\displaystyle\begin{array}{rcl} w(x,t) \leq \inf \big\{\mathcal{L}(t,\eta,v,l) + k\vert \eta (t) - y\vert \, :\, (\eta,v,l) \in \mathrm{SP}(x)\big\} = u_{k}(x,t).& & {}\\ \end{array}$$

Consequently, we have

$$\displaystyle{\inf _{t>0}w(x,t) \leq \inf _{t>0}u_{k}(x,t)\ \ \ \text{ for all }\ x \in \overline{\Omega },}$$

which together with (111) yields

$$\displaystyle{W(x,y) \leq \inf _{t>0}u_{k}(x,t)\ \ \ \text{ for all }\ x \in \overline{\Omega }.}$$

By Lemma 5.11, if we set \(v(x) =\inf _{t>0}u_{k}(x,t)\) for \(x \in \overline{\Omega }\), then \(v \in C(\overline{\Omega })\) is a viscosity subsolution of (SNP). Moreover, since v(x) ≤ u k (x, 0) = k | x − y | for all \(x \in \overline{\Omega }\), we have v(y) ≤ 0. Thus, we find that v(x) ≤ v(y) + d(x, y) ≤ d(x, y) for all \(x \in \overline{\Omega }\). We now conclude that W(x, y) ≤ v(x) ≤ d(x, y) for all \(x \in \overline{\Omega }\). The proof is complete. □ 

Proposition 5.5.

Let \(y \in \overline{\Omega }\) and δ > 0. Then we have \(y \in \mathcal{A}\) if and only if

$$\displaystyle{ \inf \big\{\mathcal{L}(t,\eta,v,l)\, :\, t >\delta,\ (\eta,v,l) \in \mathrm{SP}(y)\ \ \text{ such that }\ \ \eta (t) = y\big\} = 0. }$$
(112)

Proof.

First of all, we define the function \(u \in \mathrm{UC}(\overline{Q})\) as the viscosity solution of (ENP)–(ID), with u 0 = d( ⋅, y). By Theorem 5.5, we have

$$\displaystyle{u(x,t) =\inf \big\{ \mathcal{L}(t,\eta,v,l) + d(\eta (t),y)\, :\, (\eta,v,l) \in \mathrm{SP}(x)\big\}\ \ \text{ for all }(x,t) \in Q.}$$

We combine this formula and Proposition 5.4, to get

$$\displaystyle\begin{array}{rcl} u(x,t) =\inf \Big\{ \mathcal{L}(r,\eta,v,l)\, :\,& r >& t,\,(\eta,v,l) \in \mathrm{SP}(x)\ \text{ such that }\ \eta (r) = y\Big\} \\ & & \qquad \qquad \text{ for all }(x,t) \in Q.\qquad \qquad {}\end{array}$$
(113)

Now, we assume that \(y \in \mathcal{A}\). The function d( ⋅, y) is then a viscosity solution of (SNP) and u is a viscosity solution of (ENP)–(ID), with u 0 = d( ⋅, y). Hence, by Theorem 3.1, we have d(x, y) = u(x, t) for all \((x,t) \in \overline{Q}\). Thus,

$$\displaystyle\begin{array}{rcl} 0 = d(y,y) =\inf \big\{ \mathcal{L}(r,\eta,v,l)\, :\, r > t,\ & & (\eta,v,l) \in \mathrm{SP}(y)\ \ \text{ such that }\ \ \eta (r) = y\big\} {}\\ & & \qquad \qquad \text{ for all }t > 0. {}\\ \end{array}$$

This shows that (112) is valid.

Now, we assume that (112) holds. This assumption and (5.5) show that u(y, δ) = 0. Formula (5.5) shows as well that for each \(x \in \overline{\Omega }\), the function u(x,  ⋅) is nondecreasing in \(\overline{\mathbb{R}}_{+}\). In particular, we have  d(x, y) ≤ u(x, t) for all (x, t) ∈ Q. Let \(p \in D_{x}^{-}d(x,y)\vert _{x=y}\). Then we have  (p, 0) ∈ D  −  u(y, δ) and

$$\displaystyle{\left \{\begin{array}{llll} &H(y,p) \geq 0 &&\mathrm{if}\ y \in \Omega,\\ &\max \{H(y, p),\,\gamma (y) \cdot p - g(y)\} \geq 0 & &\mathrm{if } \ y \in \partial \Omega. \end{array} \right.}$$

This shows that d( ⋅, y) is a viscosity solution of (SNP). Hence, we have  \(y \in \mathcal{A}\). □ 

6 Large-Time Asymptotic Solutions

We discuss the large-time behavior of solutions of (ENP)–(ID) following [8, 38, 39].

There has been much interest in the large time behavior of solutions of Hamilton–Jacobi equations since Namah and Roquejoffre in [53] have first established a general convergence result for solutions of

$$\displaystyle{ u_{t}(x,t) + H(x,D_{x}u(x,t)) = 0\ \ \ \mathrm{in}\ \ (x,t) \in \Omega \times \mathbb{R}_{+} }$$
(1.2)

under (A5), (A6) and the assumptions

$$\displaystyle\begin{array}{rcl} & & H(x,p) \geq H(x,0)\ \ \ \text{ for all }\ (x,p) \in \Omega \times {\mathbb{R}}^{n}, \\ & & \max _{\Omega }H(x,0) = 0, {}\end{array}$$
(114)

where Ω is a smooth compact n-dimensional manifold without boundary. Fathi in [26] has then established a similar convergence result but under different type hypotheses, where (114) replaced by a strict convexity of the Hamiltonian H(x, p) in p, by the dynamical approach based on weak KAM theory [25, 27]. Barles and Souganidis have obtained in [3] more general results in the periodic setting (i.e., in the case where Ω is n-dimensional torus), for possibly non-convex Hamiltonians, by using a PDE-viscosity solutions approach, which does not depend on the variational formula for the solutions like the one in Theorem 5.5. We refer to [7] for a recent view on this approach.

The approach of Fathi has been later modified and refined by Roquejoffre [54], Davini and Siconolfi in [21], and others. The same asymptotic problem in the whole domain \({\mathbb{R}}^{n}\) has been investigated by Barles and Roquejoffre in [10], Fujita et al., Ichihara and the author in [30, 3437] in various situations.

There has been as well a considerable interest in the large time asymptotic behavior of solutions of Hamilton–Jacobi equation with boundary conditions. The investigations in this direction are papers: Mitake [48] (the state-constraint boundary condition), Roquejoffre [54] (the Dirichlet boundary condition in the classical sense), Mitake [49, 50] (the Dirichlet boundary condition in the viscosity framework). More recent studies are due to Barles, Mitake and the author in [8, 9, 38], where the Neumann boundary conditions including the dynamical boundary conditions are treated. In [8, 9], the PDE-viscosity solutions approach of Barles–Souganidis is adapted to problems with boundary conditions.

Yokoyama et al. in [58] and Giga et al. in [32, 33] have obtained some results on the large time behavior of solutions of Hamilton–Jacobi equations with noncoercive Hamiltonian which is motivated by a crystal growth model.

We also refer to the articles [13, 54] and to [16, 51, 52] for the large time behavior of solutions, respectively, of time-dependent Hamilton–Jacobi equations and of weakly coupled systems of Hamilton–Jacobi equations.

As before, we assume throughout this section that hypotheses (A1)–(A7) hold and that \(u_{0} \in C(\overline{\Omega })\). Moreover, we assume that c # = 0. Throughout this section u = u(x, t) denotes the viscosity solution of (ENP)–(ID).

We set

$$\displaystyle{Z =\{ (x,p) \in \overline{\Omega } \times {\mathbb{R}}^{n}\, :\, H(x,p) = 0\}.}$$
  • (A9) ±  There exists a function ω 0 ∈ C([0, )) satisfying ω 0(r) > 0 for all r > 0 such that if (x, p) ∈ Z, \(\xi \in D_{p}^{-}H(x,p)\) and \(q \in {\mathbb{R}}^{n}\), then

    $$\displaystyle{H(x,p + q) \geq \xi \cdot q +\omega _{0}((\xi \cdot q)_{\pm }).}$$

The following proposition describes the long time behavior of solutions of (ENP)–(ID).

Theorem 6.1.

Assume that either (A9)+ or (A9) holds. Then there exists a viscosity solution \(w \in \mathrm{Lip}(\overline{\Omega })\) of (SNP) for which

$$\displaystyle{ \lim _{t\rightarrow \infty }u(x,t) = w(x)\ \ \ \text{ uniformly on }\ \overline{\Omega }. }$$
(115)

The following example is an adaptation of the one from Barles–Souganidis to the Neumann problem, which shows the necessity of a stronger condition like (A9) ±  beyond the convexity assumption (A7) in order to have the asymptotic behavior described in the above theorem.

Example 6.1.

Let n = 2 and Ω = B 4. Let \(\eta,\,\zeta \in {C}^{1}(\overline{\mathbb{R}}_{+})\) be functions such that 0 ≤ η(r) ≤ 1 for all \(r \in \overline{\mathbb{R}}_{+}\), η(r) = 1 for all r ∈ [0, 1], η(r) = 0 for all r ∈ [2, ), ζ(r) ≥ 0 for all \(r \in \overline{\mathbb{R}}_{+}\), ζ(r) = 0 for all r ∈ [0, 2] ∪[3, ) and ζ(r) > 0 for all r ∈ (2, 3). Fix a constant M > 0 so that \(M \geq \|\zeta ^{\prime}\|_{\infty,\mathbb{R}_{+}}\). We consider the Hamiltonian \(H\, :\, \overline{\Omega } \times {\mathbb{R}}^{2}\) given by

$$\displaystyle\begin{array}{rcl} H(x,y,p,q)& =& \vert - yp + xq +\zeta (r)\vert -\zeta (r) {}\\ & & \quad +\eta (r)\sqrt{{p}^{2 } + {q}^{2}} + (1 -\eta (r))\left (\Big\vert \frac{x} {r}p + \frac{y} {r}q\Big\vert - M\right )_{+}, {}\\ \end{array}$$

where \(r = r(x,y) := \sqrt{{x}^{2 } + {y}^{2}}\). Let \(u \in {C}^{1}(\overline{\Omega } \times \overline{\mathbb{R}}_{+})\) be the function given by

$$\displaystyle{u(x,y,t) =\zeta (r)\left (\frac{y} {r}\cos t -\frac{x} {r}\sin t\right ),}$$

where, as above, \(r = \sqrt{{x}^{2 } + {y}^{2}}\). It is easily checked that u is a classical solution of

$$\displaystyle{\left \{\begin{array}{llll} &u_{t}(x,y,t) + H(x,y,u_{x}(x,y,t),u_{y}(x,y,t)) = 0&&\mathrm{in}\ \ B_{4} \times \mathbb{R}_{+}, \\ &\nu (x,y) \cdot (u_{x}(x,y,t),u_{y}(x,y,t)) = 0 &&\mathrm{on}\ \partial B_{4} \times \mathbb{R}_{+}, \end{array} \right.}$$

where ν(x, y) denotes the outer unit normal at (x, y) ∈ ∂B 4. Note here that if we introduce the polar coordinate system

$$\displaystyle{x = r\cos \theta,\quad y = r\sin \theta }$$

and the new function

$$\displaystyle{v(r,\theta,t) = u(r\cos \theta,r\sin \theta,t)\ \ \ \text{ for }(r,\theta,t) \in \mathbb{R}_{+} \times \mathbb{R} \times \overline{\mathbb{R}}_{+},}$$

then the above Hamilton–Jacobi equation reads

$$\displaystyle{v_{t} +\widetilde{ H}(r,\theta,v_{r},v_{\theta }) = 0,}$$

where

$$\displaystyle\begin{array}{rcl} & \widetilde{H}(r,\theta,p_{r},p_{\theta }) & {}\\ & \quad = \vert p_{\theta } +\zeta (r)\vert -\zeta (r) +\eta (r)\sqrt{p_{r }^{2 } +{ \left (\frac{p_{\theta }} {r}\right )}^{2}} + (1 -\eta (r))\left (\vert p_{r}\vert - M\right )_{+},& {}\\ \end{array}$$

while the definition of u reads

$$\displaystyle{v(r,\theta,t) =\zeta (r)\sin (\theta -t).}$$

Note also that any constant function w on \(\overline{B}_{4}\) is a classical solution of

$$\displaystyle{\left \{\begin{array}{llll} &H(x,y,w_{x}(x,y,t),w_{y}(x,y,t)) = 0 &&\mathrm{in}\ \ B_{4}, \\ &\nu (x,y) \cdot (w_{x}(x,y,t),w_{y}(x,y,t)) = 0&&\mathrm{on}\ \partial B_{4}, \end{array} \right.}$$

which implies that the eigenvalue c # is zero.

It is clear that u does not have the asymptotic behavior (115). As is easily seen, the Hamiltonian H satisfies (A5)–(A7), but neither of (A9) ± .

6.1 Preliminaries to Asymptotic Solutions

According to Theorem 3.3 and Corollary 3.1, we know that \(u \in \mathrm{BUC}(\overline{Q})\). We set

$$\displaystyle{u_{\infty }(x) =\liminf _{t\rightarrow \infty }u(x,t)\ \ \ \text{ for all }\ x \in \overline{\Omega }.}$$

Lemma 6.1.

The function u is a viscosity solution of (SNP) and \(u_{\infty }\in \mathrm{Lip}(\overline{\Omega })\).

Proof.

Note that

$$\displaystyle{ u_{\infty }(x) =\lim _{t\rightarrow \infty }\inf \{u(x,t + r)\, :\, r > 0\}\ \ \ \text{ for all }\ x \in \overline{\Omega }. }$$
(116)

By Lemma 5.11, if we set

$$\displaystyle{v(x,t) =\inf \{ u(x,t + r)\, :\, r > 0\}\ \ \ \text{ for }\ (x,t) \in \overline{Q},}$$

then \(v \in \mathrm{BUC}(\overline{Q})\) and it is a viscosity solution of (ENP). For each \(x \in \overline{\Omega }\), the function v(x,  ⋅) is nondecreasing in \(\overline{\mathbb{R}}_{+}\). Hence, by the Ascoli–Arzela theorem or Dini’s lemma, we see that the convergence in (116) is uniform in \(\overline{\Omega }\). By Proposition 1.9, we see that the function u (x), as a function of (x, t), is a viscosity solution of (ENP), which means that u is a viscosity solution of (SNP). Finally, Proposition 1.14 guarantees that \(u_{\infty }\in \mathrm{Lip}(\overline{\Omega })\). □ 

We introduce the following notation:

$$\displaystyle\begin{array}{rcl} & & S =\{ (x,\xi ) \in \overline{\Omega } \times {\mathbb{R}}^{n}\, :\,\xi \in D_{ p}^{-}H(x,p)\ \text{ for some }\ (x,p) \in Z\}, {}\\ & & P(x,\xi ) =\{ p \in {\mathbb{R}}^{n}\, :\,\xi \in D_{ p}^{-}H(x,p)\}\ \ \ \text{ for }\ (x,\xi ) \in \overline{\Omega } \times {\mathbb{R}}^{n}. {}\\ \end{array}$$

Lemma 6.2.

  1. (i)

    \(Z,\,S \subset \overline{\Omega } \times B_{R_{0}}\,\) for some R 0 > 0.

  2. (ii)

    Assume that (A9) + holds. Then there exist constants δ > 0 and R 1 > 0 such that for any (x,ξ) ∈ S and any \(\varepsilon \in (0,\,\delta )\) , we have \(P(x,(1+\varepsilon )\xi )\neq \varnothing \) and \(P(x,(1+\varepsilon )\xi ) \subset B_{R_{1}}\).

  3. (iii)

    Assume that (A9) holds. Then there exist constants δ > 0 and R 1 > 0 such that for any (x,ξ) ∈ S and any \(\varepsilon \in (0,\,\delta )\) , we have \(P(x,(1-\varepsilon )\xi )\neq \varnothing \) and \(P(x,(1-\varepsilon )\xi ) \subset B_{R_{1}}\) .

Proof.

(i) It follows from coercivity (A6) that there exists a constant R 1 > 0 such that \(Z \subset {\mathbb{R}}^{n} \times B_{R_{1}}\). Next, fix any (x, ξ) ∈ S. Then, by the definition of S, we may choose a point p ∈ P(x, ξ) such that (x, p) ∈ Z. Note that | p |  < R 1. By convexity (A7), we have

$$\displaystyle{H(x,p^{\prime}) \geq H(x,p) +\xi \cdot (p^{\prime} - p)\quad \text{ for all }p^{\prime} \in {\mathbb{R}}^{n}.}$$

Assuming that ξ≠0 and setting p′ = p + ξ ∕ | ξ | in the above, we get

$$\displaystyle{\vert \xi \vert =\xi \cdot (p^{\prime} - p) \leq H(x,p^{\prime}) - H(x,p) <\sup _{\overline{\Omega }\times B_{ R_{1}+1}}H -\inf _{\overline{\Omega }\times B_{ R_{1}}}H.}$$

We may choose a constant R 2 > 0 so that the right-hand side is less than R 2, and therefore \(\xi \in B_{R_{2}}\). Setting \(R_{0} =\max \{ R_{1},R_{2}\}\), we conclude that \(Z,\,S \subset {\mathbb{R}}^{n} \times B_{R_{0}}\).

(ii) By (i), there is a constant R 0 > 0 such that \(Z,S \subset \overline{\Omega } \times B_{R_{0}}\). We set δ = ω 0(1), where ω 0 is from (A9)\(_{_{+}}\). In view of coercivity (A6), replacing R 0 > 0 by a larger constant if necessary, we may assume that H(x, p) ≥ 1 + ω 0(1) for all \((x,p) \in \overline{\Omega } \times \left ({\mathbb{R}}^{n} \setminus B_{R_{0}}\right )\).

Fix any (x, ξ) ∈ S, p ∈ P(x, ξ) and \(\varepsilon \in (0,\delta )\). Note that \(\xi,p \in B_{R_{0}}\). By (A9) + , for all \(x \in {\mathbb{R}}^{n}\) we have

$$\displaystyle{ H(x,q) \geq \xi \cdot (q - p) +\omega _{0}\left ((\xi \cdot (q - p))_{+}\right ). }$$

We set \(V :=\{ q \in \overline{B}_{2R_{0}}(p)\, :\, \vert \xi \cdot (q - p)\vert \leq 1\}\). Let q ∈ V and observe the following: if \(q \in \partial B_{2R_{0}}(p)\), which implies that | q | ≥ R 0, then \(H(x,q) \geq 1 +\omega _{0}(1) > 1+\varepsilon \geq (1+\varepsilon )\xi \cdot (q - p)\). If ξ ⋅(q − p) = 1, then \(H(x,q) \geq 1 +\omega _{0}(1) > 1+\varepsilon = (1+\varepsilon )\xi \cdot (q - p)\). Also, if ξ ⋅(q − p) =  − 1, then \(H(x,q) \geq \xi \cdot (q - p) > (1+\varepsilon )\xi \cdot (q - p)\). Accordingly, the function \(G(q) := H(x,q) - (1+\varepsilon )\xi \cdot (q - p)\) on \({\mathbb{R}}^{n}\) is positive on ∂V while it vanishes at q = p ∈ V, and hence it attains a minimum over the set V at an interior point of V. Thus, \(P(x,(1+\varepsilon )\xi )\not =\varnothing \). By the convexity of G, we see easily that G(q) > 0 for all \(q \in {\mathbb{R}}^{n} \setminus V\) and conclude that \(P(x,(1+\varepsilon )\xi ) \subset B_{2R_{0}}\).

(iii) Let ω 0 be the function from (A9)\(_{_{-}}\). As before, we choose R 0 > 0 so that \(Z,S \subset \overline{\Omega } \times B_{R_{0}}\) and H(x, p) ≥ 1 + ω 0(1) for all \((x,p) \in \overline{\Omega } \times \left ({\mathbb{R}}^{n} \setminus B_{R_{0}}\right )\), and set δ = ω 0(1) ∧ 1. Note that for all \(x \in {\mathbb{R}}^{n}\),

$$\displaystyle{ H(x,q) \geq \xi \cdot (q - p) +\omega _{0}\left ((\xi \cdot (q - p))_{_{-}}\right ). }$$

Fix any (x, ξ) ∈ S, p ∈ P(x, ξ) and \(\varepsilon \in (0,\delta )\). Set \(V :=\{ q \in \overline{B}_{2R_{0}}(p)\, :\, \vert \xi \cdot (q - p)\vert \leq 1\}\). Let q ∈ V and observe the following: if \(q \in \partial B_{2R_{0}}(p)\), then \(H(x,q) \geq 1 +\omega _{0}(1) > 1+\varepsilon \geq (1-\varepsilon )\xi \cdot (q - p)\). If ξ ⋅(q − p) =  − 1, then \(H(x,q) \geq -1 +\omega _{0}(1) > -1+\varepsilon = (1-\varepsilon )\xi \cdot (q - p)\). If ξ ⋅(q − p) = 1, then \(H(x,q) \geq \xi \cdot (q - p) > (1-\varepsilon )\xi \cdot (q - p)\). As before, the function \(G(q) := H(x,q) - (1-\varepsilon )\xi \cdot (q - p)\) attains a minimum over V at an interior point of V. Consequently, \(P(x,(1-\varepsilon )\xi )\not =\varnothing \). Moreover, we get \(P(x,(1-\varepsilon )\xi ) \subset B_{2R_{0}}\). □ 

Lemma 6.3.

Assume that (A9) \(_{_{+}}\) (resp., (A9) \(_{_{-}}\) ) holds. Then there exist a constant δ 1 > 0 and a modulus ω 1 such that for any \(\varepsilon \in [0,\delta _{1}]\) and (x,ξ) ∈ S,

$$\displaystyle{ L(x,(1+\varepsilon )\,\xi ) \leq (1+\varepsilon )L(x,\xi ) +\varepsilon \,\omega _{1}(\varepsilon ) }$$
(117)

(resp.,

$$\displaystyle{ L(x,(1-\varepsilon )\,\xi ) \leq (1-\varepsilon )L(x,\xi ) +\varepsilon \,\omega _{1}(\varepsilon )\,\text{)}. }$$
(118)

Before going into the proof, we make the following observation: under the assumption that H, L are smooth, for any (x, ξ) ∈ S, if we set p : = D ξ L(x, ξ), then

$$\displaystyle\begin{array}{rcl} & & H(x,p) = 0, {}\\ & & p\cdot \xi = H(x,p) + L(x,\xi ) = L(x,\xi ), {}\\ \end{array}$$

and, as \(\varepsilon \rightarrow 0\),

$$\displaystyle\begin{array}{rcl} L(x,(1+\varepsilon )\xi )& =& L(x,\xi ) +\varepsilon p \cdot \xi +o(\varepsilon ) {}\\ & =& L(x,\xi ) +\varepsilon L(x,\xi ) + o(\varepsilon ) = (1+\varepsilon )L(x,\xi ) + o(\varepsilon ). {}\\ \end{array}$$

Proof.

Assume that (A9)\(_{_{+}}\) holds. Let R 0 > 0, R 1 > 0 and δ > 0 be the constants from Lemma 6.2. Fix any (x, ξ) ∈ S and \(\varepsilon \in [0,\delta )\). In view of Lemma 6.2, we may choose a \(p_{\varepsilon } \in P(x,(1+\varepsilon )\xi )\). Then we have \(\vert p_{\varepsilon } - p_{0}\vert < 2R_{1}\), | ξ |  < R 0 and \(\vert \xi \cdot (p_{\varepsilon } - p_{0})\vert < 2R_{0}R_{1}\).

Note by (A9)\(_{_{+}}\) that

$$\displaystyle{ H(x,p_{\varepsilon }) \geq \xi \cdot (p_{\varepsilon } - p_{0}) +\omega _{0}\left ((\xi \cdot (p_{\varepsilon } - p_{0}))_{+}\right ). }$$

Hence, we obtain

$$\displaystyle\begin{array}{rcl} L(x,(1+\varepsilon )\,\xi )& =& (1+\varepsilon )\,\xi \cdot p_{\varepsilon } - H(x,p_{\varepsilon }) \leq (1+\varepsilon )\,\xi \cdot p_{\varepsilon } {}\\ & & \quad -\xi \cdot (p_{\varepsilon } - p_{0}) -\omega _{0}\left ((\xi \cdot (p_{\varepsilon } - p_{0}))_{+}\right ) {}\\ & \leq & (1+\varepsilon )[\xi \cdot p_{0} - H(x,p_{0})] {}\\ & & \quad +\varepsilon \,\xi \cdot (p_{\varepsilon } - p_{0}) -\omega _{0}\left ((\xi \cdot (p_{\varepsilon } - p_{0}))_{+}\right ) {}\\ & \leq & (1+\varepsilon )L(x,\xi ) +\varepsilon \max _{0\leq r\leq 2R_{0}R_{1}}\left (r -\frac{1} {\varepsilon } \omega _{0}(r)\right ). {}\\ \end{array}$$

We define the function ω 1 on [0, ) by setting \(\omega _{1}(s) =\max _{0\leq r\leq 2R_{0}R_{1}}(r -\omega _{0}(r)/s)\) for s > 0 and ω 1(0) = 0 and observe that ω 1 ∈ C([0, )). We have also \(L(x,(1+\varepsilon )\xi ) \leq (1+\varepsilon )L(x,\xi ) +\varepsilon \omega _{1}(\varepsilon )\) for all \(\varepsilon \in (0,\delta )\). Thus (117) holds with δ 1 : = δ ∕ 2.

Next, assume that (A9)\(_{_{-}}\) holds. Let R 0 > 0, R 1 > 0 and δ > 0 be the constants from Lemma 6.2. Fix any (x, ξ) ∈ S and \(\varepsilon \in [0,\delta )\).

As before, we may choose a \(p_{\varepsilon } \in P(x,(1-\varepsilon )\xi )\), and observe that \(\vert p_{\varepsilon } - p_{0}\vert < 2R_{1}\), | ξ |  < R 0 and \(\vert \xi \cdot (p_{\varepsilon } - p_{0})\vert < 2R_{0}R_{1}\). Noting that

$$\displaystyle{ H(x,p_{\varepsilon }) \geq \xi \cdot (p_{\varepsilon } - p_{0}) +\omega _{0}\left ((\xi \cdot (p_{\varepsilon } - p_{0}))_{_{-}}\right ), }$$

we obtain

$$\displaystyle\begin{array}{rcl} L(x,(1-\varepsilon )\,\xi )& =& (1-\varepsilon )\,\xi \cdot p_{\varepsilon } - H(x,p_{\varepsilon }) \leq (1-\varepsilon )\,\xi \cdot p_{\varepsilon } {}\\ & & \quad -\xi \cdot (p_{\varepsilon } - p_{0}) -\omega _{0}\left ((\xi \cdot (p_{\varepsilon } - p_{0}))_{_{-}}\right ) {}\\ & \leq & (1-\varepsilon )[\xi \cdot p_{0} - H(x,p_{0})] {}\\ & & \quad -\varepsilon \,\xi \cdot (p_{\varepsilon } - p_{0}) -\omega _{0}\left ((\xi \cdot (p_{\varepsilon } - p_{0}))_{_{-}}\right ) {}\\ & \leq & (1+\varepsilon )L(x,\xi ) +\varepsilon \max _{0\leq r\leq 2R_{0}R_{1}}\left (r -\frac{1} {\varepsilon } \omega _{0}(r)\right ). {}\\ \end{array}$$

Setting \(\omega _{1}(s) =\max _{0\leq r\leq 2R_{0}R_{1}}(r -\omega _{0}(r)/s)\) for s > 0 and ω 1(0) = 0, we find a function ω 1 ∈ C([0, )) vanishing at the origin for which \(L(x,(1-\varepsilon )\xi ) \leq (1-\varepsilon )L(x,\xi ) +\varepsilon \omega _{1}(\varepsilon )\) for all \(\varepsilon \in (0,\delta )\). Thus (118) holds with δ 1 : = δ ∕ 2. □ 

Theorem 6.2.

Let \(u \in \mathrm{Lip}(\overline{\Omega })\) be a subsolution of (SNP) . Let \(\eta \in \mathrm{AC}(\mathbb{R}_{+},\, {\mathbb{R}}^{n})\) be such that \(\eta (t) \in \overline{\Omega }\) for all \(t \in \mathbb{R}_{+}\) . Set \(\mathbb{R}_{+,\text{b}} =\{ t \in \mathbb{R}_{+}\, :\,\eta (t) \in \partial \Omega \}\) . Then there exists a function \(p \in {L}^{\infty }(\mathbb{R}_{+},\, {\mathbb{R}}^{n})\) such that

$$\displaystyle{\left \{\begin{array}{llll} & \frac{\mathrm{d}} {\mathrm{d}t}u \circ \eta (t) = p(t) \cdot \dot{\eta } (t)&&\text{ for a.e. }t \in \mathbb{R}_{+}, \\ &H(\eta (t),\,p(t)) \leq 0 &&\text{ for a.e. }t \in \mathbb{R}_{+}, \\ &\gamma (\eta (t)) \cdot p(t) \leq g(\eta (t)) &&\text{ for a.e. }t \in \mathbb{R}_{+,\text{b}}. \end{array} \right.}$$

Proof.

According to Theorem 4.2, there is a collection \(\{u_{\varepsilon }\}_{\varepsilon \in (0,\,1)} \subset {C}^{1}(\overline{\Omega })\) such that

$$\displaystyle{\left \{\begin{array}{lllll} &H(x,\mathit{Du}_{\varepsilon }(x)) \leq \varepsilon &&\text{ for all }x \in \overline{\Omega }, & \\ &\frac{\partial u_{\varepsilon }} {\partial \gamma } (x) \leq g(x) &&\text{ for all }x \in \partial \Omega,&\\ &\|u_{ \varepsilon } - u\|_{\infty,\Omega } <\varepsilon, && & \\ &\sup _{0<\varepsilon <1}\|\mathit{Du}_{\varepsilon }\|_{{L}^{\infty }(\Omega )} < \infty.&& &\end{array} \right.}$$

If we set \(p_{\varepsilon }(t) = \mathit{Du}_{\varepsilon } \circ \eta (t)\) for all \(t \in \overline{\mathbb{R}}_{+}\), then we have

$$\displaystyle{ \left \{\begin{array}{ll} &u_{\varepsilon } \circ \eta (t) - u_{\varepsilon } \circ \eta (0) =\int _{ 0}^{t}p_{\varepsilon }(s) \cdot \dot{\eta } (s)\mathrm{d}s\ \ \text{ for all }t \in \mathbb{R}_{+}, \\ &H(\eta (t),\,p_{\varepsilon }(t)) \leq \varepsilon \ \ \text{ for all }t \in \mathbb{R}_{+}, \\ &\gamma (\eta (t)) \cdot p_{\varepsilon }(t) \leq g(\eta (t))\ \ \text{ for all }t \in \mathbb{R}_{+,\text{b}}. \end{array} \right. }$$
(119)

Since \(\{p_{\varepsilon }\}_{\varepsilon \in (0,1)}\) is bounded in \({L}^{\infty }(\mathbb{R}_{+})\), there is a sequence \(\{\varepsilon _{j}\}_{j\in \mathbb{N}}\) converging to zero such that, as j → , the sequence \(\{p_{\varepsilon _{j}}\}\) converges weakly-star in \({L}^{\infty }(\mathbb{R}_{+})\) to some function \(p \in {L}^{\infty }(\mathbb{R}_{+})\). It is clear from (119) that

$$\displaystyle{\left \{\begin{array}{ll} &u \circ \eta (t) - u \circ \eta (0) =\int _{ 0}^{t}p(s) \cdot \dot{\eta } (s)\mathrm{d}s\ \ \text{for all }t \in \mathbb{R}_{+}, \\ &\gamma (\eta (t)) \cdot p(t) \leq g(\eta (t))\ \ \text{ for a.e. }t \in \mathbb{R}_{+,\text{b}}. \end{array} \right.}$$

Now, we fix an \(i \in \mathbb{N}\) so that \(i >\| p\|_{{L}^{\infty }(\mathbb{R}_{+})}\) and any 0 < T < , and set J = [0, T]. Using Lemma 5.6, for each \(m \in \mathbb{N}\), we find a function \(v_{m} \in {L}^{\infty }(J, {\mathbb{R}}^{n})\) so that

$$\displaystyle\begin{array}{rcl} & & H(\eta (s),p(s)) + L_{i}(\eta (s),-v_{m}(s)) < -v_{m}(s) \cdot p(s) + 1/m\ \ \ \text{ for a.e. }\ s \in J.{}\end{array}$$
(120)

By the convex duality, we have

$$\displaystyle{H(x,q) =\sup _{\xi \in {\mathbb{R}}^{n}}(\xi \cdot q - L_{i}(x,\xi ))\ \ \ \text{ for all }(x,q) \in \overline{\Omega } \times B_{i}.}$$

(Note that L i (x,  ⋅) is the convex conjugate of the function \(H(x,\cdot ) +\delta _{\overline{B}_{ i}}\), where \(\delta _{\overline{B}_{ i}}(p) = 0\) if \(p \in \overline{B}_{i}\) and =  otherwise.) Hence, for any nonnegative function \(\psi \in {L}^{\infty }(J, \mathbb{R})\) and any \((j,m) \in {\mathbb{N}}^{2}\), by (119) we get

$$\displaystyle\begin{array}{rcl} \varepsilon _{j}\int _{J}\psi (s)\mathrm{d}s& \geq & \int _{J}\psi (s)H(\eta (s),p_{\varepsilon _{j}}(s))\mathrm{d}s {}\\ & \geq & \int _{J}\psi (s)[-v_{m}(s) \cdot p_{\varepsilon _{j}}(s) - L_{i}(\eta (s),-v_{m}(s))]\mathrm{d}s. {}\\ \end{array}$$

Combining this observation with (120), after sending j → , we obtain

$$\displaystyle{0 \geq \int _{J}\psi (s)(H(\eta (s),p(s)) - 1/m)\mathrm{d}s,}$$

which implies that H(η(s), p(s)) ≤ 0 for a.e. s ∈ [0, T]. Since T > 0 is arbitrary, we see that

$$\displaystyle{H(\eta (s),p(s)) \leq 0\ \ \ \text{ for a.e. }\ s \in \mathbb{R}_{+}.}$$

The proof is complete. □ 

6.2 Proof of Convergence

This subsection is devoted to the proof of Theorem 6.1.

Proof (Theorem 6.1). 

It is enough to show that

$$\displaystyle{ \limsup _{t\rightarrow \infty }u(x,t) \leq u_{\infty }(x)\ \ \text{ for all }\ x \in \overline{\Omega }. }$$
(121)

Indeed, once this is proved, it is obvious that lim t →  u(x, t) = u (x) for all \(x \in \overline{\Omega }\), and moreover, since u ∈ BUC(Q), by the Ascoli–Arzela theorem, it follows that the convergence, \(\lim _{t\rightarrow \infty }u(x,t) = u_{\infty }(x)\), is uniform in \(\overline{\Omega }\).

Fix any \(z \in \overline{\Omega }\). According to Lemma 6.1 and Corollary 5.2, we may choose a (η, v, l) ∈ SP(z) such that for all t > 0,

$$\displaystyle{ u_{\infty }(z) - u_{\infty }(\eta (t)) = \mathcal{L}(t,\eta,v,l). }$$
(122)

Due to Theorem 6.2, there exists a function \(q \in {L}^{\infty }(\mathbb{R}_{+},\, {\mathbb{R}}^{n})\) such that

$$\displaystyle{ \left \{\begin{array}{llll} & \frac{\mathrm{d}} {\mathrm{d}s}u_{\infty }(\eta (s)) = q(s) \cdot \dot{\eta } (s)\ \ &&\text{ for a.e. }s \in \mathbb{R}_{+}, \\ &H(\eta (s),\,q(s)) \leq 0\ \ &&\text{ for a.e. }s \in \mathbb{R}_{+}, \\ &\gamma (\eta (s)) \cdot q(s) \leq g(\eta (s))\ \ &&\text{ for a.e. }s \in \mathbb{R}_{+,\text{b}}, \end{array} \right. }$$
(123)

where \(\mathbb{R}_{+,\text{b}} :=\{ s \in \mathbb{R}_{+}\, :\,\eta (s) \in \partial \Omega \}\).

We now show that

$$\displaystyle{ \left \{\begin{array}{llll} &H(\eta (s),\,q(s)) = 0 &&\text{ for a.e. s } \in \mathbb{R}_{+}, \\ &l(s)\gamma (\eta (s)) \cdot q(s) = l(s)g(\eta (s)) &&\text{ for a.e. }s \in \mathbb{R}_{+,\text{b}}, \\ & - q(s) \cdot v(s) = H(\eta (s),\,q(s)) + L(\eta (s),\,-v(s))&&\text{ for a.e. s } \in \mathbb{R}_{+}. \end{array} \right. }$$
(124)

We remark here that the last equality in (124) is equivalent to saying that

$$\displaystyle{-v(s) \in D_{p}^{-}H(\eta (s),\,q(s))\ \ \text{ for a.e. }s \in \mathbb{R}_{ +},}$$

(or

$$\displaystyle{q(s) \in D_{\xi }^{-}L(\eta (s),\,-v(s))\ \ \text{ for a.e. }s \in \mathbb{R}_{ +}.)}$$

By differentiating (122), we get

$$\displaystyle{-\frac{\mathrm{d}} {\mathrm{d}s}u_{\infty }(\eta (s)) = L(\eta (s),-v(s)) + l(s)g(\eta (s))\ \ \ \text{ for a.e. }s \in \mathbb{R}_{+}.}$$

Combining this with (123), we calculate

$$\displaystyle\begin{array}{rcl} 0& =& \,q(s) \cdot \dot{\eta } (s) + L(\eta (s),-v(s)) + l(s)g(\eta (s)) {}\\ & =& \,q(s) \cdot (v(s) - l(s)\gamma (\eta (s))) + L(\eta (s),-v(s)) + l(s)g(\eta (s)) {}\\ & \geq & \,-H(\eta (s),q(s)) - l(s)(q(s) \cdot \gamma (\eta (s)) - g(\eta (s))) \geq 0 {}\\ \end{array}$$

for a.e. \(s \in \mathbb{R}_{+}\), which guarantees that (124) holds.

Fix any \(\varepsilon > 0\). We prove that there is a constant τ > 0 and for each \(x \in \overline{\Omega }\) a number σ(x) ∈ [0, τ] for which

$$\displaystyle{ u_{\infty }(x)+\varepsilon > u(x,\sigma (x)). }$$
(125)

In view of the definition of u , for each \(x \in \overline{\Omega }\) there is a constant t(x) > 0 such that

$$\displaystyle{u_{\infty }(x)+\varepsilon > u(x,\,t(x)).}$$

By continuity, for each fixed \(x \in \overline{\Omega }\), we can choose a constant r(x) > 0 so that

$$\displaystyle{u_{\infty }(y)+\varepsilon > u(y,\,t(x))\ \ \text{ for }y \in \overline{\Omega } \cap B_{r(x)}(x),}$$

where \(B_{\rho }(x) :=\{ y \in {\mathbb{R}}^{n}\, :\, \vert y - x\vert <\rho \}\). By the compactness of \(\overline{\Omega }\), there is a finite sequence x i , i = 1, 2, , N, such that

$$\displaystyle{\overline{\Omega } \subset \bigcup _{1\leq i\leq N}B_{r(x_{i})}(x_{i}),}$$

That is, for any \(y \in \overline{\Omega }\) there exists x i , with 1 ≤ i ≤ N, such that \(y \in B_{r(x_{i})}(x_{i})\), which implies

$$\displaystyle{u_{\infty }(y)+\varepsilon > u(y,\,t(x_{i})).}$$

Thus, setting

$$\displaystyle{\tau =\max _{1\leq i\leq N}t(x_{i}),}$$

we find that for each \(x \in \overline{\Omega }\) there is a constant σ(x) ∈ [0, τ] such that (125) holds.

In what follows we fix τ > 0 and σ(x) ∈ [0, τ] as above. Also, we choose a constant δ 1 > 0 and a modulus ω 1 as in Lemma 6.3.

We divide our argument into two cases according to which hypothesis is valid, (A9) +  or (A9) − . We first argue under hypothesis (A9) + . Choose a constant T > τ so that τ ∕ (T − τ) ≤ δ 1. Fix any t ≥ T, and set θ = σ(η(t)) ∈ [0, τ]. We set δ = θ ∕ (t − θ) and note that δ ≤ τ ∕ (t − τ) ≤ δ 1. We define functions η δ , v δ , l δ on \(\overline{\mathbb{R}}_{+}\) by

$$\displaystyle\begin{array}{rcl} \eta _{\delta }(s)& =& \,\eta ((1+\delta )s), {}\\ v_{\delta }(s)& =& \,(1+\delta )v((1+\delta )s), {}\\ l_{\delta }(s)& =& \,(1+\delta )l((1+\delta )s), {}\\ \end{array}$$

and note that \((\eta _{\delta },\,v_{\delta },\,l_{\delta }) \in \mathrm{SP}(z)\).

By (124) together with the remark after (124), we know that H(η(s), q(s)) = 0 and \(-v(s) \in D_{p}^{-}H(\eta (s),\,q(s))\) for a.e. \(s \in \mathbb{R}_{+}\). That is, (η(s),  − v(s)) ∈ S for a.e. \(s \in \mathbb{R}_{+}\). Therefore, by (117), we get for a.e. \(s \in \mathbb{R}_{+}\),

$$\displaystyle{L(\eta _{\delta }(s),\,-v_{\delta }(s)) \leq (1+\delta )L\big(\eta ((1+\delta )s),\,-v((1+\delta )s)\big) +\delta \omega _{1}(\delta ).}$$

Integrating this over (0, t − θ), making a change of variables in the integral and noting that (1 + δ)(t − θ) = t, we get

$$\displaystyle\begin{array}{rcl} \int _{0}^{t-\theta }L(\eta _{\delta }(s),\,-v_{\delta }(s))\mathrm{d}s& \leq & \,\int _{ 0}^{t}L(\eta (s),\,-v(s))\mathrm{d}s + (t-\theta )\delta \omega _{ 1}(\delta ) {}\\ & =& \,\int _{0}^{t}L(\eta (s),\,-v(s))\mathrm{d}s +\theta \omega _{ 1}(\delta ), {}\\ \end{array}$$

as well as

$$\displaystyle{\int _{0}^{t-\theta }l_{\delta }(s)g(\eta _{\delta }(s))\mathrm{d}s =\int _{ 0}^{t}l(s)g(\eta (s))\mathrm{d}s.}$$

Moreover,

$$\displaystyle\begin{array}{rcl} u(z,t)& \leq & \,\mathcal{L}(t-\theta,\eta _{\delta },v_{\delta },l_{\delta }) + u(\eta _{\delta }(t-\theta ),\,\theta ) {}\\ & \leq & \,\int _{0}^{t}\big(L(\eta (s),\,-v(s)) + l(s)g(\eta (s))\big)\mathrm{d}s +\theta \omega _{ 1}(\delta ) + u\big(\eta (t),\,\sigma (\eta (t))\big) {}\\ & <& \,u_{\infty }(z) - u_{\infty }(\eta (t)) +\tau \omega _{1}(\delta ) + u_{\infty }(\eta (t)) +\varepsilon {}\\ & =& \,u_{\infty }(z) +\tau \omega _{1}(\delta ) +\varepsilon. {}\\ \end{array}$$

Thus, recalling that δ ≤ τ ∕ (t − τ), we get

$$\displaystyle{ u(z,t) \leq u_{\infty }(z) +\tau \omega _{1}\Big( \frac{\tau } {t-\tau }\Big) +\varepsilon. }$$
(126)

Next, we assume that (A9) −  holds. We choose T > τ as before, and fix t ≥ T. Set θ = σ(η(t − τ)) ∈ [0, τ] and δ = (τ − θ) ∕ (t − θ). Observe that (1 − δ)(t − θ) = t − τ and δ ≤ τ ∕ (t − τ) ≤ δ 1.

We set η δ (s) = η((1 − δ)s), v δ (s) = (1 − δ)v((1 − δ)s) and l δ (s) = (1 − δ)l((1 − δ)s) for \(s \in \overline{\mathbb{R}}_{+}\) and observe that \((\eta _{\delta },\,v_{\delta },\,l_{\delta }) \in \mathrm{SP}(z)\). As before, thanks to (118), we have

$$\displaystyle{L(\eta _{\delta }(s),\,-v_{\delta }(s)) \leq (1-\delta )L(\eta ((1-\delta )s),\,-v(1-\delta )s)) +\delta \omega _{1}(\delta )\ \ \text{ for a.e. }s \in \mathbb{R}_{+}.}$$

Hence, we get

$$\displaystyle\begin{array}{rcl} \int _{0}^{t-\theta }L(\eta _{\delta }(s),\,-v_{\delta }(s))\mathrm{d}s& \leq & \,\int _{ 0}^{t-\tau }L(\eta (s),\,-v(s))\mathrm{d}s + (t-\theta )\delta \omega _{ 1}(\delta ) {}\\ & =& \,\int _{0}^{t-\tau }L(\eta (s),\,-v(s))\mathrm{d}s + (\tau -\theta )\omega _{ 1}(\delta ), {}\\ \end{array}$$

and

$$\displaystyle{\int _{0}^{t-\theta }l_{\delta }(s)g(\eta _{\delta }(s))\mathrm{d}s =\int _{ 0}^{t-\tau }l(s)g(\eta (s))\mathrm{d}s.}$$

Furthermore, we calculate

$$\displaystyle\begin{array}{rcl} u(z,t)& \leq & \,\mathcal{L}(t-\theta,\eta _{\delta },\,v_{\delta },l_{\delta }) + u(\eta _{\delta }(t-\theta ),\,\theta ) {}\\ & \leq & \,\mathcal{L}(t-\tau,\eta,v,l) +\tau \omega _{1}(\delta ) + u(\eta (t-\tau ),\,\sigma (\eta (t-\tau ))) {}\\ & <& \,u_{\infty }(z) +\tau \omega _{1}(\delta ) +\varepsilon. {}\\ \end{array}$$

Thus, we get

$$\displaystyle{u(z,t) \leq u_{\infty }(z) +\tau \omega _{1}\Big( \frac{\tau } {t-\tau }\Big)+\varepsilon,}$$

From the above inequality and (126) we see that (121) is valid. □ 

6.3 Representation of the Asymptotic Solution u

According to Theorem 6.1, if either (A9) +  or (A9) −  holds, then the solution u(x, t) of (ENP)–(ID) converges to the function u (x) in \(C(\overline{\Omega })\) as t → , where the function u is given by

$$\displaystyle{u_{\infty }(x) =\liminf _{t\rightarrow \infty }u(x,t)\ \ \ \text{ for }\ x \in \overline{\Omega }.}$$

In this subsection, we do not assume (A9) +  or (A9) −  and give two characterizations of the function u .

Let \({\mathcal{S}}^{-}\) and \(\mathcal{S}\) denote the sets of all viscosity subsolutions of (SNP) and of all viscosity solutions of (SNP), respectively.

Theorem 6.3.

Set

$$\displaystyle\begin{array}{rcl} \mathcal{F}_{1}& =& \{v \in {\mathcal{S}}^{-}\, :\, v \leq u_{ 0}\mathrm{in}\ \overline{\Omega }\}, {}\\ u_{0}^{-}& =& \sup \mathcal{F}_{ 1}, {}\\ \mathcal{F}_{2}& =& \{w \in \mathcal{S}\, :\, w \geq u_{0}^{-}\mathrm{in}\ \overline{\Omega }\}. {}\\ \end{array}$$

Then   \(u_{\infty } =\inf \mathcal{F}_{2}\).

Proof.

By Proposition 1.10, we have \(u_{0}^{-}\in {\mathcal{S}}^{-}\). It is clear that \(u_{0}^{-}\leq u_{0}\) in \(\overline{\Omega }\). Hence, by Theorem 3.1 applied to the functions \(u_{0}^{-}\) and u, we get \(u_{0}^{-}(x) \leq u(x,t)\) for all (x, t) ∈ Q, which implies that \(u_{0}^{-}\leq u_{\infty }\) in \(\overline{\Omega }\). This together with Lemma 6.1 ensures that \(u_{\infty }\in \mathcal{F}_{2}\), which shows that \(\inf \mathcal{F}_{2} \leq u_{\infty }\) in \(\overline{\Omega }\).

Next, we set

$$\displaystyle{{u}^{-}(x,t) =\inf _{ r>0}u(x,t + r)\ \ \ \text{ for all }(x,t) \in \overline{Q}.}$$

By Lemma 5.11, the function u  −  is a solution of (ENP) and the function u  − ( ⋅, 0) is a viscosity subsolution of (SNP). Also, it is clear that \({u}^{-}(x,0) \leq u_{0}(x)\) for all \(x \in \overline{\Omega }\), which implies that \({u}^{-}(\cdot,0) \leq u_{0}^{-}\leq \inf \mathcal{F}_{2}\) in \(\overline{\Omega }\). We apply Theorem 3.1 to the functions u  −  and \(\inf \mathcal{F}_{2}\), to obtain  \({u}^{-}(x,t) \leq \inf \mathcal{F}_{2}(x)\) for all (x, t) ∈ Q, from which we get  \(u_{\infty }\leq \inf \mathcal{F}_{2}\) in \(\overline{\Omega }\), and conclude the proof. □ 

Let \(d\, :\,{ \overline{\Omega }\,}^{2} \rightarrow \mathbb{R}\) and \(\mathcal{A}\) denote the distance-like function and the Aubry set, respectively, as in Sect. 4.

Theorem 6.4.

We have the formula:

$$\displaystyle{u_{\infty }(x) =\inf \{ d(x,y) + d(y,z) + u_{0}(z)\, :\, z \in \overline{\Omega },\ y \in \mathcal{A}\}\ \ \ \text{ for all }\ x \in \overline{\Omega }.}$$

Proof.

We first show that

$$\displaystyle{u_{0}^{-}(x) =\inf \{ u_{ 0}(y) + d(x,y)\, :\, y \in \overline{\Omega }\}\ \ \ \text{ for all }\ x \in \overline{\Omega },}$$

where \(u_{0}^{-}\) is the function defined in Theorem 6.3.

Let \(u_{d}^{-}\) denote the function given by the right hand side of the above formula. Since \(u_{0}^{-}\in {\mathcal{S}}^{-}\), we have

$$\displaystyle{u_{0}^{-}(x) - u_{ 0}^{-}(y) \leq d(x,y)\ \ \ \text{ for all }\ x,y \in \overline{\Omega },}$$

which ensures that \(u_{0}^{-}\leq u_{d}^{-}\) in \(\overline{\Omega }\).

By Theorem 4.4, we have  \(u_{d}^{-}\in {\mathcal{S}}^{-}\). Also, by the definition of \(u_{d}^{-}\), we have  \(u_{d}^{-}(x) \leq u_{0}(x) + d(x,x) = u_{0}(x)\) for all \(x \in \overline{\Omega }\). Hence, by the definition of \(u_{0}^{-}\), we find that \(u_{0}^{-}\geq u_{d}^{-}\) in \(\overline{\Omega }\). Thus, we have  \(u_{0}^{-} = u_{d}^{-}\) in \(\overline{\Omega }\).

It is now enough to show that

$$\displaystyle{u_{\infty }(x) =\inf _{y\in \mathcal{A}}(u_{0}^{-}(y) + d(x,y)).}$$

Let ϕ denote the function defined by the right hand side of the above formula. The version of Proposition 1.10 for supersolutions ensures that \(\phi \in {\mathcal{S}}^{+}\), while Theorem 4.4 guarantees that \(\phi \in {\mathcal{S}}^{-}\). Hence, we have  \(\phi \in \mathcal{S}\). Observe also that

$$\displaystyle{u_{0}^{-}(x) \leq u_{ 0}^{-}(y) + d(x,y)\ \ \ \text{ for all }\ x,y \in \overline{\Omega },}$$

which yields  \(u_{0}^{-}\leq \phi\) in \(\overline{\Omega }\). Thus, we see by Theorem 6.3 that u  ≤ ϕ in \(\overline{\Omega }\).

Now, applying Theorem 4.1 to u , we observe that for all \(x \in \overline{\Omega }\),

$$\displaystyle\begin{array}{rcl} u_{\infty }(x)& =& \,\inf \{u_{\infty }(y) + d(x,y)\, :\, y \in \mathcal{A}\} {}\\ &\geq & \,\inf \{u_{0}^{-}(y) + d(x,y)\, :\, y \in \mathcal{A}\} =\phi (x). {}\\ \end{array}$$

Thus we find that u  = ϕ in \(\overline{\Omega }\). The proof is complete. □ 

Combining the above theorem and Proposition 5.4, we obtain another representation formula for u .

Corollary 6.1.

The following formula holds:

$$\displaystyle\begin{array}{rcl} u_{\infty }(x)& =& \inf \big\{\mathcal{L}(T,\eta,v,l) + u_{0}(\eta (T))\, :\, T > 0,\ (\eta,v,l) \in \mathrm{SP}(x) {}\\ & & \qquad \qquad \text{ such that }\ \eta (t) \in \mathcal{A}\ \text{ for some }t \in (0,\,T)\big\}. {}\\ \end{array}$$

Example 6.2.

As in Example 3.1, let n = 1, Ω = ( − 1, 1) and γ = ν on ∂Ω (i.e., γ( ± 1) =  ± 1). Let H = H(p) =  | p | 2 and \(g\, :\, \partial \Omega \rightarrow \mathbb{R}\) be the function given by g( − 1) =  − 1 and g(1) = 0. As in Example 3.1, we see that c # = 1. We set \(\tilde{H}(p) = H(p) - {c}^{\#} = \vert p{\vert }^{2} - 1\). Note that \(\tilde{H}\) satisfies both (A9) ± , and consider the Neumann problem

$$\displaystyle{ \tilde{H}(v^{\prime}(x)) = 0\ \ \mathrm{in}\ \ \Omega,\qquad \gamma (x) \cdot v^{\prime}(x) = g(x)\ \ \mathrm{on}\ \ \partial \Omega. }$$
(127)

It is easily seen that the distance-like function \(d\, :\,{ \overline{\Omega }}^{2} \rightarrow \mathbb{R}\) for this problem is given by d(x, y) =  | x − y | . Let \(\mathcal{A}\) denote the Aubry set for problem (127). By examining the function d, we see that \(\mathcal{A} =\{ -1\}\). For instance, by observing that

$$\displaystyle{D_{x}^{-}d(x,-1) = \left \{\begin{array}{@{}l@{\quad }l@{}} \{1\} \quad &\mathrm{if}\ x \in \Omega, \\ (-\infty,\,1]\quad &\mathrm{if}\ x = -1, \\{} [1,\,\infty ) \quad &\mathrm{if}\ x = 1, \end{array} \right.}$$

we find that \(-1 \in \mathcal{A}\). Let u 0(x) = 0. Consider the problem

$$\displaystyle{ \left \{\begin{array}{llll} &u_{t}(x,t) + H(u_{x}(x,t)) = 0\ \ &&\text{ for }\ (x,t) \in \Omega \times \mathbb{R}_{+}, \\ &\gamma (x)u_{x}(x,t) = g(x)\ \ &&\text{ for }\ (x,t) \in \partial \Omega \times \mathbb{R}_{+}, \\ &u(x,0) = u_{0}(x)\ \ &&\text{ for }x \in \overline{\Omega }. \end{array} \right. }$$

If u is the viscosity solution of this problem and the function v is given by v(x, t) = u(x, t) + c # t = u(x, t) + t, then v solves in the viscosity sense

$$\displaystyle{ \left \{\begin{array}{llll} &v_{t}(x,t) +\tilde{ H}(v_{x}(x,t)) = 0\ \ &&\text{ for }\ (x,t) \in \Omega \times \mathbb{R}_{+}, \\ &\gamma (x)v_{x}(x,t) = g(x)\ \ &&\text{ for }\ (x,t) \in \partial \Omega \times \mathbb{R}_{+}, \\ &v(x,0) = u_{0}(x)\ \ &&\text{ for }x \in \overline{\Omega }. \end{array} \right. }$$

Setting

$$\displaystyle{u_{\infty }(x) =\min \{ d(x,y) + d(y,z) + u_{0}(z)\, :\, y \in \mathcal{A},\,z \in \overline{\Omega }\}\ \ \text{ for }\ x \in \overline{\Omega },}$$

we note that u (x) =  | x + 1 | for all \(x \in \overline{\Omega }\). Thanks to Theorems 6.1 and 6.4, we have

$$\displaystyle{\lim _{t\rightarrow \infty }v(x,t) = u_{\infty }(x)\ \ \text{ uniformly on }\ \overline{\Omega },}$$

which reads

$$\displaystyle{\lim _{t\rightarrow \infty }(u(x,t) + t -\vert x + 1\vert ) = 0\ \ \text{ uniformly on }\ \overline{\Omega }.}$$

That is, we have u(x, t) ≈ − t +  | x + 1 | as t → . If we replace u 0(x) = 0 by the function u 0(x) =  − 3x, then

$$\displaystyle{u_{\infty }(x) =\min _{y\in \overline{\Omega }}\{\vert x + 1\vert + \vert 1 + y\vert - 3y\} = \vert x + 1\vert - 1\ \ \text{ for all }\ x \in \overline{\Omega },}$$

and u(x, t) ≈ − t +  | x + 1 | − 1 as t → .

In some cases the variational formula in Corollary 6.1 is useful to see the convergence assertion of Theorem 6.1.

Under the hypothesis that c # = 0, which is our case, we call a point \(y \in \overline{\Omega }\) an equilibrium point if L(y, 0) = 0. This condition, L(y, 0) = 0, is equivalent to \(\min _{p\in {\mathbb{R}}^{n}}H(y,p) = 0\).

Let \(y \in \overline{\Omega }\) be an equilibrium point. If we define (η, v, l) ∈ SP(y) by setting (η, v, l)(s) = (y, 0, 0), then \(\mathcal{L}(t,\eta,v,l) = 0\) for all \(t \in \mathbb{R}_{+}\), and Propositions 5.4 and 5.5 guarantee that \(y \in \mathcal{A}\).

We now assume that \(\mathcal{A}\) consists of only equilibrium points. Fix any \(\varepsilon > 0\) and \(x \in \overline{\Omega }\). According to Corollary 6.1, we can choose \(\tau,\sigma \in \mathbb{R}_{+}\) and (η, v, l) ∈ SP(x) so that \(\eta (\tau ) \in \mathcal{A}\) and

$$\displaystyle{ \mathcal{L}(\tau +\sigma,\eta,v,l) + u_{0}(\eta (\tau +\sigma )) < u_{\infty }(x) +\varepsilon. }$$
(128)

Fix any t > τ + σ. We define \((\tilde{\eta },\tilde{v},\tilde{l}) \in \mathrm{SP}(x)\) by

$$\displaystyle{(\tilde{\eta },\tilde{v},\tilde{l})(s) = \left \{\begin{array}{@{}l@{\quad }l@{}} (\eta,v,l)(s) \quad &\text{ for }\ s \in [0,\tau ), \\ (\eta (\tau ),0,0) \quad &\text{ for }\ s \in [\tau,\tau +\theta ), \\ (\eta,v,l)(s-\theta )\quad &\text{ for }\ s \in [\tau +\theta,\infty ), \end{array} \right.}$$

where θ = t − (τ + σ). Using (128), we get

$$\displaystyle{u_{\infty }(x)+\varepsilon > \mathcal{L}(t,\tilde{\eta },\tilde{v},\tilde{l}) + u_{0}(\tilde{\eta }(t)) \geq u(x,t).}$$

Therefore, recalling that liminf t →  u(x, t) = u (x), we see that lim t →  u(x, t) = u (x) for all \(x \in \overline{\Omega }\).

6.4 Localization of Conditions (A9) ± 

In this subsection we explain briefly that the following versions of (A9) ±  localized to the Aubry set \(\mathcal{A}\) may replace the role of (A9) ±  in Theorem 6.1.

  • (A10) ±  Let

    $$\displaystyle{Z_{\mathcal{A}} =\{ (x,p) \in \mathcal{A}\times {\mathbb{R}}^{n}\, :\, H(x,p) = 0\}.}$$

    There exists a function ω 0 ∈ C([0, )) satisfying ω 0(r) > 0 for all r > 0 such that if \((x,p) \in Z_{\mathcal{A}}\), \(\xi \in D_{p}^{-}H(x,p)\) and \(q \in {\mathbb{R}}^{n}\), then

    $$\displaystyle{H(x,p + q) \geq \xi \cdot q +\omega _{0}((\xi \cdot q)_{\pm }).}$$

As before, assume that c # = 0 and let u be the solution of (ENP)–(ID) and u (x) : = liminf t →  u(x, t).

Theorem 6.5.

Assume that either (A10)+ or (A10) holds. Then

$$\displaystyle{ \lim _{t\rightarrow \infty }u(x,t) = u_{\infty }(x)\ \ \ \text{ uniformly on }\ \overline{\Omega }. }$$
(129)

If we set

$$\displaystyle{u_{\infty }^{+}(x) =\limsup _{ t\rightarrow \infty }u(x,t)\ \ \ \text{ for }\ x \in \overline{\Omega },}$$

we see by Theorem 1.3 that the function \(u_{\infty }^{+}(x)\) is a subsolution of (ENP), as a function of (x, t), and hence a subsolution of (SNP). That is, \(u_{\infty }^{+} \in {\mathcal{S}}^{-}\). Since \(u_{\infty }\in {\mathcal{S}}^{+}\), once we have shown that \(u_{\infty }^{+} \leq u_{\infty }\) on \(\mathcal{A}\), then, by Theorem 4.6, we get

$$\displaystyle{u_{\infty }^{+} \leq u_{ \infty }\ \ \ \mathrm{in}\ \ \overline{\Omega },}$$

which shows that the uniform convergence (129) is valid. Thus we only need to show that \(u_{\infty }^{+} \leq u_{\infty }\) on \(\mathcal{A}\).

Following [21] (see also [39]), one can prove the following lemma.

Lemma 6.4.

For any \(z \in \mathcal{A}\) there exists an α = (η,v,l) ∈SP (z) such that

$$\displaystyle{d(z,\eta (t)) = \mathcal{L}(t,\alpha ) = -d(\eta (t),z)\ \ \ \text{ for all }\ t > 0.}$$

Proof.

By Proposition 5.5, for each \(k \in \mathbb{N}\) there are an \(\alpha _{k} = (\eta _{k},v_{k},l_{k}) \in \mathrm{SP}(z)\) and τ k  ≥ k such that

$$\displaystyle{\mathcal{L}(\tau _{k},\alpha _{k}) < \frac{1} {k}\ \ \ \text{ and }\ \ \ \eta _{k}(\tau _{k}) = z.}$$

Observe that for any \(j,k \in \mathbb{N}\) with j < k,

$$\displaystyle\begin{array}{rcl} \frac{1} {k}& >& \mathcal{L}(j,\alpha _{k}) +\int _{ j}^{\tau _{k} }[L(\eta _{k}(s),-v_{k}(s)) + l_{k}(s)g(\eta _{k}(s))]\mathrm{d}s \\ & \geq & \mathcal{L}(j,\alpha _{k}) + d(\eta _{k}(j),\eta _{k}(\tau _{k})), {}\end{array}$$
(130)

and hence

$$\displaystyle{\sup _{k\in \mathbb{N}}\mathcal{L}(j,\alpha _{k}) < \infty \ \ \ \text{ for all }\ j \in \mathbb{N}.}$$

We apply Theorem 5.4, with \(T = j \in \mathbb{N}\), and use the diagonal argument, to conclude from (130) that there is an α = (η, v, l) ∈ SP(z) such that for all \(j \in \mathbb{N}\),

$$\displaystyle{\mathcal{L}(j,\alpha ) \leq \liminf _{k\rightarrow \infty }\mathcal{L}(j,\alpha _{k}) \leq -d(\eta (j),z).}$$

Let 0 < t < , and choose a \(j \in \mathbb{N}\) such that t < j. Using Propositions 5.4 and 4.1 (ii) (the triangle inequality for d), we compute that

$$\displaystyle\begin{array}{rcl} d(z,\eta (t))& \leq & \mathcal{L}(t,\alpha ) = \mathcal{L}(j,\alpha ) -\int _{t}^{j}[L(\eta (s),-v(s)) + l(s)g(\eta (s))]\mathrm{d}s {}\\ & \leq & \mathcal{L}(j,\alpha ) - d(\eta (t),\eta (j)) \leq -d(\eta (j),z) - d(\eta (t),\eta (j)) {}\\ & \leq & -d(\eta (t),z). {}\\ \end{array}$$

Moreover, by the triangle inequality, we get

$$\displaystyle{-d(\eta (t),z) \leq d(z,\eta (t)).}$$

These together yield

$$\displaystyle{d(z,\eta (t)) = \mathcal{L}(t,\alpha ) = -d(\eta (t),z)\ \ \ \text{ for all }t > 0,}$$

which completes the proof. □ 

The above assertion is somehow related to the idea of the quotient Aubry set (see [46, 47]). Indeed, if we introduce the equivalence relation \(\equiv \) on \(\mathcal{A}\) by

$$\displaystyle{x \equiv y\ \ \;\Longleftrightarrow\;\ \ d(x,y) + d(y,x) = 0,}$$

and consider the quotient space \(\hat{\mathcal{A}}\) consisting of the equivalence classes

$$\displaystyle{[x] =\{ y \in \mathcal{A}\, :\, y \equiv x\},\ \ \text{ with }\ x \in \mathcal{A},}$$

then the space \(\hat{\mathcal{A}}\) is a metric space with its distance given by

$$\displaystyle{\hat{d}([x],[y]) = d(x,y) + d(y,x).}$$

The property of the curve η in the above lemma that d(z, η(t)) =  − d(η(t), z)  is now stated as:  η(t) ≡ η(0).

Lemma 6.5.

Let \(\psi \in {\mathcal{S}}^{-}\) and \(x,y \in \mathcal{A}\) . If x ≡ y, then

$$\displaystyle{\psi (x) -\psi (y) = d(x,y).}$$

Proof.

By the definition of d, we have

$$\displaystyle{\psi (x) -\psi (y) \leq d(x,y)\ \ \ \text{ and }\ \ \ \psi (y) -\psi (x) \leq d(y,x).}$$

Hence,

$$\displaystyle{\psi (x) -\psi (y) \leq d(x,y) = -d(y,x) \leq \psi (x) -\psi (y),}$$

which shows that ψ(x) − ψ(y) = d(x, y) =  − d(y, x). □ 

Proof (Theorem 6.5). 

As we have noted above, we need only to show that

$$\displaystyle{u_{\infty }^{+}(x) \leq u_{ \infty }(x)\ \ \ \text{ for all }\ x \in \mathcal{A}.}$$

To this end, we fix any \(z \in \mathcal{A}\). Let α = (η, v, l) ∈ SP(z) be as in Lemma 6.4. In view of Lemma 6.5, we have

$$\displaystyle{u_{\infty }(z) - u_{\infty }(\eta (t)) = d(z,\eta (t)) = \mathcal{L}(t,\alpha )\ \ \ \text{ for all }\ t > 0.}$$

It is obvious that the same assertion as Lemma 6.3 holds if we replace S by

$$\displaystyle{S_{\mathcal{A}} :=\{ (x,\xi ) \in \mathcal{A}\times {\mathbb{R}}^{n}\, :\,\xi \in D_{ p}^{-}H(x,p)\ \ \text{ for some }(x,p) \in Z_{ \mathcal{A}}\}.}$$

We now just need to follow the arguments in Sect. 6.2, to conclude that

$$\displaystyle{u_{\infty }^{+}(z) \leq u_{ \infty }(z).}$$

The details are left to the interested reader. □