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

Discontinuous eikonal equations
in metric measure spaces

Qing Liu Qing Liu, Geometric Partial Differential Equations Unit, Okinawa Institute of Science and Technology Graduate University, 1919-1, Tancha Onna-son, Okinawa 904-0495, Japan
Email: qing.liu@oist.jp
Nageswari Shanmugalingam Nageswari Shanmugalingam, Department of Mathematical Sciences, University of Cincinnati, P.O.Box 210025, Cincinnati, OH 45221-0025, USA
Email: shanmun@uc.edu
 and  Xiaodan Zhou Xiaodan Zhou, Analysis on Metric Spaces Unit, Okinawa Institute of Science and Technology G raduate University, 1919-1 Tancha, Onna-son, Okinawa 904-0495, Japan
Email: xiaodan.zhou@oist.jp
Abstract.

In this paper, we study the eikonal equation in metric measure spaces, where the inhomogeneous term is allowed to be discontinuous, unbounded and merely pp-integrable in the domain with a finite pp. For continuous eikonal equations, it is known that the notion of Monge solutions is equivalent to the standard definition of viscosity solutions. Generalizing the notion of Monge solutions in our setting, we establish uniqueness and existence results for the associated Dirichlet boundary problem. The key in our approach is to adopt a new metric, based on the optimal control interpretation, which integrates the discontinuous term and converts the eikonal equation to a standard continuous form. We also discuss the Hölder continuity of the unique solution with respect to the original metric under regularity assumptions on the space and the inhomogeneous term.

Key words and phrases:
eikonal equation, metric spaces, discontinuous inhomogeneous term, viscosity solutions
2010 Mathematics Subject Classification:
35R15, 49L25, 35F30, 35D40

1. Introduction

1.1. Background and motivation

In this paper we study the eikonal equation in a metric measure space (𝐗,d,μ)({\mathbf{X}},d,\mu) with a possibly discontinuous inhomogenous term, where μ\mu is assumed to be a nonnegative, locally finite Borel measure. We assume that 𝐗{\mathbf{X}} is a complete length space and Ω𝐗\Omega\subsetneq{\mathbf{X}} is a bounded domain. We consider

|u|(x)=f(x)xΩ|\nabla u|(x)=f(x)\quad\text{$x\in\Omega$} (1.1)

with the Dirichlet boundary condition

u=gon Ω,u=g\quad\text{on $\partial\Omega$,} (1.2)

where ff is a given positive Borel measurable function in Ω¯\overline{\Omega} and gg is a given bounded function on Ω\partial\Omega. Further assumptions on ff and gg will be introduced later.

Well-posedness of the eikonal equation and more general Hamilton-Jacobi equations is a classical topic. The theory of viscosity solutions provides a successful tool to solve such fully nonlinear PDEs in the Euclidean space. It is well known that when Ωn\Omega\subset{\mathbb{R}}^{n} and fC(Ω¯)f\in C(\overline{\Omega}) satisfies the lower bound condition

α:=infΩf>0,\alpha:=\inf_{\Omega}f>0, (1.3)

there exists a unique viscosity solution uC(Ω¯)u\in C(\overline{\Omega}) of (1.1) with Dirichlet data (1.2) under appropriate regularity condition on gg. The uniqueness result is obtained by establishing a comparison principle while the existence of solutions can be proved via several different approaches including Perron’s method and a control-based formula.

Let us give more details on the connection with the optimal control theory, as it largely motivates our exploration and plays a central role in this work. It is widely known [40, 6] that, under appropriate conditions on ff and gg, the unique viscosity solution uu of (1.1) with the Dirichlet boundary condition (1.2) in the Euclidean space can be expressed by

u(x)=inf{g(y)+γf𝑑s:γ is a curve in Ω¯ joining xΩ¯ and yΩ}.u(x)=\inf\left\{g(y)+\int_{\gamma}f\,ds:\ \text{$\gamma$ is a curve in $\overline{\Omega}$ joining $x\in\overline{\Omega}$ and $y\in\partial\Omega$}\right\}. (1.4)

This formula represents the so-called value function of an optimal control problem, where one tries to move a point from xΩ¯x\in\overline{\Omega} toward the boundary and minimize the total payoff comprising a running cost ff integrated along the trajectory and a terminal cost gg at the first hit on the boundary. See [38, 31] and references therein for more recent developments on representation formulas for general Hamilton-Jacobi equations.

The representation formula not only provides a clear understanding of the solution and promotes various control-related practical applications, but also suggests possible generalization of the theory for a possibly discontinuous function ff. In fact, much progress has been made in this direction in the Euclidean space. The study on discontinuous Hamilton-Jacobi equations in the Euclidean space was initiated in the work [36] and later found applications in geometric optics with different layers, image processing, shape from shading, etc. Further development in different contexts to treat discontinuous Hamiltonians including the time-dependent case can be found in [55, 47, 48, 53, 12, 11, 19, 8, 13, 54, 17, 27, 28]. The aforementioned papers in the Euclidean case, reduced to our eikonal problem, seem to need at least local essential boundedness of ff for their well-posedness results except for the papers [55, 53, 54]. In these three papers, the results regarding uniqueness and existence results rely on regularity conditions on the discontinuity set of ff or certain controllability assumptions for the optimal control formulation. Such conditions can hardly be extended further to our relaxed geometric setting due to the lack of smooth structure in general metric spaces.

In the present paper we aim to study in a direct manner the eikonal equation with potentially wilder discontinuity than those considered in [55, 53, 54]. We drop the local boundedness of the inhomogeneous data ff and allow it to be as discontinuous as a function in Lp(Ω)L^{p}(\Omega). It is worth adding that an important tool called LpL^{p} viscosity solution theory has been developed for second order uniformly elliptic equations with measurable data in [10, 18], but it does not seem to apply to the first order case. For the Dirichlet problem (1.1)(1.2), our primary observation is that the formula (1.4) actually requires nothing more than the path integrability of ff along a curve connecting xx to the boundary, which is a much weaker assumption than the continuity or even boundedness of ff. Let us present a simple one-dimensional example to clarify this point.

Let Ω=(1,1)\Omega=(-1,1)\subset{\mathbb{R}} and set f:Ωf:\Omega\to{\mathbb{R}} such that

f(x)=1|x|,x[1,1]{0}.f(x)=\frac{1}{\sqrt{|x|}},\quad x\in[-1,1]\setminus\{0\}. (1.5)

The value f(0)f(0) can be chosen to be any positive real number in order to meet the requirement (1.3). Such type of singular eikonal equations finds applications in lens design for wireless communication systems [16]. It is clear that fLp(Ω)f\in L^{p}(\Omega) with p[1,2)p\in[1,2). Therefore the formula in (1.4) still makes sense for given boundary data gg at ±1\pm 1. For instance, setting g(±1)=0g(\pm 1)=0, we see that the formula immediately yields

u(x)=2(1|x|)for x[1,1].u(x)=2(1-\sqrt{|x|})\quad\text{for $x\in[-1,1]$.} (1.6)

It seems to be the correct solution of (1.1)(1.2), for it satisfies the standard definition of viscosity solutions if we consider f(0)=f(0)=\infty. Also, one can verify the validity of approximation via truncation of the running cost. Indeed, taking fM=min{f,M}f_{M}=\min\{f,M\} for M>0M>0 large, we can obtain a unique viscosity solution for the truncated problem with the bounded function fMf_{M}. By direct calculations, we easily see that the approximate solution converges uniformly to uu in [1,1][-1,1]. Moreover, although in general we cannot expect Lipschitz regularity of solutions due to the loss of boundness of ff, the example shows that the solution enjoys Hölder regularity that depends on the integrability of ff. This can be viewed as a natural consequence corresponding to the Morrey-Sobolev inequality, as the eikonal equation suggests uW1,p(Ω)u\in W^{1,p}(\Omega) at least formally.

To the best of our knowledge, the well-posedness and regularity issue for this type of discontinuous eikonal equations have not been carefully examined even in the Euclidean case. We are therefore motivated to tackle these problems in a general way that permits applications to a broad class of metric measure spaces with length structure. We develop the notion of Monge solutions used in [47, 8, 43] and show uniqueness and existence of solutions in this general setting. We also establish Hölder regularity of the solution under certain regularity assumptions on ff and 𝐗{\mathbf{X}}. More details about our results will be given momentarily.

First-order Hamilton-Jacobi equations in metric spaces have recently attracted great attention for applications in optimal transport [5, 57], mean field games [14], topological networks [49, 35, 1, 33, 34] etc. Concerning the equations that depend on the gradient in terms of its norm, we refer to [4, 29, 26, 45] for several different notions of viscosity solutions as well as the associated uniqueness and existence results. In our previous paper [43], we proved the equivalence of these solutions and introduce another alternative notion, extending the approach of Monge solutions proposed in [47] to the eikonal equation in complete length spaces; see also an application of this Monge-type notion to the study of the first eigenvalue problem for the infinity Laplacian in metric spaces [41]. Our current work further develops the notion of Monge solutions so that viscosity solution theory will be available to handle a class of nonlinear equations in general metric measure spaces with data having a wider range of discontinuities.

Our work is new even from the point of view of pure analysis. There has been great progress on various aspects of first order analysis on metric measure spaces, including first order Sobolev spaces and their relation to variational problems and partial differential equations. One central notion that generalizes the norm of the gradient of a Sobolev function in Euclidean spaces is the so-called upper gradient. We refer the reader to [30] for a survey of the Sobolev spaces theory built upon this notion. From this point of view, the Dirichlet problem (1.1) with boundary data (1.2) can be interpreted as an extension problem for a function uu to Ω¯\overline{\Omega} with boundary value gg assigned on Ω\partial\Omega and the upper gradient ff prescribed in Ω\Omega. In general, many possible choices can be expected, among which the viscosity solution constructed by the formula (1.4) is usually considered to be the most “physical”. This perspective can also be found in the work [15]. There is certainly no need to restrict the prescribed gradient to a continuous or bounded function class. We are thus inspired to address the extension problem for fLp(Ω)f\in L^{p}(\Omega) and investigate properties of the extended function uu.

1.2. General results on existence and uniqueness of solutions

Our presentation can be divided into two parts. The first part, consisting of Sections 24, is a general PDE theory with only the length structure and curve-wise integrals of ff involved. In our main well-posedness results, we impose certain key conditions on the closure of the domain Ω¯\overline{\Omega} and the inhomogeneous term ff. In the second part, adding a measure structure to the space, we provide more specific sufficient conditions for those assumptions on Ω¯\overline{\Omega} and ff.

One of the crucial steps in the first part is to find an appropriate notion of viscosity solutions in this setting that agrees with the representation formula (1.4) and justifies a comparison principle for us to prove the uniqueness. As mentioned above, we adopt the notion of Monge solutions, which in the case of a continuous ff, requires

|u|(x)=f(x)for all xΩ,|\nabla^{-}u|(x)=f(x)\quad\text{for all $x\in\Omega$}, (1.7)

where, for a locally Lipschitz function uu, the subslope |u||\nabla^{-}u| is defined by

|u|(x)=lim supd(x,y)0max{u(x)u(y),0}d(x,y).|\nabla^{-}u|(x)=\limsup_{d(x,y)\to 0}\frac{\max\{u(x)-u(y),0\}}{d(x,y)}.

In the Euclidean space, this definition is known to be equivalent to the usual viscosity solutions [47]. One of its advantages is that it avoids using smooth test functions, whose lack of availability is indeed a key difficulty in the study of Hamilton-Jacobi equations in general metric spaces.

When ff is of class Lp(Ω)L^{p}(\Omega), the definition in (1.7) may not apply directly. For example, if one changes the value of f(0)f(0) to be finite in the aforementioned example, then the expected solution (1.6) does not satisfy definition (1.7) at 0. We thus need to somehow exploit collective information of ff rather than its pointwise value. Also, we always assume that the lower bound condition (1.3) holds for ff in order to obtain uniqueness. Heuristically speaking, since ff is uniformly positive in Ω\Omega, we can rewrite the equation (1.1) as

|u|(x)f(x)=1in Ω\frac{|\nabla u|(x)}{f(x)}=1\quad\text{in $\Omega$}

and adopt a new pointwise slope by incorporating the value of f(x)f(x) into the metric. This approach enables us to study the eikonal equation as if the right hand side is a constant. Such connection can be rigorously realized by introducing the following new metric, which is also called optical length function in [47, 8] etc.,

Lf(x,y):=inf{γf𝑑s:γΓf(x,y)}for any x,yΩ¯,L_{f}(x,y):=\inf\left\{\int_{\gamma}f\,ds:\ \text{$\gamma\in\Gamma_{f}(x,y)$}\right\}\quad\text{for any $x,y\in\overline{\Omega}$,} (1.8)

where Γf(x,y)\Gamma_{f}(x,y) denotes the collection of all curves γ\gamma connecting xx and yy in Ω¯\overline{\Omega} on which the path integral is finite, namely,

Γf(x,y)={γ:[0,]Ω¯:γf𝑑s<,γ(0)=x,γ()=y and |γ|(s)=1 for a.e. s}.\Gamma_{f}(x,y)=\left\{\gamma:[0,\ell]\to\overline{\Omega}:\int_{\gamma}f\,ds<\infty,\ \gamma(0)=x,\gamma(\ell)=y\text{ and }|\gamma^{\prime}|(s)=1\text{ for a.e. }s\right\}.

We say that the curves γ\gamma in Γf(x,y)\Gamma_{f}(x,y) are admissible with respect to ff. We shall assume throughout Sections 24 that

Γf(x,y)for all x,yΩ¯\Gamma_{f}(x,y)\neq\emptyset\quad\text{for all $x,y\in\overline{\Omega}$} (1.9)

so that the formula in (1.4) is well defined for any xΩ¯x\in\overline{\Omega}. In Section 5, we discuss some sufficient assumptions on Ω¯\overline{\Omega} implying (1.9).

Adopting the subslope with respect to this newly defined metric LfL_{f}, defined by

|fu|(x):=lim supLf(x,y)0max{u(x)u(y),0}Lf(x,y)for xΩ,|\nabla_{f}^{-}u|(x):=\limsup_{L_{f}(x,y)\to 0}\frac{\max\{u(x)-u(y),0\}}{L_{f}(x,y)}\quad\text{for $x\in\Omega$,} (1.10)

one can simply define the Monge solution of (1.1) in Definition 3.1 by asking that uu satisfies

|fu|(x)=1for all xΩ.|\nabla_{f}^{-}u|(x)=1\quad\text{for all $x\in\Omega$.} (1.11)

The same idea of changing metrics was also outlined in [9] in connection with metric geometry and in [56] for applications in homogenization.

This new definition of Monge solutions is consistent with the well-studied case where ff is continuous. In fact, if ff is continuous at x0Ωx_{0}\in\Omega, then

Lf(x,x0)d(x,x0)f(x0)as d(x,x0)0.\frac{L_{f}(x,x_{0})}{d(x,x_{0})}\to f(x_{0})\quad\text{as $d(x,x_{0})\to 0$.} (1.12)

It is then easily seen that |fu|(x0)=1|\nabla_{f}^{-}u|(x_{0})=1 and |u|(x0)=f(x0)|\nabla^{-}u|(x_{0})=f(x_{0}) are equivalent.

Switching the metric from dd to LfL_{f} enables us to reduce the possibly discontinuous inhomogeneous data in (1.7) to the continuous standard form (1.11). As a consequence, most of our arguments in [43] can be applied to establish the comparison principle and prove existence of solutions of (1.1), (1.2) under the compatibility condition

g(x)Lf(x,y)+g(y)for all x,yΩ.g(x)\leq L_{f}(x,y)+g(y)\quad\text{for all $x,y\in\partial\Omega$.} (1.13)

However, a major challenge concerning the topological change of the space appears in the proof of comparison principle. Note that the metrics LfL_{f} and dd are not topologically equivalent in general, especially when ff is unbounded. One can find examples where d(x,y)0d(x,y)\to 0 fails to imply Lf(x,y)0L_{f}(x,y)\to 0; see Example 2.4. For most of our analysis, we impose the following key compatibility condition

sup{Lf(x,y):x,yΩ¯,d(x,y)r}0as r0,\sup\{L_{f}(x,y):x,y\in\overline{\Omega},\ d(x,y)\leq r\}\to 0\quad\text{as $r\to 0$}, (1.14)

which ensures that the topology induced by LfL_{f} is in agreement with the metric topology inherited from (𝐗,d)({\mathbf{X}},d), and enables us to obtain not only the uniqueness but also the uniform continuity of Monge solutions.

In general, if (1.14) fails to hold, then, because of the change in the metric, the domain Ω\Omega that is bounded in dd may become unbounded in LfL_{f}. The notions of interior and boundary points of Ω\Omega may also change accordingly. However, it is still possible to show uniqueness and existence of Monge solutions if we have boundedness of Ω¯\overline{\Omega} with respect to LfL_{f} as well as a weaker version of (1.14) as below:

sup{Lf(x,y):x,yΩ¯Ωr,d(x,y)r}0as r0,\sup\{L_{f}(x,y):x,y\in\overline{\Omega}\setminus\Omega_{r},\ d(x,y)\leq r\}\to 0\quad\text{as $r\to 0$}, (1.15)

where Ωr={xΩ:d(x,Ω)>r}\Omega_{r}=\{x\in\Omega:d(x,\partial\Omega)>r\}. Here, we denote d(x,Ω)=infyΩd(x,y)d(x,\partial\Omega)=\inf_{y\in\partial\Omega}d(x,y). In this case, we can get a unique Monge solution that is Lipschitz continuous in LfL_{f} but possibly discontinuous in dd; see Example 4.4 for a concrete example. Note that the completeness of Ω¯\overline{\Omega} is preserved under our metric change (Lemma 2.5), which enables us to establish a comparison principle based on Ekeland’s variational principle.

Our main result can be summarized as follows. Its proof is given in Section 4.

Theorem 1.1 (Existence and uniqueness of solutions).

Let (𝐗,d)({\mathbf{X}},d) be a complete length space and Ω𝐗\Omega\subsetneq{\mathbf{X}} be a bounded domain. Assume that (1.3) and (1.9) hold. Let LfL_{f} be given by (1.8), and g:Ωg:\partial\Omega\to{\mathbb{R}} be bounded. Consider the function uu defined by

u(x):=infyΩ{Lf(x,y)+g(y)},xΩ¯.u(x):=\inf_{y\in\partial\Omega}\{L_{f}(x,y)+g(y)\},\quad x\in\overline{\Omega}. (1.16)

Then the following results hold.

  1. (i)

    uu is Lipschitz in Ω¯\overline{\Omega} with respect to LfL_{f} and is a Monge solution of (1.1).

  2. (ii)

    If gg further satisfies the compatibility condition (1.13), (Ω¯,Lf)(\overline{\Omega},L_{f}) is bounded, and (1.15) holds, then uu is the unique Monge solution of the Dirichlet problem (1.1), (1.2) with (1.2) interpreted as

    sup{|u(x)g(y)|:xΩ¯,yΩ,d(x,y)δ}0as δ0.\sup\{|u(x)-g(y)|:x\in\overline{\Omega},\ y\in\partial\Omega,\ d(x,y)\leq\delta\}\to 0\quad\text{as $\delta\to 0$}. (1.17)
  3. (iii)

    If gg satisfies (1.13), and (1.14) holds, then uu is the unique Monge solution, uniformly continuous with respect to dd, of the Dirichlet problem (1.1), (1.2) with (1.2) interpreted as (1.17).

The interpretation (1.17) is automatically guaranteed by (1.2) if (Ω¯,Lf)(\overline{\Omega},L_{f}) is assumed to be compact. Our comparison results with and without compactness of (Ω¯,Lf)(\overline{\Omega},L_{f}) are presented in Section 3.2. We also include a discussion in Section 4.2 about the case when the compatibility condition (1.13) on gg fails to hold. In the Euclidean space, this corresponds to loss of Dirichlet boundary data, and one needs to relax the boundary condition in the viscosity sense; see for example [51, 52, 37] for the well-posedness results under certain regularity conditions on Ω\partial\Omega. Thanks to the general setting considered in this present paper, we treat this problem in a more direct way. If the boundary data gg on a part of Ω\partial\Omega is lost, we take the remaining piece, denoted by Σg\Sigma_{g}, as the real boundary and regard Ω¯Σg\overline{\Omega}\setminus\Sigma_{g} as the interior of the domain. In fact, the formula (1.4) does yield the Monge solution property of uu in Ω¯Σg\overline{\Omega}\setminus\Sigma_{g}. Such flexibility in changing the notion of boundary and interior points leads us to uniqueness and existence of solutions to the reduced Dirichlet problem.

1.3. Regularity of solutions and further discussions

The second half of our presentation is devoted to understanding more deeply the key assumptions (1.9) and (1.14) in Theorem 1.1. We study this regularity problem in the setting of a general metric measure space (𝐗,d,μ)({\mathbf{X}},d,\mu) with the measure μ\mu assumed to be doubling. We recall that a locally finite Borel measure μ\mu is doubling if there exists Cd1C_{d}\geq 1 such that μ(B2r(x))Cdμ(Br(x))\mu(B_{2r}(x))\leq C_{d}\mu(B_{r}(x)) for all r>0r>0, where Br(x)B_{r}(x) denotes the open metric ball centered at x𝐗x\in{\mathbf{X}} with radius r>0r>0, i.e.,

Br(x):={y𝐗:d(x,y)<r}.B_{r}(x):=\{y\in{\mathbf{X}}\,:\,d(x,y)<r\}.

We also call Q=log2CdQ=\log_{2}C_{d} the homogeneous dimension of the doubling metric measure space. It is not difficult to see that by the doubling property of μ\mu, we have, for any metric balls Br(x)BR(x)𝐗B_{r}(x)\subset B_{R}(x)\subset{\mathbf{X}},

μ(Br(x))μ(BR(x))C(rR)Q.\frac{\mu(B_{r}(x))}{\mu(B_{R}(x))}\geq C{\left(\frac{r}{R}\right)}^{Q}. (1.18)

We prove the property (1.14) and Hölder regularity with respect to the metric dd of the solution uu defined by (1.16) under either of the following two assumptions.

  • (A1)

    Ω¯\overline{\Omega} satisfies a pp-Poincaré inequality for some finite p>max{1,Q}p>\max\{1,Q\} and fLp(Ω¯)f\in L^{p}(\overline{\Omega}), where Q>0Q>0 is the homogeneous dimension from (1.18) above.

  • (A2)

    Ω¯\overline{\Omega} satisfies an \infty-Poincaré inequality and fL(Ω¯)f\in L^{\infty}(\overline{\Omega}).

Theorem 1.2 (Regularity of solutions).

Let (Ω¯,d,μ)(\overline{\Omega},d,\mu) be a complete bounded metric measure space with μ\mu a doubling measure. Let g:Ωg:\partial\Omega\to{\mathbb{R}} be bounded and uu be the Monge solution defined by (1.16). If Ω¯\overline{\Omega} and ff satisfy (A1), then uu is (1Qp)(1-\frac{Q}{p})-Hölder continuous in Ω¯\overline{\Omega} with respect to the metric dd. If Ω¯\overline{\Omega} and ff satisfy (A2), then uu is Lipschtiz continuous in Ω¯\overline{\Omega} with respect to the metric dd.

Here, the pp-Poincaré inequality is a condition that indicates the richness of curves of a space; see Section 5 for a more precise description and [30] for more detailed introduction. In our assumptions, a balance is taken between the regularity of the space Ω¯\overline{\Omega} and the function ff. In order to consider merely pp-integrable ff with a finite pp, we assume pp-Poincaré inequality in (A1), which is stronger than the \infty-Poincaré inequality in (A2),

Our results in Theorem 1.2 resemble the Morrey-Sobolev embedding theorem. In the proof, we apply a geometric characterization of the Poincaré inequality provided by [21] for both cases of (A1) and (A2) with adaptations for our eikonal equation. We can also show that uu is in the Sobolev class N1,p(Ω¯)N^{1,p}(\overline{\Omega}) under the conditions of Theorem 1.2. While (A1) and (A2) seem close to optimal for us to obtain (1.14), both of them are actually too strong to directly show (1.9), which is far weaker than (1.14) and simply requires the existence of one curve on which ff is integrable. We construct an example in 2{\mathbb{R}}^{2}, showing that (1.9) can be obtained even for some fL2(Ω)f\notin L^{2}(\Omega); see Example 5.4. It would be interesting to find more general sufficient conditions for (1.9).

It is worth remarking that the LpL^{p} class in our study is only used to describe the extent of discontinuity of ff and cannot be understood as the usual pp-integrable function space. Since the Monge solution obtained by (1.4) substantially depends on path integrals of ff, changing, especially decreasing, the value of ff even on a null set will drastically affect the solution. In other words, our Monge solution strongly depends on the choice of representatives of the equivalence class for fLp(Ω¯)f\in L^{p}(\overline{\Omega}).

Section 6 is devoted to discussions on how to alleviate the instability of solutions with respect to ff increasing on null sets. Our method generalizes the approach in [12, 8]. Instead of taking the solution for LpL^{p} function ff, we generate a class of solutions for different representatives from its equivalence class under the assumption (A1) or (A2). More precisely, for any null set NΩ¯N\subset\overline{\Omega}, we can find a Monge solution of (1.1)(1.2) with ff replaced by fN=f+χNf_{N}=f+\infty\chi_{N}, where χN\chi_{N} denotes the characteristic function of NN. This change in ff amounts to restricting ff to be integrable only on curves transversal to NN. Here, a curve is said to be transversal to NN if 1(γ1(N))=0\mathcal{H}^{1}(\gamma^{-1}(N))=0, where 1\mathcal{H}^{1} stands for the one-dimensional Hausdorff measure. By substituting Γf(x,y)\Gamma_{f}(x,y) for each pair of points x,yΩ¯x,y\in\overline{\Omega} with the class of transversal curves

ΓfN(x,y)={γ:[0,]Ω¯:\displaystyle{\Gamma}^{N}_{f}(x,y)=\bigg{\{}\gamma:[0,\ell]\to\overline{\Omega}\ : γf𝑑s<,γ(0)=x,γ()=y and\displaystyle\ \int_{\gamma}f\,ds<\infty,\,\gamma(0)=x,\gamma(\ell)=y\ \text{ and } (1.19)
|γ|(s)=1 for a.e. s,with1(γ1(N))=0}\displaystyle|\gamma^{\prime}|(s)=1\text{ for a.e. }s,\ \text{with}\ \mathcal{H}^{1}(\gamma^{-1}(N))=0\bigg{\}}

and taking the associated optical length function

LfN(x,y)=inf{γf𝑑s:γΓfN(x,y)},L^{N}_{f}(x,y)=\inf\left\{\int_{\gamma}f\,ds\,:\,\gamma\in{\Gamma}_{f}^{N}(x,y)\right\}, (1.20)

we can obtain the same existence and uniqueness results for any given null set NN.

We are particularly interested in the maximal solution over all null sets NN, as it corresponds to the choice of the largest function in the equivalence class of ff and represents the strongest possible instability. It can be defined by

u~(x)=infyΩ{L~f(x,y)+g(y)},xΩ¯,\tilde{u}(x)=\inf_{y\in\partial\Omega}\{\tilde{L}_{f}(x,y)+g(y)\},\quad x\in\overline{\Omega}, (1.21)

where L~f\tilde{L}_{f} is the maximal optical length function over all NN, that is,

L~f(x,y)=supμ(N)=0LfN(x,y)=supμ(N)=0inf{γf𝑑s:γΓfN(x,y)}.\tilde{L}_{f}(x,y)=\sup_{\mu(N)=0}L_{f}^{N}(x,y)=\sup_{\mu(N)=0}\inf\left\{\int_{\gamma}f\,ds\,:\,\gamma\in{\Gamma}_{f}^{N}(x,y)\right\}. (1.22)

We proved in Proposition 6.9 and Remark 6.11 that u~\tilde{u} is the maximal weak solution of (1.1), (1.2) under (A2) together with appropriate assumptions on Ω¯\overline{\Omega} and boundary data gg as well as almost everywhere continuity of ff. Here, the notion of weak solutions is a generalization of the Euclidean counterpart, requiring u~\tilde{u} to satisfy |u~|=f|\nabla\tilde{u}|=f almost everywhere in Ω\Omega, where |u~||\nabla\tilde{u}| denotes the slope of uu, i.e., for xΩx\in\Omega,

|u~|(x):=lim supd(x,y)0|u~(y)u~(x)|d(x,y).|\nabla\tilde{u}|(x):=\limsup_{d(x,y)\to 0}\frac{|\tilde{u}(y)-\tilde{u}(x)|}{d(x,y)}.

This result extends the discussion about the well-known maximal weak solution characterization of viscosity solutions in the Euclidean space [47, 8] to the setting of metric measure spaces and possibly unbounded discontinuous inhomogeneous term ff. It is however not clear to us how to handle the case when ff is not continuous almost everywhere. Another important open question is under what assumptions can we can eliminate the instability, i.e., u=u~u=\tilde{u} in Ω¯\overline{\Omega}. Some discussions regarding this issue can be found in [53] for the Euclidean space but the case for general metric spaces seems more difficult.

Let us conclude the introduction by mentioning that in this work we choose not to pursue more general Hamilton-Jacobi equations. One can easily generalize our method if a control-based formula is available and a new metric incorporating the discontinuity of the Hamiltonian can be found. Even for the eikonal equation itself, our general setup actually implicitly allows more general dependence of the Hamiltonian on the space variable. Typical examples of 𝐗{\mathbf{X}} include the sub-Riemannian manifolds. For instance, when (𝐗,d)({\mathbf{X}},d) is taken to be the first Heisenberg group with the Carnot-Carathéodory metric, the eikonal equation (1.1) can be written in the Euclidean coordinates as

(uxy2uz)2+(uy+x2uz)2=f(x,y,z)for (x,y,z)3.\sqrt{\left(u_{x}-{y\over 2}u_{z}\right)^{2}+\left(u_{y}+{x\over 2}u_{z}\right)^{2}}=f(x,y,z)\quad\text{for $(x,y,z)\in{\mathbb{R}}^{3}$.}

We refer to [7, 2, 3] etc. for discussions on the eikonal equation with f1f\equiv 1 in the sub-Riemannian setting.

Acknowledgements

The authors thank Professors Yoshikazu Giga, Nao Hamamuki and Atsushi Nakayasu for valuable comments on the first draft of the paper. The work of the first author was supported by JSPS Grant-in-Aid for Scientific Research (No. 19K03574, No. 22K03396). The work of the second author is partially supported by the grant DMS-#2054960 of the National Science Foundation (U.S.A.). The work of the third author was supported by JSPS Grant-in-Aid for Research Activity Start-up (No. 20K22315) and JSPS Grant-in-Aid for Early-Career Scientists (No. 22K13947).

2. Optical length function

The notion of optical length dates back to the paper of Houstoun [32], see also [50, (3)] in the context of the behavior of light rays in general relativity. Briani and Davini [8] use this notion to define the Monge solution for discontinuous Hamiltonians in the Euclidean space. See recent developments of this approach on Carnot groups in [25]. In the context of metric measure spaces, the notion of optical length is constructed in [15], but there the terminology of optical length was not used. Other applications can be found in [30, 39].

We define the optical length function Lf:Ω¯×Ω¯L_{f}:\overline{\Omega}\times\overline{\Omega}\to{\mathbb{R}} as in (1.8). Hereafter, we sometimes adopt the notation If(γ):=γf𝑑sI_{f}(\gamma):=\int_{\gamma}f\,ds for a rectifiable curve γ\gamma. Under our standing assumption (1.9), it is not difficult to see that LfL_{f} satisfies all the axioms of a metric on Ω¯\overline{\Omega}.

Lemma 2.1 (Metric properties of optical length).

Let (𝐗,d)({\mathbf{X}},d) be a complete length space and Ω𝐗\Omega\subsetneq{\mathbf{X}} be a bounded domain. Assume that ff satisfies (1.3). Let LfL_{f} be the optical length function defined by (1.8). Assume in addition that (1.9) holds. Then LfL_{f} satisfies the following:

  1. (1)

    Lf(x,y)0L_{f}(x,y)\geq 0 for all x,yΩ¯x,y\in\overline{\Omega}, and Lf(x,y)=0L_{f}(x,y)=0 holds if and only if x=yx=y;

  2. (2)

    Lf(x,y)=Lf(y,x)L_{f}(x,y)=L_{f}(y,x) for all x,yΩ¯x,y\in\overline{\Omega};

  3. (3)

    Lf(x,y)Lf(x,z)+Lf(y,z)L_{f}(x,y)\leq L_{f}(x,z)+L_{f}(y,z) for all x,y,zΩ¯x,y,z\in\overline{\Omega}.

Moreover,

Lf(x,y)αd(x,y)for all x,yΩ.L_{f}(x,y)\geq\alpha d(x,y)\quad\text{for all $x,y\in\Omega$.} (2.1)
Proof.

By (1.3), it is clear that Lf0L_{f}\geq 0 in Ω¯×Ω¯\overline{\Omega}\times\overline{\Omega} and in addition, Lf(x,y)=0L_{f}(x,y)=0 if and only if x=yx=y. Moreover, the definition (1.8) immediately implies (2.1). This completes the proof of (1). The statement (2) is obvious. We finally prove (3), which is also quite straightforward. For any ε>0\varepsilon>0, there exist γ1Γf(x,z)\gamma_{1}\in\Gamma_{f}(x,z) and γ2Γf(z,y)\gamma_{2}\in\Gamma_{f}(z,y) such that

Lf(x,z)If(γ1)ε,Lf(z,y)If(γ2)ε.L_{f}(x,z)\geq I_{f}(\gamma_{1})-\varepsilon,\quad L_{f}(z,y)\geq I_{f}(\gamma_{2})-\varepsilon.

By connecting the curve γ1\gamma_{1} and γ2\gamma_{2} to build a curve joining xx and yy, we can easily see that

Lf(x,y)If(γ1)+If(γ2)Lf(x,z)+Lf(z,y)+2ε.L_{f}(x,y)\leq I_{f}(\gamma_{1})+I_{f}(\gamma_{2})\leq L_{f}(x,z)+L_{f}(z,y)+2\varepsilon.

We conclude the proof by sending ε0\varepsilon\to 0. ∎

Remark 2.2.

Under the assumption (1.9), using the metric LfL_{f}, we can introduce a new notion of length of curves in Ω¯\overline{\Omega}. For any curve γ:[a,b]Ω¯\gamma:[a,b]\to\overline{\Omega}, define

f(γ)=supa=t0<t1<<tk=bj=0k1Lf(γ(tj),γ(tj+1)).\ell_{f}(\gamma)=\sup_{a=t_{0}<t_{1}<\cdots<t_{k}=b}\sum_{j=0}^{k-1}L_{f}(\gamma(t_{j}),\gamma(t_{j+1})).

We note that when f=1f=1 in Ω¯\overline{\Omega}, LfL_{f} defines the intrinsic metric and Ω¯\overline{\Omega} equipped with this metric is a length space. For a general function ff one can still verify that Ω¯\overline{\Omega} is a length space with metric LfL_{f} and the length of curves in this metric is given by f\ell_{f} defined above.

As an immediate consequence of (2.1), the closure/boundary of Ω\Omega with respect to LfL_{f} is contained in the closure/boundary with respect to dd. We can obtain the bi-Lipschitz equivalence between dd and LfL_{f} if a reverse version of (2.1) holds. It is the case when ff is uniformly bounded. However we cannot expect the bi-Lipschitz equivalence when ff is unbounded. A simple example is as follows.

Example 2.3.

Let 𝐗={\mathbf{X}}={\mathbb{R}} with dd the Euclidean metric and μ\mu the one-dimensional Lebesgue measure. Let Ω=(1,1)\Omega=(-1,1) and ff be the function given by (1.5). In this case, we have

Lf(x,0)=0xf(s)𝑑s=2|x|L_{f}(x,0)=\int_{0}^{x}f(s)\,ds=2\sqrt{|x|}

for all x(1,1)x\in(-1,1) and therefore

Lf(x,0)d(x,0)=2|x|as x0.\frac{L_{f}(x,0)}{d(x,0)}=\frac{2}{\sqrt{|x|}}\to\infty\quad\text{as $x\to 0$.}

Note however that (1,1)(-1,1) is bounded with respect to the metric LfL_{f}, and the topology generated by LfL_{f} agrees with the Euclidean topology on (1,1)(-1,1); moreover, [1,1][-1,1] is complete and compact in both metrics.

One can further consider the case when LfL_{f} is unbounded in Ω¯\overline{\Omega}; in fact, in Example 2.3, replacing ff with the following function

f(x)=1|x|,x(1,1){0},f(x)=\frac{1}{|x|},\quad x\in(-1,1)\setminus\{0\},

we see that Lf(0,x)=L_{f}(0,x)=\infty for any x(1,1){0}x\in(-1,1)\setminus\{0\}. Observe that this choice of ff does not belong to Lp((1,1))L^{p}((-1,1)) for any p1p\geq 1. Since such situation is not our focus in this work, we do not pursue this direction here and leave it for future discussions.

Even if LfL_{f} is bounded in Ω¯×Ω¯\overline{\Omega}\times\overline{\Omega}, due to the unboundedness of ff in general it may happen that Lf(x,y)↛0L_{f}(x,y)\not\to 0 as d(x,y)0d(x,y)\to 0. An example on metric graphs is as follows.

Example 2.4.

In 2{\mathbb{R}}^{2}, let e0e_{0} be the closed line segment between points (0,0)(0,0) and (1,0)(1,0), and eje_{j} be the line segments connecting Pj=(1/j,0)P_{j}=(1/j,0) and Qj=(1/j,1/j)Q_{j}=(1/j,1/j) for j=1,2,j=1,2,\ldots We take 𝐗=Ω¯=j=0,1,ej{\mathbf{X}}=\overline{\Omega}=\bigcup_{j=0,1,\ldots}e_{j} with dd being the intrinsic metric of this graph. See Figure 1.

Refer to caption
Figure 1. Example 2.4

It is not difficult to see that (𝐗,d)({\mathbf{X}},d) is a complete geodesics space. For x𝐗x\in{\mathbf{X}} with coordinates (x1,x2)(x_{1},x_{2}) in 2{\mathbb{R}}^{2}, let

f(x1,x2)={1/x1if x2>0;1if x2=0.f(x_{1},x_{2})=\begin{dcases}1/x_{1}&\text{if $x_{2}>0$;}\\ 1&\text{if $x_{2}=0$.}\end{dcases}

Denoting O=(0,0)O=(0,0), by direct calculations, for any j2j\geq 2 we have

Lf(O,Qj)=j01j𝑑s+01j𝑑s=1+1j.L_{f}(O,Q_{j})=j\int_{0}^{1\over j}\,ds+\int_{0}^{1\over j}ds=1+{1\over j}.

(Here we choose the optimal integration path from QjQ_{j} to PjP_{j} and then to OO.) We thus observe that as jj\to\infty, d(O,Qj)0d(O,Q_{j})\to 0 but Lf(O,Qj)1L_{f}(O,Q_{j})\to 1.

Lemma 2.5 (Completeness).

Let (𝐗,d)({\mathbf{X}},d) be a complete length space and Ω𝐗\Omega\subsetneq{\mathbf{X}} be an open set. Assume that ff satisfies (1.3). Let LfL_{f} be the optical length function defined by (1.8). Then (Ω¯,Lf)(\overline{\Omega},L_{f}) is complete.

Proof.

Suppose that xjΩ¯x_{j}\in\overline{\Omega} is a Cauchy sequence with respect to the metric LfL_{f}. Thanks to (2.1) in Lemma 2.1, we see that {xj}\{x_{j}\} is also a Cauchy sequence with respect to the metric dd. Since Ω¯\overline{\Omega} is a closed set in the complete space 𝐗{\mathbf{X}}, there exists x0Ω¯x_{0}\in\overline{\Omega} such that d(xj,x0)0d(x_{j},x_{0})\to 0 as jj\to\infty.
By definition of Cauchy sequences, for any ε>0\varepsilon>0, we can take xj1x_{j_{1}} such that Lf(xj1,xi)<ε/2L_{f}(x_{j_{1}},x_{i})<\varepsilon/2 for all ij1i\geq j_{1}. Similarly, we can choose xj2x_{j_{2}} with j2>j1j_{2}>j_{1} satisfying Lf(xj2,xi)<ε/4L_{f}(x_{j_{2}},x_{i})<\varepsilon/4 for all ij2i\geq j_{2}. We repeat this process to obtain a sequence xjkx_{j_{k}} such that Lf(xjk,xjk+1)<2kεL_{f}(x_{j_{k}},x_{j_{k+1}})<2^{-k}\varepsilon. By definition of LfL_{f}, we can find curves γk\gamma_{k} in Ω¯\overline{\Omega} joining xjkx_{j_{k}} and xjk+1x_{j_{k}+1} such that

γkf𝑑s<2kε.\int_{\gamma_{k}}f\,ds<2^{-k}\varepsilon.

Concatenating these curves in order, we build a curve γ\gamma connecting xj1x_{j_{1}} to x0x_{0} satisfying

γf𝑑sk12kε=ε,\int_{\gamma}f\,ds\leq\sum_{k\geq 1}2^{-k}\varepsilon=\varepsilon,

which yields Lf(xj1,x0)εL_{f}(x_{j_{1}},x_{0})\leq\varepsilon. In view of the arbitrariness of ε>0\varepsilon>0, we have actually proved the convergence of xjx_{j} to x0x_{0} in the metric LfL_{f}. ∎

We cannot guarantee the compactness of (Ω¯,Lf)(\overline{\Omega},L_{f}) without imposing further assumptions. Indeed, in Example 2.4 we see that the sequence {Qj}j2\{Q_{j}\}_{j\geq 2} is bounded but without a convergent subsequence in (Ω¯,Lf)(\overline{\Omega},L_{f}); indeed, for each jkj\neq k we have that Lf(Qj,Qk)2L_{f}(Q_{j},Q_{k})\geq 2. It follows that (Ω¯,Lf)(\overline{\Omega},L_{f}) is not compact.

3. Monge Solutions and Comparison Principle

3.1. Definition of Monge solutions

Let us now study the Dirichlet problem (1.1), (1.2). We begin with the definition of Monge solutions to (1.1).

Definition 3.1 (Monge solutions).

We say that a locally bounded function u:Ωu:\Omega\to{\mathbb{R}} is a Monge solution (resp. subsolution, supersolution) to (1.1) in Ω\Omega if for every x0Ωx_{0}\in\Omega,

lim supLf(x,x0)0u(x0)u(x)Lf(x0,x)=1(resp., ,).\limsup_{L_{f}(x,x_{0})\to 0}\frac{u(x_{0})-u(x)}{L_{f}(x_{0},x)}=1\quad(\text{resp., }\leq,\geq). (3.1)

We stress that the topology in the local boundedness of uu and limsup in the definition above is taken with respect to the metric LfL_{f}. Since Lf(x,x0)0L_{f}(x,x_{0})\to 0 implies d(x,x0)0d(x,x_{0})\to 0 by (2.1), we see that

lim supLf(x,x0)0u(x0)u(x)Lf(x0,x)=1for every x0Ω\limsup_{L_{f}(x,x_{0})\to 0}\frac{u(x_{0})-u(x)}{L_{f}(x_{0},x)}=1\quad\text{for every $x_{0}\in\Omega$} (3.2)

implies

lim supd(x,x0)0u(x0)u(x)Lf(x0,x)1for every x0Ω.\limsup_{d(x,x_{0})\to 0}\frac{u(x_{0})-u(x)}{L_{f}(x_{0},x)}\geq 1\quad\text{for every $x_{0}\in\Omega$}. (3.3)

In general, we cannot expect that the reverse inequality of (3.3) holds without further assumptions like (1.14).

We also emphasize that our definition does not require any continuity or semicontinuity of uu. This relaxation constitutes a difference from the standard definition of discontinuous viscosity solutions. Recall that when defining a (possibly discontinuous) subsolution uu, we usually assume uUSC(Ω)u\in USC(\Omega) (and uLSC(Ω)u\in LSC(\Omega) for the symmetric supersolution definition). Here, USC(Ω)USC(\Omega) and LSC(Ω)LSC(\Omega) denote the classes of upper and lower semicontinuous functions in Ω\Omega respectively with respect to dd. We do not adopt such restrictions here.

Remark 3.2.

Our definition above is a direct generalization of the standard notion of Monge solutions for fC(Ω)f\in C(\Omega). Recall that 𝐗{\mathbf{X}} is assumed to be a length space. When ff is continuous at x0x_{0}, we can show that (1.12) holds and that Lf(x,x0)0L_{f}(x,x_{0})\to 0 as d(x,x0)0d(x,x_{0})\to 0. Our generalized notion is consistent with the case for continuous ff discussed in [43].

Using the subslope defined in (1.10), we can see in a straightforward manner that our Monge solutions (resp., subsolutions, supersolutions) of (1.1) can simply be understood as Monge solutions (resp., subsolutions, supersolutions) of

|fu|=1in Ω.|\nabla_{f}u|=1\quad\text{in $\Omega$}. (3.4)
Proposition 3.3.

Let (𝐗,d)({\mathbf{X}},d) be a complete length space and Ω𝐗\Omega\subsetneq{\mathbf{X}} be a bounded domain. Assume that (1.3) hold. If uu is a Monge subsolution of (1.1), then for any x0Ωx_{0}\in\Omega and any C>1C>1,

u(x)u(x0)CLf(x,x0)u(x)-u(x_{0})\geq-CL_{f}(x,x_{0})

for any xΩx\in\Omega sufficiently close to x0x_{0}. In addition, if (1.14) holds, then uu is lower semicontinuous in Ω\Omega.

On the other hand, Definition 3.1, together with (1.14), does not guarantee the upper semicontinuity of a Monge solution. However, for Monge solutions that arise from the associated Dirichlet problem, we will later use its optimal control interpretation to show the continuity of solutions.

Thanks to the metric change, one can essentially apply the results in [43] to (3.4) in the complete metric space (Ω¯,Lf)(\overline{\Omega},L_{f}) to show both uniqueness and existence of Monge solutions of (1.1)(1.2) under appropriate assumptions on the boundary data gg. The compatibility condition (1.13) we will impose on gg, combined with (1.14), basically ensures that the topology stays equivalent when converting the metric from dd to LfL_{f}.

3.2. Comparison principle

Let us prove a comparison principle, where we assume the semicontinuity of Monge sub- and supersolutions with respect to the metric LfL_{f}. We say that u:Ω¯u:\overline{\Omega}\to{\mathbb{R}} is upper (resp., lower) semicontinuous with respect to LfL_{f}, denoted by uUSCL(Ω¯)u\in USC_{L}(\overline{\Omega}) (resp., uLSCL(Ω¯)u\in LSC_{L}(\overline{\Omega})), if for every fixed xΩ¯x\in\overline{\Omega}, we have

lim supLf(x,y)0,yΩ¯u(y)u(x)(resp.,lim infLf(x,y)0,yΩ¯u(y)u(x)).\limsup_{L_{f}(x,y)\to 0,\ y\in\overline{\Omega}}\ u(y)\leq u(x)\quad\left(\text{resp.,}\liminf_{L_{f}(x,y)\to 0,\ y\in\overline{\Omega}}\ u(y)\geq u(x)\right).

It is not difficult to see that uUSCL(Ω¯)u\in USC_{L}(\overline{\Omega}) (resp., uLSCL(Ω¯)u\in LSC_{L}(\overline{\Omega})) if uUSC(Ω¯)u\in USC(\overline{\Omega}) (resp., uLSC(Ω¯)u\in LSC(\overline{\Omega})) with respect to dd, since Lf(x,y)0L_{f}(x,y)\to 0 implies that d(x,y)0d(x,y)\to 0.

Theorem 3.4 (Comparison principle for Monge solutions).

Let (𝐗,d)({\mathbf{X}},d) be a complete length space and Ω𝐗\Omega\subsetneq{\mathbf{X}} be a bounded domain. Assume that ff satisfies (1.3). Let uUSCL(Ω¯)u\in USC_{L}(\overline{\Omega}) and vLSCL(Ω¯)v\in LSC_{L}(\overline{\Omega}) be respectively a bounded Monge subsolution and a bounded Monge supersolution of (1.1). If

limδ0sup{u(x)v(x):xΩ¯,d(x,Ω)δ}0,\lim_{\delta\to 0}\sup\left\{u(x)-v(x):x\in\overline{\Omega},\ d(x,\partial\Omega)\leq\delta\right\}\leq 0, (3.5)

then uvu\leq v holds in Ω¯\overline{\Omega}.

Proof.

Since uu and vv are bounded, we may assume that u,v0u,v\geq 0 by adding a positive constant to them. It suffices to show that λuv\lambda u\leq v in Ω\Omega for all λ(0,1)\lambda\in(0,1). Assume by contradiction that there exists λ(0,1)\lambda\in(0,1) such that supΩ(λuv)>2τ\sup_{\Omega}(\lambda u-v)>2\tau for some τ>0\tau>0. By (3.5), we may take δ>0\delta>0 small such that

λu(x)v(x)u(x)v(x)τ\lambda u(x)-v(x)\leq u(x)-v(x)\leq\tau

for all xΩ¯Ωδx\in\overline{\Omega}\setminus\Omega_{\delta}, where we denote Ωr={xΩ:d(x,Ω)>r}\Omega_{r}=\{x\in\Omega:d(x,\partial\Omega)>r\} for r>0r>0. We choose ε(0,δα/2)\varepsilon\in(0,\delta\alpha/2) small, where α>0\alpha>0 is the lower bound of ff as in (1.3), such that

supΩ(λuv)>2τ+ε2\sup_{\Omega}(\lambda u-v)>2\tau+\varepsilon^{2}

and

ε<1λ.\varepsilon<1-\lambda. (3.6)

Thus there exists x0Ωx_{0}\in\Omega such that λu(x0)v(x0)supΩ(λuv)ε2>2τ\lambda u(x_{0})-v(x_{0})\geq\sup_{\Omega}(\lambda u-v)-\varepsilon^{2}>2\tau and therefore x0Ωδx_{0}\in\Omega_{\delta}.

In view of Lemma 2.5, we see that (Ω¯,Lf)(\overline{\Omega},L_{f}) is complete. Since λuv\lambda u-v are bounded from above and upper semicontinuous in Ω¯\overline{\Omega} with respect to the metric LfL_{f}, by Ekeland’s variational principle (cf. [23, Theorem 1.1], [24, Theorem 1]), there exists xεΩx_{\varepsilon}\in\Omega such that

Lf(xε,x0)<ε,L_{f}(x_{\varepsilon},x_{0})<\varepsilon, (3.7)
λu(xε)v(xε)λu(x0)v(x0),\lambda u(x_{\varepsilon})-v(x_{\varepsilon})\geq\lambda u(x_{0})-v(x_{0}),

and

λu(x)v(x)εLf(xε,x)λu(xε)v(xε)for all xΩ.\lambda u(x)-v(x)-\varepsilon L_{f}(x_{\varepsilon},x)\leq\lambda u(x_{\varepsilon})-v(x_{\varepsilon})\quad\text{for all $x\in\Omega$.} (3.8)

Note that by (2.1), the relation (3.7) combined with the choice of ε<δα/2\varepsilon<\delta\alpha/2 implies that xεBε/α(x0)Ωδ/2x_{\varepsilon}\in B_{\varepsilon/\alpha}(x_{0})\subset\Omega_{\delta/2}. Then from (3.8), it follows that

v(xε)v(x)λu(xε)λu(x)+εLf(xε,x)for all xΩv(x_{\varepsilon})-v(x)\leq\lambda u(x_{\varepsilon})-\lambda u(x)+\varepsilon L_{f}(x_{\varepsilon},x)\quad\text{for all $x\in\Omega$} (3.9)

when ε>0\varepsilon>0 is small enough. Hence, by (2.1) we get

lim supLf(x,xε)0v(xε)v(x)Lf(xε,x)lim supLf(x,xε)0λ(u(xε)u(x))Lf(xε,x)+ε.\limsup_{L_{f}(x,x_{\varepsilon})\to 0}\frac{v(x_{\varepsilon})-v(x)}{L_{f}(x_{\varepsilon},x)}\leq\limsup_{L_{f}(x,x_{\varepsilon})\to 0}\frac{\lambda(u(x_{\varepsilon})-u(x))}{L_{f}(x_{\varepsilon},x)}+\varepsilon.

By the Monge subsolution property of uu and the Monge supersolution property of vv, it follows that 1λ+ε1\leq\lambda+\varepsilon, which is a contradiction to (3.6). Hence, we obtain λuv\lambda u\leq v in Ω\Omega for all 0<λ<10<\lambda<1. Letting λ1\lambda\to 1, we end up with uvu\leq v in Ω\Omega. Our proof is thus complete. ∎

Remark 3.5.

If in Theorem 3.4 we further assume that (Ω¯,Lf)(\overline{\Omega},L_{f}) is compact, then the proof becomes simpler. Under this assumption, there is no need to use Ekeland’s variational principle, since uvu-v is upper semicontinuous and hence attains a maximum at a point x^\hat{x} in the compact set Ω¯\overline{\Omega}. The condition (3.5) can also be simplified; it is sufficient to assume that uvu\leq v on Ω\partial\Omega, since it guarantees a positive maximum of λuv\lambda u-v in Ω\Omega for 0<λ<10<\lambda<1 in our proof above. We summarize this observation in the following theorem.

Theorem 3.6 (Comparison principle in a compact space).

Let (𝐗,d)({\mathbf{X}},d) be a complete length space and Ω𝐗\Omega\subsetneq{\mathbf{X}} be a bounded domain. Assume that ff satisfies (1.3) and that (Ω¯,Lf)(\overline{\Omega},L_{f}) is compact. Let uUSCL(Ω¯)u\in USC_{L}(\overline{\Omega}) and vLSCL(Ω¯)v\in LSC_{L}(\overline{\Omega}) be respectively a bounded Monge subsolution and a bounded Monge supersolution of (1.1). If uvu\leq v on Ω\partial\Omega, then uvu\leq v in Ω¯\overline{\Omega}.

We would also like to remark that the comparison results may be applied to the situation with only partial boundary data (or the so-called boundary condition in the viscosity sense). In general, even though gg is prescribed on Ω\partial\Omega, a Monge solution may only fulfill the condition on a subset of Ω\partial\Omega. We discuss such situation in detail in Section 4.2.

We concluding this section by remarking that the boundedness condition on uu and vv in Theorem 3.4 is essential. In fact, without assuming the boundedness, we have the following simple counterexample for uniqueness of Monge solutions.

Example 3.7.

Let 𝐗{\mathbf{X}} be the unit circle in 2{\mathbb{R}}^{2}, centered at O=(0,0)O=(0,0), that is,

𝐗=𝕊1={(cosθ,sinθ):π<θπ}.{\mathbf{X}}=\mathbb{S}^{1}=\{(\cos\theta,\sin\theta):-\pi<\theta\leq\pi\}.

Here (𝐗,d)({\mathbf{X}},d) is a complete geodesic space with dd its intrinsic length metric. Let Ω=𝐗{(1,0)}\Omega={\mathbf{X}}\setminus\{(1,0)\} and g(1,0)=0g(1,0)=0. Take f:Ωf:\Omega\to{\mathbb{R}} to be

f(cosθ,sinθ)={1/πif 0θπ,1/(θ+π)if π<θ<0.f(\cos\theta,\sin\theta)=\begin{cases}1/\pi&\text{if $0\leq\theta\leq\pi$,}\\ 1/(\theta+\pi)&\text{if $-\pi<\theta<0$}.\end{cases}

See Figure 2.

Refer to caption
Figure 2. Example 3.7

Note that

u(cosθ,sinθ)={θ/πif 0θπ,log(π/π+θ)if π<θ<0u(\cos\theta,\sin\theta)=\begin{cases}\theta/\pi&\text{if $0\leq\theta\leq\pi$,}\\ \log(\pi/\pi+\theta)&\text{if $-\pi<\theta<0$}\end{cases}

is a Monge solution of (1.1), (1.2), which can be derived from the formula (1.16) in Theorem 1.1. Note that as (π,0)θπ(-\pi,0)\ni\theta\to-\pi, we must have Lf((cosθ,sinθ),(1,0))L_{f}((\cos\theta,\sin\theta),(-1,0))\to\infty. On the other hand, we observe that

v(cosθ,sinθ)={θ/πif 0θπ,log(π/π+θ)if π<θ<0v(\cos\theta,\sin\theta)=\begin{cases}\theta/\pi&\text{if $0\leq\theta\leq\pi$,}\\ -\log(\pi/\pi+\theta)&\text{if $-\pi<\theta<0$}\end{cases}

is also a Monge solution that also satisfies (1.17). This example shows that in general we may have multiple solutions, continuous but unbounded in (Ω¯,Lf)(\overline{\Omega},L_{f}). In this example, once we equip 𝐗=Ω¯{\mathbf{X}}=\overline{\Omega} with the metric LfL_{f}, the topology on 𝐗{\mathbf{X}} changes so that Ω\Omega now is homeomorphic to the Euclidean set (,0)(0,1](-\infty,0)\cup(0,1], with boundary {0}\{0\}. In effect, Ω\Omega is no longer connected with respect to the topology generated by the metric LfL_{f}, allowing us to construct two distinct solutions.

4. Existence of Monge solutions

In this section, we establish an optimal control interpretation of the Dirichlet boundary condition (1.2) for the problem (1.1) and prove our main result, Theorem 1.1. We also discuss the case when the boundary compatibility condition (1.13) fails to hold and turn it into another Dirichlet problem of the same type but with reduced boundary data.

4.1. Optimal control formulation for the Dirichlet problem

Recall that the standing assumption (1.9) still holds. In other words, there exist curves in Ω¯\overline{\Omega} connecting any x,yΩ¯x,y\in\overline{\Omega} such that Lf(x,y)<L_{f}(x,y)<\infty. It is a condition on Ω\Omega as well as the regularity of ff.

Lemma 4.1 (Lipschitz continuity in LfL_{f}).

Let (𝐗,d)({\mathbf{X}},d) be a complete length space, Ω𝐗\Omega\subsetneq{\mathbf{X}} be a bounded domain, ff satisfy (1.3), and g:Ωg:\partial\Omega\to{\mathbb{R}} be bounded. Then the function uu, as defined by (1.16), satisfies

u(x)u(y)+Lf(x,y)for all x,yΩ¯.u(x)\leq u(y)+L_{f}(x,y)\quad\text{for all $x,y\in\overline{\Omega}$}. (4.1)

In particular, uu is Lipschitz in Ω¯\overline{\Omega} with respect to LfL_{f} and hence locally bounded with respect to the topology generated by LfL_{f}. Moreover, if (1.14) holds, then uu is uniformly continuous in Ω¯\overline{\Omega} with respect to dd.

Proof.

By Lemma 2.1(3), we have for any x,yΩ¯x,y\in\overline{\Omega} and zΩz\in\partial\Omega,

u(x)Lf(x,z)+g(z)Lf(y,z)+g(z)+Lf(x,y).u(x)\leq L_{f}(x,z)+g(z)\leq L_{f}(y,z)+g(z)+L_{f}(x,y).

Then taking infima over zΩz\in\partial\Omega, we get (4.1). As an immediate consequence, the Lipschitz continuity of uu in Ω¯\overline{\Omega} with respect to LfL_{f} holds. Also, we can obtain the uniform continuity of uu with respect to dd if we additionally assume that (1.14) holds. ∎

Let us now prove our main result, Theorem 1.1.

Proof of Theorem 1.1.

We first show (i). By Lemma 4.1, we see that uu is Lipschitz with respect to LfL_{f} and is a Monge subsolution; indeed, by (4.1) we have that for any x0Ωx_{0}\in\Omega,

u(x0)u(x)Lf(x0,x)for all xΩu(x_{0})-u(x)\leq L_{f}(x_{0},x)\quad\text{for all $x\in\Omega$} (4.2)

and thus

lim supLf(x,x0)0u(x0)u(x)Lf(x0,x)1.\limsup_{L_{f}(x,x_{0})\to 0}\frac{u(x_{0})-u(x)}{L_{f}(x_{0},x)}\leq 1. (4.3)

On the other hand, in view of the definition of uu from (1.16), for any ε>0\varepsilon>0 and any r(0,d(x0,Ω))r\in(0,d(x_{0},\partial\Omega)), there exist yΩy\in\partial\Omega and γΓf(x0,y)\gamma\in\Gamma_{f}(x_{0},y) such that

u(x0)γf𝑑s+g(y)εr.u(x_{0})\geq\int_{\gamma}f\,ds+g(y)-\varepsilon r. (4.4)

Let xx be a point on γ\gamma with d(x0,x)=rd(x_{0},x)=r and denote by γx\gamma_{x} the portion of γ\gamma between x0x_{0} and xx. Then by (1.16) again, we deduce that

u(x0)γxf𝑑s+u(x)εd(x0,x)u(x)+Lf(x0,x)εd(x0,x).u(x_{0})\geq\int_{\gamma_{x}}f\,ds+u(x)-\varepsilon d(x_{0},x)\geq u(x)+L_{f}(x_{0},x)-\varepsilon d(x_{0},x).

It follows that

u(x0)u(x)Lf(x0,x)1εd(x0,x)Lf(x0,x),\frac{u(x_{0})-u(x)}{L_{f}(x_{0},x)}\geq 1-\varepsilon\frac{d(x_{0},x)}{L_{f}(x_{0},x)},

which, by (1.3), yields

lim supLf(x,x0)0u(x0)u(x)Lf(x0,x)1εα.\limsup_{L_{f}(x,x_{0})\to 0}\frac{u(x_{0})-u(x)}{L_{f}(x_{0},x)}\geq 1-\frac{\varepsilon}{\alpha}.

Due to the arbitrariness of ε>0\varepsilon>0 and x0Ωx_{0}\in\Omega, we obtain the Monge supersolution property of uu. We have thus shown that uu is a Monge solution of (1.1) in the sense of Definition 3.1.

Let us show (ii) and (iii). We prove (1.17) under the additional compatibility condition (1.13) on the boundary value gg. Indeed, fixing xΩ¯x\in\overline{\Omega} and yΩy\in\partial\Omega arbitrarily, by the definition of uu in (1.16) we can see that

u(x)g(y)Lf(x,y)u(x)-g(y)\leq L_{f}(x,y) (4.5)

For any δ>0\delta>0 small, the definition of uu also implies the existence of yδΩy_{\delta}\in\partial\Omega such that

u(x)Lf(x,yδ)+g(yδ)max{δ,d(x,Ω)}.u(x)\geq L_{f}(x,y_{\delta})+g(y_{\delta})-\max\{\delta,d(x,\partial\Omega)\}. (4.6)

Since the compatibility condition (1.13) yields

g(y)g(yδ)+Lf(y,yδ),g(y)\leq g(y_{\delta})+L_{f}(y,y_{\delta}),

applying Lemma 2.1 we deduce from (4.6) that

u(x)g(y)Lf(x,yδ)Lf(y,yδ)d(x,Ω)Lf(x,y)max{δ,d(x,Ω)}.u(x)-g(y)\geq L_{f}(x,y_{\delta})-L_{f}(y,y_{\delta})-d(x,\partial\Omega)\geq-L_{f}(x,y)-\max\{\delta,d(x,\partial\Omega)\}.

Combining (4.5) with the above inequality we obtain

|u(x)g(y)|Lf(x,y)+max{δ,d(x,Ω)}|u(x)-g(y)|\leq L_{f}(x,y)+\max\{\delta,d(x,\partial\Omega)\} (4.7)

for all xΩx\in\Omega and yΩy\in\partial\Omega. Therefore,

sup{|u(x)g(y)|:\displaystyle\sup\{|u(x)-g(y)|: xΩ¯,yΩ,d(x,y)δ}\displaystyle x\in\overline{\Omega},\ y\in\partial\Omega,\ d(x,y)\leq\delta\} (4.8)
sup{Lf(x,y):x,yΩ¯Ωδ,d(x,y)δ}+δ.\displaystyle\leq\sup\{L_{f}(x,y):x,y\in\overline{\Omega}\setminus\Omega_{\delta},\ d(x,y)\leq\delta\}+\delta.

By the additional assumption (1.15), we obtain (1.17) by passing to the limit of (4.8) as δ0\delta\to 0. In particular, we have u=gu=g on Ω\partial\Omega. The uniqueness of solutions follow from the comparison principle, Theorem 3.4. Suppose that there is another such solution vv satisfying (1.17). For any xΩ¯x\in\overline{\Omega} with d(x,Ω)δd(x,\partial\Omega)\leq\delta, we can find yΩy\in\partial\Omega such that d(x,y)2δd(x,y)\leq 2\delta and

sup{u(x)v(x):xΩ¯,d(x,Ω)δ}\displaystyle\sup\left\{u(x)-v(x):x\in\overline{\Omega},\ d(x,\partial\Omega)\leq\delta\right\}
sup{|u(x)g(y)|+|v(x)g(y)|:xΩ¯,yΩ,d(x,y)2δ}.\displaystyle\leq\sup\{|u(x)-g(y)|+|v(x)-g(y)|:x\in\overline{\Omega},y\in\partial\Omega,d(x,y)\leq 2\delta\}.

By (1.17), this yields (3.5). Using Theorem 3.4, we end up with u=vu=v in Ω¯\overline{\Omega}, which completes the proof of (ii).

The statement (iii) can be immediately proved, since (1.14) implies (1.15). We obtain the uniform continuity of uu with respect to dd by (1.14) and Lemma  4.1. ∎

Remark 4.2.

Following Theorem 3.6, we see that if (𝐗,Lf)({\mathbf{X}},L_{f}) is complete and locally compact, then the function uu obtained by (1.16) is the unique Monge solution of (1.1)(1.2) provided that (1.13) holds. We do not need to adopt interpretation (1.17) for (1.2).

Following Example 2.3, we now give an explicit example of unique Monge solution of the Dirichlet problem in one dimension with LpL^{p} inhomogeneous term.

Example 4.3.

Let (𝐗,d)({\mathbf{X}},d), Ω\Omega and ff be given as in Example 2.3. We recall that fLp(Ω)f\in L^{p}(\Omega) with p[1,2)p\in[1,2). Setting g(±1)=0g(\pm 1)=0, by Theorem 1.1(iii) we can show that the unique Monge solution of the Dirichlet problem is as in (1.6).

The statement (i) in Theorem 1.1 is a very general existence result for Monge solutions. It allows us to have a discontinuous (with respect to dd) Monge solution of (1.1) even when (1.14) fails to hold. In this case, under the assumptions of (ii), we have unique existence of a solution that is Lipschitz with respect to LfL_{f} but possibly discontinuous in the metric dd. Below we give a typical example illustrating such behavior based on Example 2.4.

Example 4.4.

Let (𝐗,d)({\mathbf{X}},d) and ff be given as in Example 2.4. Let Ω=𝐗{Q1}\Omega={\mathbf{X}}\setminus\{Q_{1}\}; in other words, we set Ω={Q1}\partial\Omega=\{Q_{1}\}. Also, we take g=0g=0 at Q1Q_{1}. We have shown in Example 2.4 that 𝐗=Ω¯{\mathbf{X}}=\overline{\Omega} is complete with respect to the metric LfL_{f}. Applying Theorem 1.1(ii), we obtain a unique Monge solution uu of (1.1)(1.2) satisfying (1.17). In particular, by (1.16) we have u(O)=2u(O)=2, u(Qj)=31/ju(Q_{j})=3-1/j for all j2j\geq 2. This clearly shows that uu is not continuous at OO with respect to the metric dd. However, uu is Lipschitz continuous with respect to the metric LfL_{f}.

Let us present an additional property of the Monge solution uu, which is a direct generalization of [43, Proposition 4.15].

Proposition 4.5 (Additional regularity).

Let (𝐗,d)({\mathbf{X}},d) be a complete length space and Ω𝐗\Omega\subsetneq{\mathbf{X}} be a bounded domain. Assume that ff satisfies (1.3), (1.9) and that gg satisfies (1.13) and (1.15) holds. Then any uu that satisfies (1.17) is the unique Monge solution of (1.1) if and only if

|fu|=1and|f+u||fu| in Ω.|\nabla_{f}u|=1\quad\text{and}\quad|\nabla_{f}^{+}u|\leq|\nabla_{f}^{-}u|\quad\text{ in $\Omega$.} (4.9)

Here, for any locally bounded function u:Ωu:\Omega\to{\mathbb{R}} and xΩx\in\Omega, we set

|fu|(x)=lim supLf(x,y)0|u(y)u(x)|Lf(x,y),\displaystyle|\nabla_{f}u|(x)=\limsup_{L_{f}(x,y)\to 0}\frac{|u(y)-u(x)|}{L_{f}(x,y)}, (4.10)
|f+u|(x)=lim supLf(x,y)0max{u(y)u(x),0}Lf(x,y).\displaystyle|\nabla_{f}^{+}u|(x)=\limsup_{L_{f}(x,y)\to 0}\frac{\max\{u(y)-u(x),0\}}{L_{f}(x,y)}.
Proof.

The implication \Leftarrow is obvious, as (4.9) also implies |fu|=1|\nabla_{f}^{-}u|=1 in Ω\Omega. It suffices to prove \Rightarrow. The unique Monge solution uu can be expressed by (1.16). By Lemma 4.1, it satisfies (4.2). It follows that for every fixed x0Ωx_{0}\in\Omega, we have u(x)u(x0)+Lf(x0,x)u(x)\leq u(x_{0})+L_{f}(x_{0},x) for for xΩx\in\Omega, which immediately yields |f+u|(x0)1|\nabla_{f}^{+}u|(x_{0})\leq 1. Since

|fu|=max{|f+u|,|fu|},|\nabla_{f}u|=\max\left\{|\nabla_{f}^{+}u|,\ |\nabla_{f}^{-}u|\right\},

we immediately get (4.9). ∎

4.2. Loss of boundary data and reduced Dirichlet problem

In general, the condition (1.13) may not hold. As a result, it is possible that u=gu=g holds only on a subset of Ω\partial\Omega. This set may also depend on ff but we write Σg\Sigma_{g} to emphasize that it is the part where the Dirichlet condition is maintained. The loss of boundary data occurs on ΩΣg\partial\Omega\setminus\Sigma_{g}. At points x0ΩΣgx_{0}\in\partial\Omega\setminus\Sigma_{g} the Monge condition (3.1) holds, and so the function uu constructed in (1.16) solves the problem in the new domain Ω¯Σg\overline{\Omega}\setminus\Sigma_{g}. We thus can use the comparison results above with Ω\partial\Omega replaced by Σg\Sigma_{g} to guarantee the uniqueness of solutions. The following typical, well-understood example in {\mathbb{R}} reveals such a situation.

Example 4.6.

Let Ω=(0,1)\Omega=(0,1)\subset{\mathbb{R}} and f1f\equiv 1 in Ω\Omega. Set g(0)=0g(0)=0 and g(1)=2g(1)=2. It turns out that there are no solutions that are continuous in [0,1][0,1] and satisfy both (1.1) and (1.2) simultaneously. However, u(x)=xu(x)=x is the unique solution that satisfies only the partial boundary condition g(0)=0g(0)=0. In fact, |u|(1)=1|\nabla^{-}u|(1)=1 holds and we can still use the comparison principle if we consider Ω=(0,1]\Omega=(0,1] and Ω={0}\partial\Omega=\{0\}. One can certainly take another solution u(x)=x+1u(x)=x+1 for [0,1][0,1], which is a Monge solution in [0,1)[0,1) and satisfies the boundary condition u(1)=2u(1)=2. However, this solution does not satisfy the Monge condition (3.1) at 0, and moreover, such solutions are not necessarily unique from the PDE viewpoint; u(x)=x+3u(x)=-x+3 is also a Monge solution in [0,1)[0,1) with u(1)=2u(1)=2. Indeed, it is the only Monge solution in [0,1)[0,1) with boundary data u(1)=2u(1)=2. We resolve the uniqueness issue by taking Σg={0}\Sigma_{g}=\{0\} so that u(x)=xu(x)=x is the only solution in Ω=(0,1]\Omega=(0,1] satisfying u(0)=0u(0)=0. The function u(x)=xu(x)=x is regarded as a natural choice of the unique solution here also due to the optimal control interpretation given by (1.16).

Let us provide a more general result for the case when (1.13) fails to hold and ugu\neq g on Ω\partial\Omega. As explained at the end of Section 3, we still expect that there exists a solution uu such that u=gu=g on a subset Σg\Sigma_{g} of Ω\partial\Omega. In fact, we take

Σg={xΩ:g(x)infyΩ{g(y)+Lf(x,y)}}.\Sigma_{g}=\left\{x\in\partial\Omega:g(x)\leq\inf_{y\in\partial\Omega}\{g(y)+L_{f}(x,y)\}\right\}. (4.11)

By (1.16), it is clear that

Σg={xΩ:g(x)u(x)}={xΩ:g(x)=u(x)}.\Sigma_{g}=\left\{x\in\partial\Omega:g(x)\leq u(x)\right\}=\left\{x\in\partial\Omega:g(x)=u(x)\right\}. (4.12)

In general, the set Σg\Sigma_{g} may be empty. However, Σg\Sigma_{g}\neq\emptyset if Ω\partial\Omega is compact with respect to LfL_{f} and and gLSCL(Ω)g\in LSC_{L}(\partial\Omega), or if Ω\partial\Omega is compact with respect to dd and gLSCd(Ω)g\in LSC_{d}(\partial\Omega). In either case, Σg\Sigma_{g} contains the minimizers of gg on Ω\partial\Omega. Also, Σg\Sigma_{g} is a closed set with respect to the metric LfL_{f}.

Using Σg\Sigma_{g} we can reduce the original Dirichlet problem to

|u|(x)=f(x)in Ω¯Σg|\nabla u|(x)=f(x)\quad\text{in $\overline{\Omega}\setminus\Sigma_{g}$} (4.13)

with boundary condition

u=gfor xΣg.u=g\quad\text{for $x\in\Sigma_{g}$.} (4.14)

It turns out that uu given by (1.16) is the unique Monge solution of (4.13),  (4.14) when Σg\Sigma_{g}\neq\emptyset. Here, for these new Monge solutions, we certainly need to extend the definition of Monge solutions to Ω¯Σg\overline{\Omega}\setminus\Sigma_{g}. More precisely, we say a locally bounded function u:Ω¯Σgu:\overline{\Omega}\setminus\Sigma_{g}\to{\mathbb{R}} is a Monge solution of (4.13) if it satisfies the property in Definition 3.1 with Ω¯Σg\overline{\Omega}\setminus\Sigma_{g} playing the role of the domain Ω\Omega there, and with (3.1) replaced by

lim supxΩ¯Σg,Lf(x,x0)0u(x0)u(x)Lf(x0,x)=1.\limsup_{x\in\overline{\Omega}\setminus\Sigma_{g},\ L_{f}(x,x_{0})\to 0}\frac{u(x_{0})-u(x)}{L_{f}(x_{0},x)}=1. (4.15)

One can define Monge subsolutions and supersolutions of (4.13) in a similar way.

Proposition 4.7 (Well-posedness with possible loss of boundary data).

Let (𝐗,d)({\mathbf{X}},d) be a complete length space and Ω𝐗\Omega\subsetneq{\mathbf{X}} be a bounded domain. Assume that (1.3) and (1.9) hold. Let g:Ωg:\partial\Omega\to{\mathbb{R}} be bounded and uu be defined by (1.16). Let Σg\Sigma_{g} be given by (4.11). Then uu is a Monge solution of (4.13),  (4.14). If in addition Σg\Sigma_{g}\neq\emptyset and (1.15) holds, then uu is the unique Lipschitz (with respect to LfL_{f}) Monge solution of the Dirichlet problem (4.13)(4.14) satisfying

sup{|u(x)g(y)|:xΩ¯,yΣg,d(x,y)δ}0as δ0.\sup\{|u(x)-g(y)|:x\in\overline{\Omega},\ y\in\Sigma_{g},\ d(x,y)\leq\delta\}\to 0\quad\text{as $\delta\to 0$}. (4.16)
Proof.

We have shown in Theorem 1.1 that uu is a Monge solution in Ω\Omega. Part of the proof holds also for boundary points. In fact, by (4.1) we get, for any fixed x0Ωx_{0}\in\partial\Omega,

u(x0)u(x)Lf(x0,x)for all xΩ¯,u(x_{0})-u(x)\leq L_{f}(x_{0},x)\quad\text{for all $x\in\overline{\Omega}$,}

which yields (4.3) with x0Ωx_{0}\in\partial\Omega. Now, for any x0ΩΣgx_{0}\in\partial\Omega\setminus\Sigma_{g}, since u(x0)<g(x0)u(x_{0})<g(x_{0}), by (1.16), for any ε>0\varepsilon>0 we still have (4.4) for some yΩy\in\partial\Omega and γΓf(x0,y)\gamma\in\Gamma_{f}(x_{0},y). Then we can follow the same argument in the proof of Theorem 1.1 to prove that

lim supxΩ¯Σg,Lf(x,x0)0u(x0)u(x)Lf(x0,x)1.\limsup_{x\in\overline{\Omega}\setminus\Sigma_{g},\ L_{f}(x,x_{0})\to 0}\frac{u(x_{0})-u(x)}{L_{f}(x_{0},x)}\geq 1.

Hence, (4.15) holds for all x0Ω¯Σgx_{0}\in\overline{\Omega}\setminus\Sigma_{g}. In view of (4.12), we also have u=gu=g on Σg\Sigma_{g}. Thus uu is a Monge solution of (4.13),  (4.14).

If Σg\Sigma_{g}\neq\emptyset, we can further obtain (4.16). Indeed, we have (4.5)–(4.8) with Ω\Omega and Ω\partial\Omega replaced by Ω¯Σg\overline{\Omega}\setminus\Sigma_{g} and Σg\Sigma_{g} respectively. Then we deduce (4.16) under the condition (1.15).

The comparison principle, Theorem 3.4, can also be extended to the current case with Ω\Omega and Ω\partial\Omega replaced by Ω¯Σg\overline{\Omega}\setminus\Sigma_{g} and Σg\Sigma_{g} respectively. Hence, the uniqueness of Monge solutions of (4.13)(4.14) with (4.16) holds. ∎

Remark 4.8.

As u<gu<g on ΩΣg\partial\Omega\setminus\Sigma_{g}, by (4.1) we have u(x)<g(y)+Lf(x,y)u(x)<g(y)+L_{f}(x,y) for all xΩ¯x\in\overline{\Omega} and yΩΣgy\in\partial\Omega\setminus\Sigma_{g}. This means that the optimal control formula (1.16) can be rewritten as

u(x)=infyΣg{Lf(x,y)+g(y)}u(x)=\inf_{y\in\Sigma_{g}}\{L_{f}(x,y)+g(y)\}

if Σg\Sigma_{g}\neq\emptyset. In other words, increasing the value of gg at points in ΩΣg\partial\Omega\setminus\Sigma_{g} will not change the solution uu.

In the Euclidean space, when ΣgΩ\Sigma_{g}\neq\partial\Omega and fC(Ω)f\in C(\Omega), we usually relax the meaning of the Dirichlet condition on ΩΣg\partial\Omega\setminus\Sigma_{g}, requiring uu to satisfy

|u|fon ΩΣg|\nabla u|\geq f\quad\text{on $\partial\Omega\setminus\Sigma_{g}$} (4.17)

in the viscosity sense. Under appropriate assumptions on the regularity of Ω\Omega, one can show that the generalized Dirichlet (or state constraint) problem still admits a unique viscosity solution [51, 52, 37]. Our result in Proposition 4.7 essentially handles this type of generalized Dirichlet boundary problems in metric spaces. In fact, we have (4.15) at each x0ΩΣgx_{0}\in\partial\Omega\setminus\Sigma_{g}, which corresponds to the viscosity inequality (4.17). Such topological change was also observed in [42, Remark 5.10] for evolutionary Hamilton-Jacobi equations in metric spaces.

5. Existence of admissible curves and regularity of solutions

We have seen that the existence of admissible curves (1.9) plays an important role in finding the Monge solutions of (1.1). The consistency of local topology condition (1.14), which guarantees the continuity with respect to the original metric dd and uniqueness of the Monge solution uu of the Dirichlet problem, is also related to the existence of admissible curves. We already provided an example, Example 2.4, to show that (1.14) fails to hold in general. It is natural to ask under what assumptions we can obtain (1.9) and (1.14). In this section, we attempt to answer this question.

We introduce some notations and definitions for our use in this section. Let (𝐗,d,μ)({\mathbf{X}},d,\mu) be a metric measure space with μ\mu being a locally finite Borel measure. We use BB to denote an open ball Br(x)B_{r}(x) when the center xx and radius r>0r>0 are irrelevant to our analysis. In this case, we also write λB=Bλr(x)\lambda B=B_{\lambda r}(x) for any λ>0\lambda>0 for simplicity of notation.

Let p1p\geq 1. Let Γ\Gamma be a collection of nonconstant rectifiable curves in (𝐗,d,μ)({\mathbf{X}},d,\mu). We say that a curve connecting x,yXx,y\in X is a CC-quasiconvex curve if the length of the curve is at most Cd(x,y)C\,d(x,y). Let Γ(x,y;C)\Gamma(x,y;C) denote the collection of CC-quasiconvex curves connecting x,y𝐗x,y\in{\mathbf{X}}. A metric space is said to be quasiconvex if there exists C1C\geq 1 such that every pair of points x,y𝐗x,y\in{\mathbf{X}} can be connected by a CC-quasiconvex curve.

Let F(Γ)F(\Gamma) be the family of all Borel measurable functions ρ:X[0,]\rho:X\to[0,\infty] such that γρ1\int_{\gamma}\rho\geq 1 for every γΓ\gamma\in\Gamma. For each 1p<1\leq p<\infty, the pp-modulus of Γ\Gamma is defined as

Modp(Γ)=infρF(Γ)𝐗ρp𝑑μ,\operatorname{Mod}_{p}(\Gamma)=\inf_{\rho\in F(\Gamma)}\int_{\mathbf{X}}\rho^{p}\ d\mu,

and the \infty-modulus of Γ\Gamma is defined as

Mod(Γ)=infρF(Γ)ρL(𝐗).\operatorname{Mod}_{\infty}(\Gamma)=\inf_{\rho\in F(\Gamma)}\|\rho\|_{L^{\infty}({\mathbf{X}})}.

For any given function u:𝐗u:{\mathbf{X}}\to\mathbb{R}, a Borel function ρ:𝐗[0,]\rho:{\mathbf{X}}\to[0,\infty] is said to be a pp-weak upper gradient of uu if

|u(γ(a))u(γ(b))|γρ𝑑s|u(\gamma(a))-u(\gamma(b))|\leq\int_{\gamma}\rho\ ds

holds for all rectifiable paths γ:[a,b]𝐗\gamma:[a,b]\to{\mathbf{X}} outside a family of curves with pp-modulus zero. Consult [30] for an introduction about modulus of curve family and upper gradients.

Let 1p<1\leq p<\infty. A metric measure space (𝐗,d,μ)({\mathbf{X}},d,\mu) is said to support a pp-Poincaré inequality if there exist constants C>0C>0 and λ1\lambda\geq 1 such that for every measurable function u:𝐗u:{\mathbf{X}}\to\mathbb{R} and every upper gradient ρ:𝐗[0,]\rho:{\mathbf{X}}\to[0,\infty] of uu, the pair (u,ρ)(u,\rho) satisfies

B|uuB|𝑑μCdiamB(λBρp𝑑μ)1/p\fint_{B}|u-u_{B}|\,d\mu\leq C{\rm diam\,}B\left(\fint_{\lambda B}\rho^{p}\,d\mu\right)^{1/p} (5.1)

on every ball BXB\subset X. If p=p=\infty, the right hand side above is replaced by CdiamBgL(λB).C{\rm diam\,}B\|g\|_{L^{\infty}(\lambda B)}. Here and in the sequel, for an integrable function ff and a measurable set AA of finite measure, we take

Af𝑑μ:=1μ(A)Af𝑑μ.\fint_{A}f\,d\mu:={1\over\mu(A)}\int_{A}f\,d\mu.

We say that a metric measure space (𝐗,d,μ)({\mathbf{X}},d,\mu) is a PI-space if the measure μ\mu is doubling and the space supports a pp-Poincaré inequality for some p1p\geq 1.

It turns out that (1.9) and (1.14) hold on a broad class of PI-spaces.

Theorem 5.1 (Regularity in pp-PI spaces).

Let (𝐗,d,μ)({\mathbf{X}},d,\mu) be a complete doubling metric measure space supporting a pp-Poincaré inequality with max{1,Q}<p<\max\{1,Q\}<p<\infty, where Q>0Q>0 is the homogeneous dimension satisfying (1.18). Then there exists a constant C>0C>0 depending on the constants of PI-space and R>0R>0 such that for every function uu and its pp-weak upper gradient ρLp(𝐗)\rho\in L^{p}({\mathbf{X}}), the following inequality holds for each open ball BR(z)B_{R}(z) and every x,yBR(z)x,y\in B_{R}(z):

|u(x)u(y)|Cd(x,y)1QpρLp(B5λR(z)),|u(x)-u(y)|\leq Cd(x,y)^{1-\frac{Q}{p}}\|\rho\|_{L^{p}(B_{5\lambda R}(z))}, (5.2)

where λ\lambda is the scaling constant in Poincaré inequality. Furthermore, for any pair of distinct points x,yBR𝐗x,y\in B_{R}\subset{\mathbf{X}}, there exists C>0C>0 depending on the constants of PI-space and R>0R>0 such that

Modp(Γ(x,y))Cd(x,y)pQ,\operatorname{Mod}_{p}(\Gamma(x,y))\geq{C\over d(x,y)^{p-Q}}, (5.3)

where Γ(x,y)\Gamma(x,y) is the family of all rectifiable curves in B5λRB_{5\lambda R} joining xx and yy.

A stronger version of the above result on Ahlfors QQ-regular space with Q>1Q>1 can be found in [21, Theorem 5.1].

Proof of Theorem 5.1.

The proof of (5.2) can be found in [30, Theorem 9.2.14]. We only show the modulus estimate (5.3) below. We only consider a pair of distinct points x,yBR𝐗x,y\in B_{R}\subset{\mathbf{X}} such that Modp(Γ(x,y))<\operatorname{Mod}_{p}(\Gamma(x,y))<\infty. Otherwise, the proof is complete. We pick ρF(Γ(x,y))\rho\in F(\Gamma(x,y)), that is, γρ𝑑s1\int_{\gamma}\rho\,ds\geq 1 for every rectifiable curve γΓ(x,y)\gamma\in\Gamma(x,y). Define

vρ(z)=infγΓ(x,z)γρ𝑑s,v_{\rho}(z)=\inf_{\gamma\in\Gamma(x,z)}\int_{\gamma}\rho\ ds,

and then vρv_{\rho} is measurable [39, Corollary 1.10] and ρ\rho is an upper gradient of vρv_{\rho}. In fact, it is clear that vρ(x)=0v_{\rho}(x)=0 and

vρ(z)γρ𝑑sfor any γΓ(x,z).v_{\rho}(z)\leq\int_{\gamma}\rho\ ds\quad\text{for any $\gamma\in\Gamma(x,z)$}.

Therefore ρ\rho is an upper gradient for vρv_{\rho} in 𝐗{\mathbf{X}}. It follows that ρ\rho is also an upper gradient for min{vρ,2}\min\{v_{\rho},2\}.

Let η:𝐗\eta:{\mathbf{X}}\to\mathbb{R} be a Lipschitz function satisfying that 0η10\leq\eta\leq 1, η=1\eta=1 on 5λB5\lambda B and η=0\eta=0 on 𝐗10λB{\mathbf{X}}\setminus 10\lambda B. Define a function u=ηmin{vρ,2}u=\eta\min\{v_{\rho},2\}. It easily follows that u(x)=0u(x)=0 and u(y)1u(y)\geq 1. One can also verify that g=2|η|+ρ{g}=2|\nabla\eta|+\rho is an upper gradient of uu. In particular, since η=1\eta=1 on 5λB5\lambda B, we obtain ρ\rho is an upper gradient for uu on 5λB5\lambda B. Hence, it follows that

1|u(x)u(y)|\displaystyle 1\leq|u(x)-u(y)| Cd(x,y)1QpρLp(B5λR(z)).\displaystyle\leq Cd(x,y)^{1-\frac{Q}{p}}\|\rho\|_{L^{p}(B_{5\lambda R}(z))}.

Since ρF(Γ(x,y))\rho\in F(\Gamma(x,y)) is arbitrary, by definition we obtain (5.3). ∎

This theorem implies the following result.

Proposition 5.2 (Hölder regularity).

Let (Ω¯,d,μ)(\overline{\Omega},d,\mu) be a complete bounded metric measure space with μ\mu a doubling measure. Assume that Ω¯\overline{\Omega} and ff satisfy (A1). Then for any distinct x,yΩ¯x,y\in\overline{\Omega}, there exists an admissible curve γΓf(x,y)\gamma\in\Gamma_{f}(x,y), and in particular, (1.9) holds. Furthermore, there exists a constant Cf>0C_{f}>0 such that

Lf(x,y)Cfd(x,y)1Qpfor all x,yΩ¯,L_{f}(x,y)\leq C_{f}\,d(x,y)^{1-\frac{Q}{p}}\quad\text{for all $x,y\in\overline{\Omega}$,} (5.4)

which implies that (1.14) holds.

Proof.

Note that for any fLp(Ω¯)f\in L^{p}(\overline{\Omega}) with 1p<1\leq p<\infty, the collection Γ\Gamma of nonconstant rectifiable curves such that γf𝑑s=\int_{\gamma}f\ ds=\infty has pp-modulus zero; see for instance [30, Lemma 5.2.8]. By (5.3) in Theorem 5.1, we get Modp(Γ(x,y))>0\operatorname{Mod}_{p}(\Gamma(x,y))>0 for any distinct points x,yΩ¯x,y\in\overline{\Omega}, which implies the existence of rectifiable curves joining xx and yy on which ff is integrable.

For any fixed xΩ¯x\in\overline{\Omega}, writing v(y)=infγΓf(x,y)γf𝑑sv(y)=\inf_{\gamma\in\Gamma_{f}(x,y)}\int_{\gamma}f\ ds for yΩ¯y\in\overline{\Omega}, we can deduce that vv is measurable and locally pp-integrable by [39, Theorem 1.11] and fLp(Ω¯)f\in L^{p}(\overline{\Omega}) is an upper gradient of vv. Applying Theorem 5.1, we are led to

Lf(x,y)=v(y)=v(y)v(x)Cd(x,y)1QpfLp(Ω¯).L_{f}(x,y)=v(y)=v(y)-v(x)\leq Cd(x,y)^{1-\frac{Q}{p}}\|f\|_{L^{p}(\overline{\Omega})}.

Hence, we also obtain (1.14). ∎

In general, if fLp(Ω)f\notin L^{p}(\Omega) for p>Qp>Q, then (1.9) may not hold in general, as can be seen in the following simple example, which is an adaptation of Example 2.3.

Example 5.3.

Let 𝐗={\mathbf{X}}={\mathbb{R}} with dd being the Euclidean metric and μ\mu being the one-dimensional Lebesgue measure. Let Ω=(1,1)\Omega=(-1,1) and

f(x)=1x,x(1,1){0}.f(x)={1\over x},\quad x\in(-1,1)\setminus\{0\}.

It is clear that fLp(Ω)f\notin L^{p}(\Omega) for any p1p\geq 1. Then for any x(1,1){0}x\in(-1,1)\setminus\{0\}, Γf(x,0)=\Gamma_{f}(x,0)=\emptyset and Lf(x,0)=L_{f}(x,0)=\infty.

The critical case p=Qp=Q is complicated. For n{\mathbb{R}}^{n} with n=1n=1, we can prove the conditions (1.9) and (1.14) if p1p\geq 1, since the pp-modulus of the curve collection containing just one non-constant curve is positive. In general, we cannot expect the results in Proposition 5.2 to always hold if fLQ(Ω)f\in L^{Q}(\Omega) but fLp(Ω)f\notin L^{p}(\Omega) for any p>Qp>Q. On the other hand, (1.9) is only about the existence of one curve on which ff is integrable and can be obtained even for some functions fLQ(Ω)f\notin L^{Q}(\Omega). A simple example is as below.

Example 5.4.

Let 𝐗=2{\mathbf{X}}={\mathbb{R}}^{2} and Ω=B1(O)𝐗\Omega=B_{1}(O)\subset{\mathbf{X}} with OO denoting the origin (0,0)(0,0). That is, Ω\Omega is the unit disk in 2{\mathbb{R}}^{2} centered at OO. We set

e1:={(x1,x2):x1(0,1),x2=0}.e_{1}:=\{(x_{1},x_{2}):x_{1}\in(0,1),x_{2}=0\}.

Let f:Ωf:\Omega\to{\mathbb{R}} be given by

f(x)={1|x1|if xe1,1|x|if xe1 and xO.f(x)=\begin{dcases}{1\over\sqrt{|x_{1}|}}&\text{if $x\in e_{1}$,}\\ {1\over|x|}&\text{if $x\notin e_{1}$ and $x\neq O$.}\end{dcases}

Observe that fLp(Ω)f\in L^{p}(\Omega) for 1p<21\leq p<2 but fL2(Ω)f\notin L^{2}(\Omega); the value of ff on e1e_{1} does not affect the integral of ff in Ω\Omega. However, for each pair of distinct points x,yΩ¯x,y\in\overline{\Omega}, one can find a curve γΓf(x,y)\gamma\in\Gamma_{f}(x,y) in Ω¯\overline{\Omega}. For zOz\neq O, we can take a curve connecting zz first to (|z|,0)(|z|,0) along the circular arc and then to OO along the horizontal line segment. This choice enables us to guarantee Lf(O,z)π+2L_{f}(O,z)\leq\pi+2, which shows that Ω¯\overline{\Omega} a bounded metric space with respect to the metric LfL_{f}. The topology induced by LfL_{f}, however, is distinct from the Euclidean topology. Indeed, for zj:=(0,1/j)z_{j}:=(0,1/j) and any curve γ\gamma joining OO and zjz_{j}, denoting by γ¯\underline{\gamma} the portion of γ\gamma in the Euclidean disk B1/j(zj)B_{1/j}(z_{j}), we have γf𝑑sγ¯f𝑑s(γ¯)/(2j)1/2\int_{\gamma}f\,ds\geq\int_{\underline{\gamma}}f\,ds\geq\ell(\underline{\gamma})/(2j)\geq 1/2, which yields Lf(O,zj)1/2L_{f}(O,z_{j})\geq 1/2 for all j1j\geq 1. Hence, Lf(O,zj)↛0L_{f}(O,z_{j})\not\to 0 as jj\to\infty, although zjz_{j} converges to OO in the Euclidean topology. (As the metric LfL_{f} is locally bi-Lipschitz equivalent to the Euclidean metric in Ω¯{O}\overline{\Omega}\setminus\{O\} as well, we see that this sequence cannot converge in Ω¯\overline{\Omega} with respect to LfL_{f}.)

The following result, which can be found in [21, Theorem 3.1] and [22, Theorem 2.10], is a variant of Theorem 5.1 for the case p=p=\infty.

Theorem 5.5 (Regularity in \infty-PI spaces).

Let (𝐗,d,μ)({\mathbf{X}},d,\mu) be a complete doubling metric measure space. Then the following statements are equivalent.

  1. (1)

    𝐗{\mathbf{X}} supports an \infty-Poincaré inequality.

  2. (2)

    There exists C1C\geq 1 such that Mod(Γ(x,y;C))>0\operatorname{Mod}_{\infty}(\Gamma(x,y;C))>0 for all x,y𝐗x,y\in{\mathbf{X}} satisfying d(x,y)>0d(x,y)>0, where we recall that Γ(x,y;C)\Gamma(x,y;C) denotes the collection of CC-quasiconvex curves connecting x,yx,y.

  3. (3)

    There exists C1C\geq 1 such that whenever N𝐗N\subset{\mathbf{X}} with μ(N)=0\mu(N)=0 and x,y𝐗x,y\in{\mathbf{X}} with xyx\neq y, there is a quasiconvex curve γ\gamma joining x,yx,y such that (γ)Cd(x,y)\ell(\gamma)\leq Cd(x,y) and 1(γ1(N))=0\mathcal{H}^{1}(\gamma^{-1}(N))=0, where 1\mathcal{H}^{1} denotes the one-dimensional Hausdorff measure.

The following is the version of Proposition 5.2 corresponding to p=p=\infty.

Proposition 5.6 (Lipschitz regularity).

Let (Ω¯,d,μ)(\overline{\Omega},d,\mu) be a complete bounded metric measure space with μ\mu a doubling measure. Assume that Ω¯\overline{\Omega} and ff satisfy (A2). Then for any distinct x,yΩ¯x,y\in\overline{\Omega}, there exists a curve γΓ(x,y;C)Γf(x,y)\gamma\in\Gamma(x,y;C)\cap\Gamma_{f}(x,y) in Ω¯\overline{\Omega}. In particular, (1.9) and (1.14) hold with

Lf(x,y)CfL(Ω)d(x,y)for all x,yΩ¯.L_{f}(x,y)\leq C\|f\|_{L^{\infty}(\Omega)}d(x,y)\quad\text{for all $x,y\in\overline{\Omega}$.} (5.5)
Proof.

It is not difficult to see that Theorem 5.5 (2) implies (1.9). The proof of (1.14) involves the condition (3) in Theorem 5.5. In fact, since fL(Ω¯)f\in L^{\infty}(\overline{\Omega}), there exist a null set NfN_{f} such that

ffL(Ω¯)in Ω¯Nf.f\leq\|f\|_{L^{\infty}(\overline{\Omega})}\quad\text{in $\overline{\Omega}\setminus N_{f}$.} (5.6)

Then by Theorem 5.5, for any x,yΩ¯x,y\in\overline{\Omega}, we can take a curve γ\gamma in Ω¯\overline{\Omega} joining them with (γ)Cd(x,y)\ell(\gamma)\leq Cd(x,y) and 1(γ1(Nf))=0\mathcal{H}^{1}(\gamma^{-1}(N_{f}))=0, which yields (5.5). Thus (1.14) clearly holds. ∎

Since the Monge solution uu defined by (1.16) is Lipschitz in the metric LfL_{f}, as shown in Theorem 1.1, we obtain Theorem 1.2 as an immediate consequence of Proposition 5.2 and Proposition 5.6. Under the same assumptions, it is not difficult to show that uu is in the Sobolev class N1,p(Ω¯)N^{1,p}(\overline{\Omega}), that is, uLp(Ω¯)u\in L^{p}(\overline{\Omega}) and uu has an upper gradient in Lp(Ω¯)L^{p}(\overline{\Omega}). In fact, by (1.8) and (4.1) we see that fLp(Ω¯)f\in L^{p}(\overline{\Omega}) is an upper gradient of uu.

6. Solutions for admissible curves transversal to null sets

In this section, we assume that ff satisfies (1.3) and (Ω¯,d,μ)(\overline{\Omega},d,\mu) is a complete doubling metric measure space with homogeneous dimention Q>0Q>0. Assuming either of the additional regularity assumptions (A1) and (A2) as stated in the discussion following (1.18), we consider a slightly different optical length function L~f\tilde{L}_{f} and the associated solutions using curves transversal to null sets in Ω¯\overline{\Omega}. This change improves the stability of solutions with respect to the perturbation on ff. We refer to [46, 44] for stability results on time-dependent Hamilton-Jacobi equations in metric spaces.

6.1. Optical length with transversal curves

Let ΓfN(x,y){\Gamma}_{f}^{N}(x,y) be the collection of arc-length parametrized rectifiable curves in 𝐗{\mathbf{X}} connecting x,yΩ¯x,y\in\overline{\Omega} with γf𝑑s<\int_{\gamma}f\ ds<\infty and transversal to a given null set NΩ¯N\subset\overline{\Omega}, as given in (1.19). The associated optical length function is defined by (1.20).

Under the standing assumptions of this section, we consider fN=f+χNf_{N}=f+\infty\chi_{N} if (A1) holds. Note that fNLp(Ω¯)f_{N}\in L^{p}(\overline{\Omega}) with max{1,Q}<p<\max\{1,Q\}<p<\infty. Proposition 5.2 then implies that ΓfN(x,y)\Gamma_{f_{N}}(x,y)\neq\emptyset if xyx\neq y, and we can verify that ΓfN(x,y)=ΓfN(x,y)\Gamma_{f_{N}}(x,y)=\Gamma_{f}^{N}(x,y) and

γf𝑑s=γfN𝑑sfor γΓfN(x,y).\int_{\gamma}f\,ds=\int_{\gamma}f_{N}\,ds\quad\text{for $\gamma\in\Gamma_{f}^{N}(x,y)$.}

If (A2) holds, then Theorem 5.5 implies that ΓfN(x,y)\Gamma^{N}_{f}(x,y)\neq\emptyset and therefore LfN(x,y)<L^{N}_{f}(x,y)<\infty for any distinct points x,yΩ¯x,y\in\overline{\Omega}. Hence, in either case we have

LfN(x,y)=inf{γf𝑑s:γΓfN(x,y)}=inf{γf𝑑s:γΓfN(x,y)}=LfN(x,y).L_{f_{N}}(x,y)=\inf\left\{\int_{\gamma}f\,ds:\ \text{$\gamma\in\Gamma_{f_{N}}(x,y)$}\right\}=\inf\left\{\int_{\gamma}f\,ds:\ \text{$\gamma\in\Gamma_{f}^{N}(x,y)$}\right\}=L_{f}^{N}(x,y). (6.1)

We define the maximal optical length function L~f:𝐗×𝐗\tilde{L}_{f}:{\mathbf{X}}\times{\mathbf{X}}\to\mathbb{R} by (1.22). It is easy to verify using Proposition 5.6 and Proposition 5.2 that the optical length function L~f\tilde{L}_{f} (as well as LfNL^{N}_{f} for any fixed null set NN) also satisfies the metric properties, as discussed for LfL_{f} in Lemma 2.1.

Moreover, we see that (1.14) holds for L~f\tilde{L}_{f} under either of the assumptions (A1), (A2). Indeed, it is an immediate consequence of Proposition 5.2 when (A1) holds. In the case of (A2), there exists a null set NfN_{f} such that (5.6) holds. Since Ω¯\overline{\Omega} supports an \infty-Poincaré inequality, by Theorem 5.5, there exists C>0C>0 such that for any null set NN and x,yΩ¯x,y\in\overline{\Omega} we can find a curve transversal to NfNN_{f}\cup N satisfying

γf𝑑sCfL(Ω)d(x,y).\int_{\gamma}f\,ds\leq C\|f\|_{L^{\infty}(\Omega)}\,d(x,y).

Recall that Nf={x:f(x)>fL(Ω¯)}N_{f}=\{x\,:\,f(x)>\|f\|_{L^{\infty}(\overline{\Omega})}\}. It follows that

L~f(x,y)CfL(Ω)d(x,y)for all x,yΩ¯\tilde{L}_{f}(x,y)\leq C\|f\|_{L^{\infty}(\Omega)}\,d(x,y)\quad\text{for all $x,y\in\overline{\Omega}$} (6.2)

and thus (1.14) holds for L~f\tilde{L}_{f}. Furthermore, under the assumption (1.3), together with (2.1), we also obtain

L~f(x,y)αd(x,y)for all x,yΩ¯.\tilde{L}_{f}(x,y)\geq\alpha d(x,y)\quad\text{for all $x,y\in\overline{\Omega}$.} (6.3)

Since ΓfN(x,y)Γf(x,y)\Gamma^{N}_{f}(x,y)\subset\Gamma_{f}(x,y) for each null set NN, it is easily seen that

Lf(x,y)L~f(x,y)for all x,yΩ¯.L_{f}(x,y)\leq\tilde{L}_{f}(x,y)\quad\text{for all $x,y\in\overline{\Omega}$.} (6.4)

It turns out that the supremum in the definition of L~f\tilde{L}_{f} can be attained at a particular null set if either (A1) or (A2) holds.

Lemma 6.1.

Let (Ω¯,d,μ)(\overline{\Omega},d,\mu) be a complete metric measure space with μ\mu a doubling measure. If (A1) or (A2) holds, then there exists a null set EE such that L~f\tilde{L}_{f} defined in (1.22) satisfies

L~f(x,y)=LfE(x,y)for all x,yΩ¯.\tilde{L}_{f}(x,y)=L^{E}_{f}(x,y)\quad\text{for all $x,y\in\overline{\Omega}$.} (6.5)
Proof.

It suffices to show L~fLfE\tilde{L}_{f}\leq L^{E}_{f}, since the reverse inequality holds obviously. We first fix a null set N0N_{0}, and we now show that for any fixed x,yΩ¯x,y\in\overline{\Omega} there exists a null set NxyN_{xy} containing N0N_{0} such that

L~f(x,y)=inf{If(γ):γΓNxy(x,y)}.\tilde{L}_{f}(x,y)=\inf\{I_{f}(\gamma):\gamma\in\Gamma^{N_{xy}}(x,y)\}.

To see this, for any kk\in\mathbb{N}, we take a null set NkN_{k} (k1k\geq 1) such that

L~f(x,y)LfNk(x,y)+1kLfNkN0(x,y)+1k.\tilde{L}_{f}(x,y)\leq L^{N_{k}}_{f}(x,y)+\frac{1}{k}\leq L^{N_{k}\cup N_{0}}_{f}(x,y)+\frac{1}{k}.

Set Nxy=k=0NkN_{xy}=\bigcup_{k=0}^{\infty}N_{k}. Then,

L~f(x,y)LfNxy(x,y)LfNk(x,y)L~f(x,y)1k\tilde{L}_{f}(x,y)\geq L^{N_{xy}}_{f}(x,y)\geq L^{N_{k}}_{f}(x,y)\geq\tilde{L}_{f}(x,y)-\frac{1}{k}

and thus Lf(x,y)=LfNxy(x,y)L_{f}(x,y)=L_{f}^{N_{xy}}(x,y) by letting kk\to\infty. The claim is proved.

Since (Ω¯,d,μ)(\overline{\Omega},d,\mu) is complete and doubling, it follows that Ω¯\overline{\Omega} is separable [30, Lemma 4.1.13], and therefore there exists a countable dense subset DΩ¯D\subset\overline{\Omega}. For every pair of points x0,y0Dx_{0},y_{0}\in D, there exists a null set Nx0y0N_{x_{0}y_{0}} containing NN such that

L~f(x0,y0)=LfNx0y0(x0,y0)=inf{If(γ):γΓNx0y0(x0,y0)}.\tilde{L}_{f}(x_{0},y_{0})=L_{f}^{N_{x_{0}y_{0}}}(x_{0},y_{0})=\inf\{I_{f}(\gamma):\gamma\in{\Gamma}^{N_{x_{0}y_{0}}}(x_{0},y_{0})\}.

We construct a null set EE in slightly different ways for the cases (A1) and (A2).

If (A1) holds, we set E=x0,y0DNx0y0E=\bigcup_{x_{0},y_{0}\in D}N_{x_{0}y_{0}}. Then we have μ(E)=0\mu(E)=0 and

L~f(x0,y0)=LfE(x0,y0)=inf{If(γ):γΓE(x0,y0)}for all x0,y0D.\tilde{L}_{f}(x_{0},y_{0})=L^{E}_{f}(x_{0},y_{0})=\inf\{I_{f}(\gamma):\gamma\in{\Gamma}^{E}(x_{0},y_{0})\}\quad\text{for all $x_{0},y_{0}\in D$.} (6.6)

Fix x,yΩ¯x,y\in\overline{\Omega} and ε>0\varepsilon>0 arbitrarily. Set ε=(ε/(4C))ppQ\varepsilon^{\prime}=\left(\varepsilon/(4C)\right)^{\frac{p}{p-Q}} with constant C>0C>0 from (5.4) and let x0DBε(x)x_{0}\in D\cap B_{\varepsilon^{\prime}}(x) and y0DBε(y)y_{0}\in D\cap B_{\varepsilon^{\prime}}(y). Proposition 5.2 implies that

L~f(x,x0)<ε/4.\tilde{L}_{f}(x,x_{0})<\varepsilon/4. (6.7)

There exists a null set N1N_{1}^{\prime} containing EE such that

L~f(x,x0)=LfN1(x,x0)=inf{If(γ):γΓN1(x,x0)}.\tilde{L}_{f}(x,x_{0})=L^{N_{1}^{\prime}}_{f}(x,x_{0})=\inf\{I_{f}(\gamma):\gamma\in{\Gamma}^{N_{1}^{\prime}}(x,x_{0})\}. (6.8)

For f~=f+χN1\tilde{f}=f+\infty\chi_{N_{1}^{\prime}}, we can apply Theorem 5.1 to find a curve transversal to N1N_{1}^{\prime} connecting xx and x0x_{0}. In particular, there exists a curve γ1\gamma_{1} such that

If(γ1)LfN1(x,x0)+ε4=L~f(x,x0)+ε4<ε2.I_{f}(\gamma_{1})\leq L^{N_{1}^{\prime}}_{f}(x,x_{0})+\frac{\varepsilon}{4}=\tilde{L}_{f}(x,x_{0})+\frac{\varepsilon}{4}<\frac{\varepsilon}{2}. (6.9)

Analogously, there exists a curve γ2\gamma_{2} transversal to a null set N2N_{2}^{\prime} containing EE and connecting y,y0y,y_{0} in Ω¯\overline{\Omega} such that

L~f(y,y0)If(γ2)<ε/2.\tilde{L}_{f}(y,y_{0})\leq I_{f}(\gamma_{2})<\varepsilon/2. (6.10)

Choose any curve γΓE(x,y)\gamma\in\Gamma^{E}(x,y) and let ξ\xi be a curve connecting x0x_{0} and y0y_{0} defined by joining γ1\gamma_{1}, γ\gamma and γ2\gamma_{2}. Therefore we obtain

L~f(x,y)\displaystyle\tilde{L}_{f}(x,y) L~f(x,x0)+L~f(x0,y0)+L~f(y0,y)\displaystyle\leq\tilde{L}_{f}(x,x_{0})+\tilde{L}_{f}(x_{0},y_{0})+\tilde{L}_{f}(y_{0},y) (6.11)
ε+I(ξ)=ε+If(γ1)+If(γ2)+If(γ)If(γ)+2ε.\displaystyle\leq\varepsilon+I(\xi)=\varepsilon+I_{f}(\gamma_{1})+I_{f}(\gamma_{2})+I_{f}(\gamma)\leq I_{f}(\gamma)+2\varepsilon.

Since ε\varepsilon is arbitratry, it follows that

L~f(x,y)inf{If(γ):γΓx,yE}=LfE(x,y)\tilde{L}_{f}(x,y)\leq\inf\{I_{f}(\gamma):\gamma\in{\Gamma}_{x,y}^{E}\}=L_{f}^{E}(x,y)

for all x,yΩ¯x,y\in\overline{\Omega}. We have completed the proof of (6.5) under the assumption (A1).

If, on the other hand, (A2) holds, then there exists a null set such that (5.6) holds. We define E=x0,y0D(Nx0y0Nf)E=\bigcup_{x_{0},y_{0}\in D}(N_{x_{0}y_{0}}\cup N_{f}) in this case. It is clear that EE is still a null set. For every ε>0\varepsilon>0, let ε=ε/(2Cβ)\varepsilon^{\prime}=\varepsilon/(2C\beta). Then for any x,yΩ¯x,y\in\overline{\Omega} and x0DBε(x)x_{0}\in D\cap B_{\varepsilon^{\prime}}(x) and y0DBε(y)y_{0}\in D\cap B_{\varepsilon^{\prime}}(y), there exists a null set N1N_{1}^{\prime} containing EE such that (6.8) holds also in this case. Thanks to the \infty-Poincaré inequality, we adopt (5.5) in Proposition 5.6 to get (6.7). The rest of the proof is the same as that under the assumption (A1). We find curves γ1\gamma_{1} and γ2\gamma_{2} satisfying estimates (6.9) and (6.10), and build ξ\xi again by concatenating γ1,γ\gamma_{1},\gamma and γ2\gamma_{2}. The same estimate as in (6.11) yields (6.5) in this case. ∎

6.2. Transversal Monge solutions

Using the optical length L~f\tilde{L}_{f} (or equivalently LfEL^{E}_{f} with EE given in Lemma 6.1), we can consider a different notion of Monge solutions, which we call transversal Monge solutions.

Definition 6.2 (Transversal Monge solutions).

We say that a locally bounded function u:Ωu:\Omega\to{\mathbb{R}} is a transversal Monge solution (resp. subsolution, supersolution) to (1.1) in Ω\Omega if for any x0Ωx_{0}\in\Omega

lim supxx0u(x0)u(x)L~f(x,x0)=1(resp.,).\limsup_{x\to x_{0}}\frac{u(x_{0})-u(x)}{\tilde{L}_{f}(x,x_{0})}=1\ \ \ (\text{resp.}\ \leq,\geq). (6.12)
Remark 6.3.

Under assumption (A2), by (6.2) and (6.3), we can show that (6.12) holds if and only if

lim supxx0u(x0)u(x)L~f(x,x0)d(x,x0)=0(resp.,).\limsup_{x\to x_{0}}\frac{u(x_{0})-u(x)-\tilde{L}_{f}(x,x_{0})}{{d}(x,x_{0})}=0\ \ \ (\text{resp.}\ \leq,\geq). (6.13)

Indeed, (6.12) is equivalent to

lim supxx0u(x0)u(x)L~f(x,x0)L~f(x,x0)=0(resp.,),\limsup_{x\to x_{0}}\frac{u(x_{0})-u(x)-\tilde{L}_{f}(x,x_{0})}{\tilde{L}_{f}(x,x_{0})}=0\ \ \ (\text{resp.}\ \leq,\geq),

which is further equivalent to (6.13), since we have

αL~f(x,x0)d(x,x0)CfL(Ω).\alpha\leq{\tilde{L}_{f}(x,x_{0})\over d(x,x_{0})}\leq C\|f\|_{L^{\infty}(\Omega)}.
Remark 6.4.

Following Remark 3.2, we can show that |u|(x0)=f(x0)|\nabla^{-}u|(x_{0})=f(x_{0}) holds for any transversal Monge solution and any x0Ωx_{0}\in\Omega where ff is continuous.

We can repeat our arguments in the previous section, replacing LfL_{f} by L~f\tilde{L}_{f}, to prove the uniqueness and existence of transversal Monge solutions of (1.1)(1.2). Note that thanks to (A1) or (A2), (Ω¯,L~f)(\overline{\Omega},\tilde{L}_{f}) is bounded if (Ω,d)(\Omega,d) is bounded. Below we give the statements without proofs.

Theorem 6.5.

Let (𝐗,d,μ)({\mathbf{X}},d,\mu) be a complete geodesic space with μ\mu doubling and let Ω𝐗\Omega\subsetneq{\mathbf{X}} be a bounded domain. Assume that ff satisfies (1.3) and that either (A1) or (A2) holds. Let uUSCL~(Ω¯)u\in USC_{\tilde{L}}(\overline{\Omega}) and vLSCL~(Ω¯)v\in LSC_{\tilde{L}}(\overline{\Omega}) be respectively a Monge subsolution and a Monge supersolution to (1.1) in the sense of Definition 6.2. If uvu\leq v on Ω\partial\Omega, then uvu\leq v in Ω¯.\overline{\Omega}.

Theorem 6.6.

Let (𝐗,d,μ)({\mathbf{X}},d,\mu) be a complete geodesic space with μ\mu doubling and let Ω𝐗\Omega\subsetneq{\mathbf{X}} be a bounded domain. Assume that ff satisfies (1.3) and that either (A1) or (A2) holds. Let g:Ωg:\partial\Omega\to{\mathbb{R}} be bounded, and let u~\tilde{u} be defined by (1.21), where L~f\tilde{L}_{f} is given by (1.22). Then u~\tilde{u} is a transversal Monge solution of (1.1), which is Lipschitz continuous with respect to L~f\widetilde{L}_{f} and uniformly continuous with respect to dd in Ω¯\overline{\Omega}. Moreover, if gg satisfies

g(x)L~f(x,y)+g(y)for all x,yΩg(x)\leq\tilde{L}_{f}(x,y)+g(y)\quad\text{for all $x,y\in\partial\Omega$, } (6.14)

then u~\tilde{u} is the unique transversal Monge solution of (1.1)(1.2).

Remark 6.7.

Recall that a metric space is said to be proper if every closed and bounded set is compact. A complete metric measure space (X,d,μ)(X,d,\mu) with μ\mu doubling is proper [30, Lemma 4.1.14]. Furthermore, Hopf-Rinow Theorem implies that a complete and proper length space is a geodesic space.

The Lipschitz continuity of u~\tilde{u} can be obtained via the relation

u~(x)u~(y)+L~f(x,y)for all x,yΩ¯,\tilde{u}(x)\leq\tilde{u}(y)+\tilde{L}_{f}(x,y)\quad\text{for all $x,y\in\overline{\Omega}$,} (6.15)

which is a counterpart of (4.1) in the current setting.

By (6.4), we see that in general the Monge solution uu and transversal Monge solution u~\tilde{u} of (1.1)(1.2) satisfy uu~u\leq\tilde{u} in Ω¯\overline{\Omega} under the same given boundary data.

6.3. Maximal weak solutions

In the Euclidean space, for bounded continuous inhomogeneous term ff, one can define weak solutions of (1.1) by requiring the function to be locally Lipschitz and satisfies the equation almost everywhere in Ω\Omega. Such kind of solutions are also called Lipschitz a.e. solutions in the literature. We extend this notion to metric measure spaces for a possibly discontinuous ff.

Definition 6.8.

We say that uLiploc(Ω)C(Ω¯)u\in{\rm Lip_{\rm loc}}(\Omega)\cap C(\overline{\Omega}) is a weak solution (resp. subsolution, supersolution) to (1.1) if the slope |u|(x)|\nabla u|(x) exists and satisfies |u|(x)=f(x)|\nabla u|(x)=f(x) (resp. \leq, \geq) at almost every xΩx\in\Omega.

When the measure μ\mu on Ω¯\overline{\Omega} is doubling and supports an \infty-Poincaré inequality, we know from [20] that uu is Lipschitz continuous with respect to the metric dd and that |u||\nabla u| is the least \infty-weak upper gradient of uu. Our construction of uu in this setting yields |u|f|\nabla u|\leq f and in general we may not have equality. However, if fL(Ω¯)f\in L^{\infty}(\overline{\Omega}) is continuous almost everywhere, we do obtain that f=|u|f=|\nabla u|. This is the focus of this subsection.

Let us denote the class of weak subsolutions to (1.1), (1.2) by

Sweak:={vLiploc(Ω)C(Ω¯):|v(x)|f(x) a.e. in Ω and vg on Ω}.S_{weak}:=\{v\in{\rm Lip_{\rm loc}}(\Omega)\cap C(\overline{\Omega}):\ \text{$|\nabla v(x)|\leq f(x)$ a.e. in $\Omega$ and $v\leq g$ on $\partial\Omega$}\}. (6.16)

We can show that the maximal weak subsolution is the transversal Monge solution provided that ff is upper semicontinuous almost everywhere and the metric space 𝐗{\mathbf{X}} satisfies the \infty-weak Fubini property. A metric measure space (𝐗,d,μ)({\mathbf{X}},d,\mu) is said to satisfy the \infty-weak Fubini property if for any null set NN and ε>0\varepsilon>0, given any distinct points x,y𝐗x,y\in{\mathbf{X}}, there exists a curve connecting x,yx,y transversal to NN such that (γ)d(x,y)+ε\ell(\gamma)\leq d(x,y)+\varepsilon. In particular, a space satisfying the \infty-weak Fubini necessarily supports an \infty-Poincaré inequality. More discussion on the \infty-weak Fubini property can be found in [22, Section 4].

Proposition 6.9.

Let (𝐗,d,μ)({\mathbf{X}},d,\mu) be a complete geodesic space with μ\mu a doubling measure. Suppose that Ω𝐗\Omega\subsetneq{\mathbf{X}} is a bounded domain with (Ω¯,d,μ)(\overline{\Omega},d,\mu) satisfying the \infty-weak Fubini property. Assume that fL(Ω¯)f\in L^{\infty}(\overline{\Omega}) and that ff satisfies (1.3). Let g:Ωg:\partial\Omega\to{\mathbb{R}} satisfy (6.14) and u~\tilde{u} be the transversal Monge solution defined in (1.21). Then, vu~v\leq\tilde{u} in Ω¯\overline{\Omega} holds for all vSweakv\in S_{weak}. If in addition ff is assumed to be upper semicontinuous almost everywhere, i.e., there exists a set NΩN\subset\Omega with μ(N)=0\mu(N)=0 such that f|ΩNf|_{{\Omega}\setminus N} is upper semicontinuous, then u~\tilde{u} is a weak subsolution of (1.1), (1.2). In particular, u~\tilde{u} is the maximal weak subsolution in the sense that

u~(x)=sup{v(x):vSweak}for all xΩ.\tilde{u}(x)=\sup\{v(x):v\in S_{weak}\}\quad\text{for all $x\in\Omega$.} (6.17)
Remark 6.10.

If we further assume that 𝐗{\mathbf{X}} supports a pp-Poincaré inequality for some 1<p<1<p<\infty, then we have gu=|u|g_{u}=|\nabla u| almost everywhere [15, Theorem 6.1], where gug_{u} denotes the least pp-weak upper gradient of uu. In particular, the function u0u_{0} defined in (1.16) is the maximal Lipschitz solution to the Dirichlet problem

{gu(x)=f(x)a.e. in Ωu=gon Ω,\begin{cases}g_{u}(x)=f(x)&\quad\quad\text{a.e. in $\Omega$}\\ u=g&\quad\quad\text{on $\partial\Omega$,}\end{cases}

where f,gf,g satisfy the conditions in Proposition 6.9.

Proof of Proposition 6.9.

We have seen in Theorem 6.6 that u~\tilde{u} is uniformly continuous in Ω¯\overline{\Omega} with respect to the metric dd. Let us show that vu~v\leq\tilde{u} in Ω\Omega for every vSweakv\in S_{weak}. Take the null set Nv={xΩ:|v|(x)>f(x)}N_{v}=\{x\in\Omega:|\nabla v|(x)>f(x)\}. Note that |v|(x)|\nabla v|(x) is an upper gradient of vv, i.e.,

|v(x)v(y)|γ|v|𝑑s|v(x)-v(y)|\leq\int_{\gamma}|\nabla v|\,ds

holds for every curve γ\gamma joining xx and yy in Ω¯\overline{\Omega}; see [30, Lemma 6.2.6]. Let N=NvEN^{\prime}=N_{v}\cup E, where EE is the null set given in Lemma 6.1. It is clear that NN^{\prime} is still a null set. We fix yΩy\in\partial\Omega. For any ε>0\varepsilon>0, we can choose a curve γ\gamma in Ω¯\overline{\Omega} connecting xx and yy such that it is transversal to NN^{\prime} and

γf𝑑s+g(y)u~(x)+ε.\int_{\gamma}f\ ds+g(y)\leq\tilde{u}(x)+\varepsilon. (6.18)

This is possible because Ω¯\overline{\Omega} supports the \infty-weak Fubini property. Due to the transversality, we get

γ|v|𝑑sγf𝑑s.\int_{\gamma}|\nabla v|\ ds\leq\int_{\gamma}f\ ds.

Hence, it follows that

v(x)=v(x)v(y)+v(y)\displaystyle v(x)=v(x)-v(y)+v(y) |v(x)v(y)|+g(y)\displaystyle\leq|v(x)-v(y)|+g(y)
0|v|𝑑s+g(y)γf𝑑s+g(y).\displaystyle\leq\int_{0}^{\ell}|\nabla v|\,ds+g(y)\leq\int_{\gamma}f\ ds+g(y).

By (6.18), we thus have v(x)u~(x)+εv(x)\leq\tilde{u}(x)+\varepsilon. Since ε\varepsilon is arbitrary, it follows that v(x)u~(x)v(x)\leq\tilde{u}(x) for every xΩx\in\Omega.

We next prove that u~\tilde{u} is a weak solution of (1.1) under the upper semicontinuity of f|ΩNf|_{\Omega\setminus N} for a null set NN. Let N=NEN_{\ast}=N\cup E and fix x0ΩNx_{0}\in\Omega\setminus N_{\ast} arbitrarily. In view of Remark 6.4, u~\tilde{u} satisfies |u~|(x0)|u~|(x0)=f(x0)|\nabla\tilde{u}|(x_{0})\geq|\nabla^{-}\tilde{u}|(x_{0})=f(x_{0}). In what follows we show that |u~|(x0)f(x0)|\nabla\tilde{u}|(x_{0})\leq f(x_{0}). Then by the upper semicontinuity of f|ΩNf|_{\Omega\setminus N}, for every 0<ε<10<\varepsilon<1 we can take r>0r>0 small such that

f(y)f(x0)+εfor all yB2r(x0)Nf(y)\leq f(x_{0})+\varepsilon\quad\text{for all $y\in B_{2r}(x_{0})\setminus N_{\ast}$} (6.19)

Using the \infty-weak Fubini property of Ω¯\overline{\Omega}, for any xBr(x0)x\in B_{r}(x_{0}) it is possible to connect x0x_{0} and xx by a curve γ\gamma in Ω\Omega that is transversal to NN_{\ast} and satisfies (γ)(1+ε)d(x,x0)<2r\ell(\gamma)\leq(1+\varepsilon)d(x,x_{0})<2r. This implies that d(x0,y)<2rd(x_{0},y)<2r for any yγy\in\gamma; in other words, γ\gamma lies in B2r(x0)B_{2r}(x_{0}). It follows from (6.19) and the transversality of γ\gamma to NN_{\ast} that

γf𝑑s(f(x0)+ε)(γ)(f(x0)+ε)(1+ε)d(x,x0).\int_{\gamma}f\,ds\leq(f(x_{0})+\varepsilon)\ell(\gamma)\leq(f(x_{0})+\varepsilon)(1+\varepsilon)d(x,x_{0}).

Applying (6.5) in Lemma 6.1, we therefore obtain

L~f(x,x0)=LfE(x,x0)LfN(x,x0)(f(x0)+ε)(1+ε)d(x,x0)\tilde{L}_{f}(x,x_{0})=L_{f}^{E}(x,x_{0})\leq L_{f}^{N_{\ast}}(x,x_{0})\leq(f(x_{0})+\varepsilon)(1+\varepsilon)d(x,x_{0})

for all xΩx\in\Omega with d(x,x0)>0d(x,x_{0})>0 small.

In view of the Lipschitz continuity of u~\tilde{u} in (6.15), we get

|u~(x)u~(x0)|(f(x0)+ε)(1+ε)d(x,x0).|\tilde{u}(x)-\tilde{u}(x_{0})|\leq(f(x_{0})+\varepsilon)(1+\varepsilon)d(x,x_{0}).

Dividing the inequality by d(x,x0)d(x,x_{0}), letting d(x,x0)0d(x,x_{0})\to 0 and then sending ε0\varepsilon\to 0, we end up with |u~|(x0)f(x0)|\nabla\tilde{u}|(x_{0})\leq f(x_{0}). Since x0ΩNx_{0}\in\Omega\setminus N_{\ast} is arbitrary and NN_{\ast} is a null set, we see that u~\tilde{u} is a weak subsolution of (1.1). Note that u~g\tilde{u}\leq g holds on Ω\partial\Omega thanks to the definition (1.21). Hence, u~Sweak\tilde{u}\in S_{weak}. Combining this with the first part of our result, we are led to (6.17). ∎

Remark 6.11.

We can define the class of weak supersolution and solution by replacing “\leq" in (6.16) by ``"``\geq" and ``="``=" respectively. If f|ΩNf|_{\Omega\setminus N} is further assumed to be continuous, then by Remark 6.4, u~\tilde{u} satisfies |u~|(x0)=f(x0)|\nabla^{-}\tilde{u}|(x_{0})=f(x_{0}) for any x0ΩNx_{0}\in\Omega\setminus N_{\ast}. It immediately follows that |u~|(x0)f(x0)|\nabla\tilde{u}|(x_{0})\geq f(x_{0}) due to the property |u~||u~||\nabla\tilde{u}|\geq|\nabla^{-}\tilde{u}|. In other words, u~\tilde{u} is a weak supersolution of (1.1) since u~g\tilde{u}\geq g follows from (6.14). Since u~\tilde{u} is also a weak subsolution, we see that u~\tilde{u} is a weak solution.

Remark 6.12.

If instead of the condition fL(Ω¯)f\in L^{\infty}(\overline{\Omega}), we impose (A1), that is, Ω¯\overline{\Omega} satisfies a pp-Poincaré inequality and fLp(Ω¯)f\in L^{p}(\overline{\Omega}) for finite p>max{1,Q}p>\max\{1,Q\}, then the problem becomes more challenging. One may replace the definition of SweakS_{weak} in (6.16) by

Sweak:={vC(Ω¯):\displaystyle S_{weak}:=\{v\in C(\overline{\Omega}): vg on Ω, there exists a null set Nv such that f\displaystyle v\leq g\text{ on }\partial\Omega,\,\text{ there exists a null set $N_{v}$ such that $f$}
is the upper gradient of v along all curves transversal to Nv}.\displaystyle\text{is the upper gradient of $v$ along all curves transversal to $N_{v}$}\}.

Then the same conclusion as in Proposition 6.9 follows from a similar argument.

References

  • [1] Y. Achdou, F. Camilli, A. Cutrì, and N. Tchou. Hamilton-Jacobi equations constrained on networks. NoDEA Nonlinear Differential Equations Appl., 20(3):413–445, 2013.
  • [2] P. Albano. On the eikonal equation for degenerate elliptic operators. Proc. Amer. Math. Soc., 140(5):1739–1747, 2012.
  • [3] P. Albano, P. Cannarsa, and T. Scarinci. Regularity results for the minimum time function with Hörmander vector fields. J. Differential Equations, 264(5):3312–3335, 2018.
  • [4] L. Ambrosio and J. Feng. On a class of first order Hamilton-Jacobi equations in metric spaces. J. Differential Equations, 256(7):2194–2245, 2014.
  • [5] L. Ambrosio, N. Gigli, and G. Savaré. Heat flow and calculus on metric measure spaces with Ricci curvature bounded below—the compact case. In Analysis and numerics of partial differential equations, volume 4 of Springer INdAM Ser., pages 63–115. Springer, Milan, 2013.
  • [6] M. Bardi and I. Capuzzo-Dolcetta. Optimal control and viscosity solutions of Hamilton-Jacobi-Bellman equations. Systems & Control: Foundations & Applications. Birkhäuser Boston Inc., Boston, MA, 1997. With appendices by Maurizio Falcone and Pierpaolo Soravia.
  • [7] T. Bieske. The Carnot-Carathéodory distance vis-à-vis the eikonal equation and the infinite Laplacian. Bull. Lond. Math. Soc., 42(3):395–404, 2010.
  • [8] A. Briani and A. Davini. Monge solutions for discontinuous Hamiltonians. ESAIM Control Optim. Calc. Var., 11(2):229–251 (electronic), 2005.
  • [9] D. Y. Burago. Periodic metrics. In Representation theory and dynamical systems, volume 9 of Adv. Soviet Math., pages 205–210. Amer. Math. Soc., Providence, RI, 1992.
  • [10] L. Caffarelli, M. G. Crandall, M. Kocan, and A. Swi\polhk ech. On viscosity solutions of fully nonlinear equations with measurable ingredients. Comm. Pure Appl. Math., 49(4):365–397, 1996.
  • [11] F. Camilli. An Hopf-Lax formula for a class of measurable Hamilton-Jacobi equations. Nonlinear Anal., 57(2):265–286, 2004.
  • [12] F. Camilli and A. Siconolfi. Hamilton-Jacobi equations with measurable dependence on the state variable. Adv. Differential Equations, 8(6):733–768, 2003.
  • [13] F. Camilli and A. Siconolfi. Time-dependent measurable Hamilton-Jacobi equations. Comm. Partial Differential Equations, 30(4-6):813–847, 2005.
  • [14] P. Cardaliaguet. Notes on mean field games (from P.-L. Lions’ lectures at collége de france). https://www.ceremade.dauphine.fr/ cardaliaguet/MFG20130420.pdf, 2013.
  • [15] J. Cheeger. Differentiability of Lipschitz functions on metric measure spaces. Geom. Funct. Anal., 9(3):428–517, 1999.
  • [16] Q. Chen, S. A. R. Horsley, N. J. G. Fonseca, T. Tyc, and O. Quevedo-Teruel. Double-layer geodesic and gradient-index lenses. Nature Communications, 13(1):2354, 2022.
  • [17] X. Chen and B. Hu. Viscosity solutions of discontinuous Hamilton-Jacobi equations. Interfaces Free Bound., 10(3):339–359, 2008.
  • [18] M. G. Crandall, M. Kocan, P. L. Lions, and A. Świ\polhkech. Existence results for boundary problems for uniformly elliptic and parabolic fully nonlinear equations. Electron. J. Differential Equations, pages No. 24, 22 pp.  1999.
  • [19] K. Deckelnick and C. M. Elliott. Uniqueness and error analysis for Hamilton-Jacobi equations with discontinuities. Interfaces Free Bound., 6(3):329–349, 2004.
  • [20] E. Durand-Cartagena and J. A. Jaramillo. Pointwise Lipschitz functions on metric spaces. J. Math. Anal. Appl., 363(2):525–548, 2010.
  • [21] E. Durand-Cartagena, J. A. Jaramillo, and N. Shanmugalingam. Geometric characterizations of pp-Poincaré inequalities in the metric setting. Publ. Mat., 60(1):81–111, 2016.
  • [22] E. Durand-Cartagena, J. A. Jaramillo, and N. Shanmugalingam. Existence and uniqueness of \infty-harmonic functions under assumption of \infty-Poincaré inequality. Math. Ann., 374(1-2):881–906, 2019.
  • [23] I. Ekeland. On the variational principle. J. Math. Anal. Appl., 47:324–353, 1974.
  • [24] I. Ekeland. Nonconvex minimization problems. Bull. Amer. Math. Soc. (N.S.), 1(3):443–474, 1979.
  • [25] F. Essebei, G. Giovannardi, and S. Verzellesi. Monge solutions for discontinuous Hamilton-Jacobi equations in Carnot groups, preprint, 2023.
  • [26] W. Gangbo and A. Świ\polhkech. Metric viscosity solutions of Hamilton-Jacobi equations depending on local slopes. Calc. Var. Partial Differential Equations, 54(1):1183–1218, 2015.
  • [27] Y. Giga, P. Górka, and P. Rybka. A comparison principle for Hamilton-Jacobi equations with discontinuous Hamiltonians. Proc. Amer. Math. Soc., 139(5):1777–1785, 2011.
  • [28] Y. Giga and N. Hamamuki. Hamilton-Jacobi equations with discontinuous source terms. Comm. Partial Differential Equations, 38(2):199–243, 2013.
  • [29] Y. Giga, N. Hamamuki, and A. Nakayasu. Eikonal equations in metric spaces. Trans. Amer. Math. Soc., 367(1):49–66, 2015.
  • [30] J. Heinonen, P. Koskela, N. Shanmugalingam, and J. T. Tyson. Sobolev spaces on metric measure spaces, volume 27 of New Mathematical Monographs. Cambridge University Press, Cambridge, 2015. An approach based on upper gradients.
  • [31] J. Hong, W. Cheng, S. Hu, and K. Zhao. Representation formulas for contact type Hamilton-Jacobi equations. J. Dynam. Differential Equations, 34(3):2315–2327, 2022.
  • [32] R. A. Houstoun. Note on Einstein’s theory of gravitation. Philos. Mag. (7), 33:899–903, 1942.
  • [33] C. Imbert and R. Monneau. Flux-limited solutions for quasi-convex Hamilton-Jacobi equations on networks. Ann. Sci. Éc. Norm. Supér. (4), 50(2):357–448, 2017.
  • [34] C. Imbert and R. Monneau. Quasi-convex Hamilton-Jacobi equations posed on junctions: The multi-dimensional case. Discrete Contin. Dyn. Syst., 37(12):6405–6435, 2017.
  • [35] C. Imbert, R. Monneau, and H. Zidani. A Hamilton-Jacobi approach to junction problems and application to traffic flows. ESAIM Control Optim. Calc. Var., 19(1):129–166, 2013.
  • [36] H. Ishii. Hamilton-Jacobi equations with discontinuous Hamiltonians on arbitrary open sets. Bull. Fac. Sci. Engrg. Chuo Univ., 28:33–77, 1985.
  • [37] H. Ishii. A boundary value problem of the Dirichlet type for Hamilton-Jacobi equations. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4), 16(1):105–135, 1989.
  • [38] H. Ishii and H. Mitake. Representation formulas for solutions of Hamilton-Jacobi equations with convex Hamiltonians. Indiana Univ. Math. J., 56(5):2159–2183, 2007.
  • [39] E. Järvenpää, M. Järvenpää, K. Rogovin, S. Rogovin, and N. Shanmugalingam. Measurability of equivalence classes and MECp{\rm MEC}_{p}-property in metric spaces. Rev. Mat. Iberoam., 23(3):811–830, 2007.
  • [40] P.-L. Lions. Generalized solutions of Hamilton-Jacobi equations, volume 69 of Research Notes in Mathematics. Pitman (Advanced Publishing Program), Boston, Mass.-London, 1982.
  • [41] Q. Liu and A. Mitsuishi. Principal eigenvalue problem for infinity Laplacian in metric spaces. Adv. Nonlinear Stud., 22(1):548–573, 2022.
  • [42] Q. Liu and A. Nakayasu. Convexity preserving properties for Hamilton-Jacobi equations in geodesic spaces. Discrete Contin. Dyn. Syst., 39(1):157–183, 2019.
  • [43] Q. Liu, N. Shanmugalingam, and X. Zhou. Equivalence of solutions of eikonal equation in metric spaces. J. Differential Equations, 272:979–1014, 2021.
  • [44] S. Makida. Stability of viscosity solutions on expanding networks. preprint, 2023.
  • [45] A. Nakayasu. Metric viscosity solutions for Hamilton-Jacobi equations of evolution type. Adv. Math. Sci. Appl., 24(2):333–351, 2014.
  • [46] A. Nakayasu and T. Namba. Stability properties and large time behavior of viscosity solutions of Hamilton-Jacobi equations on metric spaces. Nonlinearity, 31(11):5147–5161, 2018.
  • [47] R. T. Newcomb, II and J. Su. Eikonal equations with discontinuities. Differential Integral Equations, 8(8):1947–1960, 1995.
  • [48] D. N. Ostrov. Extending viscosity solutions to eikonal equations with discontinuous spatial dependence. Nonlinear Anal., 42(4, Ser. A: Theory Methods):709–736, 2000.
  • [49] D. Schieborn and F. Camilli. Viscosity solutions of Eikonal equations on topological networks. Calc. Var. Partial Differential Equations, 46(3-4):671–686, 2013.
  • [50] B. D. Seckler and J. B. Keller. Geometrical theory of diffraction in inhomogeneous media. J. Acoust. Soc. Amer., 31:192–205, 1959.
  • [51] H. M. Soner. Optimal control with state-space constraint. I. SIAM J. Control Optim., 24(3):552–561, 1986.
  • [52] H. M. Soner. Optimal control with state-space constraint. II. SIAM J. Control Optim., 24(6):1110–1122, 1986.
  • [53] P. Soravia. Boundary value problems for Hamilton-Jacobi equations with discontinuous Lagrangian. Indiana Univ. Math. J., 51(2):451–477, 2002.
  • [54] P. Soravia. Degenerate eikonal equations with discontinuous refraction index. ESAIM Control Optim. Calc. Var., 12(2):216–230, 2006.
  • [55] A. Tourin. A comparison theorem for a piecewise Lipschitz continuous Hamiltonian and application to shape-from-shading problems. Numer. Math., 62(1):75–85, 1992.
  • [56] H. V. Tran and Y. Yu. Optimal convergence rate for periodic homogenization of convex Hamilton-Jacobi equations, 2022.
  • [57] C. Villani. Optimal transport, volume 338 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 2009. Old and new.