Skip to main content

A Short Introduction to Viscosity Solutions and the Large Time Behavior of Solutions of Hamilton–Jacobi Equations

  • Chapter
  • First Online:
Hamilton-Jacobi Equations: Approximations, Numerical Analysis and Applications

Part of the book series: Lecture Notes in Mathematics ((LNMCIME,volume 2074))

Abstract

We present an introduction to the theory of viscosity solutions of first-order partial differential equations and a review on the optimal control/dynamical approach to the large time behavior of solutions of Hamilton–Jacobi equations, with the Neumann boundary condition. This article also includes some of basics of mathematical analysis related to the optimal control/dynamical approach for easy accessibility to the topics.

In memory of Riichi Iino, my former adviser at Waseda University.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 44.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 59.95
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  1. L. Ambrosio, P. Tilli, Topics on Analysis in Metric Spaces. Oxford Lecture Series in Mathematics and Its Applications, vol. 25 (Oxford University Press, Oxford, 2004), viii + 133 pp

    Google Scholar 

  2. M. Bardi, I. Capuzzo-Dolcetta, Optimal Control and Viscosity Solutions of Hamilton-Jacobi-Bellman Equations. Systems & Control: Foundations & Applications (Birkhäuser Boston, Inc., Boston, 1997), xviii + 570 pp

    Google Scholar 

  3. G. Barles, P.E. Souganidis, On the large time behavior of solutions of Hamilton-Jacobi equations. SIAM J. Math. Anal. 31(4), 925–939 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  4. G. Barles, An introduction to the theory of viscosity solutions for first-order Hamilton-Jacobi equations and applications, in Hamilton-Jacobi Equations: Approximations, Numerical Analysis and Applications, ed. by P. Loreti, N.A. Tchou. Lecture Notes in Mathematics (Springer, Berlin/Heidelberg, 2013)

    Google Scholar 

  5. G. Barles, Discontinuous viscosity solutions of first-order Hamilton-Jacobi equations: a guided visit. Nonlinear Anal. 20(9), 1123–1134 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  6. G. Barles, Solutions de Viscosité des Équations de Hamilton-Jacobi. Mathématiques & Applications (Berlin), vol. 17 (Springer, Paris, 1994), x + 194 pp

    Google Scholar 

  7. G. Barles, H. Ishii, H. Mitake, A new PDE approach to the large time asymptotics of solutions of Hamilton-Jacobi equations. Bull. Math. Sci. (to appear)

    Google Scholar 

  8. G. Barles, H. Ishii, H. Mitake, On the large time behavior of solutions of Hamilton-Jacobi equations associated with nonlinear boundary conditions. Arch. Ration. Mech. Anal. 204(2), 515–558 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  9. G. Barles, H. Mitake, A pde approach to large-time asymptotics for boundary-value problems for nonconvex Hamilton-Jacobi equations. Commun. Partial Differ. Equ. 37(1), 136–168 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  10. G. Barles, J.-M. Roquejoffre, Ergodic type problems and large time behaviour of unbounded solutions of Hamilton-Jacobi equations. Commun. Partial Differ. Equ. 31(7–9), 1209–1225 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  11. E.N. Barron, R. Jensen, Semicontinuous viscosity solutions for Hamilton-Jacobi equations with convex Hamiltonians. Commun. Partial Differ. Equ. 15(12), 1713–1742 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  12. P. Bernard, Existence of C 1, 1 critical sub-solutions of the Hamilton-Jacobi equation on compact manifolds. Ann. Sci. École Norm. Sup. (4) 40(3), 445–452 (2007)

    Google Scholar 

  13. P. Bernard, J.-M. Roquejoffre, Convergence to time-periodic solutions in time-periodic Hamilton-Jacobi equations on the circle. Commun. Partial Differ. Equ. 29(3–4), 457–469 (2004)

    MathSciNet  MATH  Google Scholar 

  14. G. Buttazzo, M. Giaquinta, S. Hildebrandt, One-Dimensional Variational Problems. An Introduction. Oxford Lecture Series in Mathematics and Its Applications, vol. 15 (The Clarendon Press/Oxford University Press, New York, 1998), viii + 262 pp

    Google Scholar 

  15. L.A. Caffarelli, X. Cabré, Fully Nonlinear Elliptic Equations. American Mathematical Society Colloquium Publications, vol. 43 (American Mathematical Society, Providence, 1995), vi + 104 pp

    Google Scholar 

  16. F. Camilli, O. Ley, P. Loreti, V.D. Nguyen, Large time behavior of weakly coupled systems of first-order Hamilton-Jacobi equations. Nonlinear Differ. Equ. Appl. (NoDEA) 19(6), 719–749 (2012)

    Google Scholar 

  17. I. Capuzzo-Dolcetta, P.-L. Lions, Hamilton-Jacobi equations with state constraints. Trans. Am. Math. Soc. 318(2), 643–683 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  18. M.G. Crandall, L.C. Evans, P.-L. Lions, Some properties of viscosity solutions of Hamilton-Jacobi equations. Trans. Am. Math. Soc. 282(2), 487–502 (1984)

    Article  MathSciNet  MATH  Google Scholar 

  19. M.G. Crandall, H. Ishii, P.-L. Lions, User’s guide to viscosity solutions of second order partial differential equations. Bull. Am. Math. Soc. (N.S.) 27(1), 1–67 (1992)

    Google Scholar 

  20. M.G. Crandall, P.-L. Lions, Viscosity solutions of Hamilton-Jacobi equations. Trans. Am. Math. Soc. 277(1), 1–42 (1983)

    Article  MathSciNet  MATH  Google Scholar 

  21. A. Davini, A. Siconolfi, A generalized dynamical approach to the large time behavior of solutions of Hamilton-Jacobi equations. SIAM J. Math. Anal. 38(2), 478–502 (2006)

    Article  MathSciNet  Google Scholar 

  22. W. E, Aubry-Mather theory and periodic solutions of the forced Burgers equation. Commun. Pure Appl. Math. 52(7), 811–828 (1999)

    Google Scholar 

  23. L.C. Evans, On solving certain nonlinear partial differential equations by accretive operator methods. Isr. J. Math. 36(3–4), 225–247 (1980)

    Article  MATH  Google Scholar 

  24. L.C. Evans, A survey of partial differential equations methods in weak KAM theory. Commun. Pure Appl. Math. 57(4) 445–480 (2004)

    Article  MATH  Google Scholar 

  25. A. Fathi, Théorème KAM faible et théorie de Mather sur les systèmes lagrangiens. C. R. Acad. Sci. Paris Sér. I Math. 324(9), 1043–1046 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  26. A. Fathi, Sur la convergence du semi-groupe de Lax-Oleinik. C. R. Acad. Sci. Paris Sér. I Math. 327(3), 267–270 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  27. A. Fathi, Weak KAM theorem in Lagrangian dynamics. Cambridge University Press (to appear)

    Google Scholar 

  28. A. Fathi, A. Siconolfi, Existence of C 1 critical subsolutions of the Hamilton-Jacobi equation. Invent. Math. 155(2), 363–388 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  29. W.H. Fleming, H. Mete Soner, Controlled Markov Processes and Viscosity Solutions, 2nd edn. Stochastic Modelling and Applied Probability, vol. 25 (Springer, New York, 2006), xviii + 429 pp

    Google Scholar 

  30. Y. Fujita, H. Ishii, P. Loreti, Asymptotic solutions of Hamilton-Jacobi equations in Euclidean n space. Indiana Univ. Math. J. 55(5), 1671–1700 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  31. Y. Giga, Surface Evolution Equations. A Level Set Approach. Monographs in Mathematics, vol. 99 (Birkhäuser, Basel, 2006), xii + 264 pp

    Google Scholar 

  32. Y. Giga, Q. Liu, H. Mitake, Singular Neumann problems and large-time behavior of solutions of non-coercive Hamiton-Jacobi equations. Tran. Amer. Math. Soc. (to appear)

    Google Scholar 

  33. Y. Giga, Q. Liu, H. Mitake, Large-time asymptotics for one-dimensional Dirichlet problems for Hamilton-Jacobi equations with noncoercive Hamiltonians. J. Differ. Equ. 252(2), 1263–1282 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  34. N. Ichihara, H. Ishii, Asymptotic solutions of Hamilton-Jacobi equations with semi-periodic Hamiltonians. Commun. Partial Differ. Equ. 33(4–6), 784–807 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  35. N. Ichihara, H. Ishii, The large-time behavior of solutions of Hamilton-Jacobi equations on the real line. Methods Appl. Anal. 15(2), 223–242 (2008)

    MathSciNet  MATH  Google Scholar 

  36. N. Ichihara, H. Ishii, Long-time behavior of solutions of Hamilton-Jacobi equations with convex and coercive Hamiltonians. Arch. Ration. Mech. Anal. 194(2), 383–419 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  37. H. Ishii, Asymptotic solutions for large time of Hamilton-Jacobi equations in Euclidean. Ann. Inst. Henri Poincaré Anal. Non Linéaire 25(2), 231–266 (2008)

    Article  MATH  Google Scholar 

  38. H. Ishii, Long-time asymptotic solutions of convex Hamilton-Jacobi equations with Neumann type boundary conditions. Calc. Var. Partial Differ. Equ. 42(1–2), 189–209 (2011)

    Article  MATH  Google Scholar 

  39. H. Ishii, Weak KAM aspects of convex Hamilton-Jacobi equations with Neumann type boundary conditions. J. Math. Pures Appl. (9) 95(1), 99–135 (2011)

    Google Scholar 

  40. H. Ishii, H. Mitake, Representation formulas for solutions of Hamilton-Jacobi equations with convex Hamiltonians. Indiana Univ. Math. J. 56(5), 2159–2183 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  41. S. Koike, A Beginner’s Guide to the Theory of Viscosity Solutions. MSJ Memoirs, vol. 13 (Mathematical Society of Japan, Tokyo, 2004), viii + 123 pp

    Google Scholar 

  42. P.-L. Lions, Generalized Solutions of Hamilton-Jacobi Equations. Research Notes in Mathematics, vol. 69 (Pitman, Boston, 1982), iv + 317 pp

    Google Scholar 

  43. P.-L. Lions, Neumann type boundary conditions for Hamilton-Jacobi equations. Duke Math. J. 52(4), 793–820 (1985)

    Article  MathSciNet  MATH  Google Scholar 

  44. P.-L. Lions, A.-S. Sznitman, Stochastic differential equations with reflecting boundary conditions. Commun. Pure Appl. Math. 37(4), 511–537 (1984)

    Article  MathSciNet  MATH  Google Scholar 

  45. P.-L. Lions, N.S. Trudinger, Linear oblique derivative problems for the uniformly elliptic Hamilton-Jacobi-Bellman equation. Math. Z. 191(1), 1–15 (1986)

    Article  MathSciNet  MATH  Google Scholar 

  46. J.N. Mather, Variational construction of connecting orbits. Ann. Inst. Fourier (Grenoble) 43(5), 1349–1386 (1993)

    Google Scholar 

  47. J.N. Mather, Total disconnectedness of the quotient Aubry set in low dimensions. Dedicated to the memory of Jürgen K. Moser. Commun. Pure Appl. Math. 56(8), 1178–1183 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  48. H. Mitake, Asymptotic solutions of Hamilton-Jacobi equations with state constraints. Appl. Math. Optim. 58(3), 393–410 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  49. H. Mitake, The large-time behavior of solutions of the Cauchy-Dirichlet problem. Nonlinear Differ. Equ. Appl. (NoDEA) 15(3), 347–362 (2008)

    Google Scholar 

  50. H. Mitake, Large time behavior of solutions of Hamilton-Jacobi equations with periodic boundary data. Nonlinear Anal. 71(11), 5392–5405 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  51. H. Mitake, H.V. Tran, Remarks on the large time behavior of viscosity solutions of quasi-monotone weakly coupled systems of Hamilton-Jacobi equations. Asymptot. Anal. 77(1–2), 43–70 (2012)

    MathSciNet  MATH  Google Scholar 

  52. H. Mitake, H.V. Tran, A dynamical approach to the large-time behavior of solutions to weakly coupled systems of Hamilton-Jacobi equations. J. Math. Pures Appl. (to appear)

    Google Scholar 

  53. G. Namah, J.-M. Roquejoffre, Remarks on the long time behaviour of the solutions of Hamilton-Jacobi equations. Commun. Partial Differ. Equ. 24(5–6), 883–893 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  54. J.-M. Roquejoffre, Convergence to steady states or periodic solutions in a class of Hamilton-Jacobi equations. J. Math. Pures Appl. (9) 80(1), 85–104 (2001)

    Google Scholar 

  55. H.M. Soner, Optimal control with state-space constraint, I. SIAM J. Control Optim. 24(3), 552–561 (1986)

    Article  MathSciNet  MATH  Google Scholar 

  56. M. Spivak, Calculus on Manifolds. A Modern Approach to Classical Theorems of Advanced Calculus (W.A. Benjamin, Inc., New York, 1965), xii + 144 pp

    Google Scholar 

  57. E.M. Stein, R. Shakarchi, Real Analysis. Measure Theory, Integration, and Hilbert Spaces. Princeton Lectures in Analysis, vol. III (Princeton University Press, Princeton, 2005), xx + 402 pp

    Google Scholar 

  58. E. Yokoyama, Y. Giga, P. Rybka, A microscopic time scale approximation to the behavior of the local slope on the faceted surface under a nonuniformity in supersaturation. Phys. D 237(22), 2845–2855 (2008)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

Supported in part by JSPS KAKENHI (#20340019, #21340032, #21224001, #23340028, and #23244015).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Hitoshi Ishii .

Appendix

Appendix

1.1 A.1 Local maxima to global maxima

We recall a proposition from [56] which is about partition of unity.

Proposition A.1.

Let \(\mathcal{O}\) be a collection of open subsets of \({\mathbb{R}}^{n}\) . Set \(W :=\bigcup _{U\in \mathcal{O}}U\) . Then there is a collection \(\mathcal{F}\) of \({C}^{\infty }\) functions in \({\mathbb{R}}^{n}\) having the following properties:

  1. (i)

    0 ≤ f(x) ≤ 1 for all x ∈ W and \(f \in \mathcal{F}\).

  2. (ii)

    For each x ∈ W there is a neighborhood V of x such that all but finitely many \(f \in \mathcal{F}\) vanish in V.

  3. (iii)

    \(\sum _{f\in \mathcal{F}}f(x) = 1\) for all x ∈ W.

  4. (iv)

    For each \(f \in \mathcal{F}\) there is a set \(U \in \mathcal{O}\) such that supp f ⊂ U.

Proposition A.2.

Let Ω be any subset of \({\mathbb{R}}^{n}\) ,\(u \in \mathrm{USC}(\Omega, \mathbb{R})\) and ϕ ∈ C 1 (Ω). Assume that u −ϕ attains a local maximum at y ∈ Ω. Then there is a function ψ ∈ C 1 (Ω) such that u −ψ attains a global maximum at y and ψ = ϕ in a neighborhood of y.

Proof.

As usual it is enough to prove the above proposition in the case when (u − ϕ)(y) = 0.

By the definition of the space C 1(Ω), there is an open neighborhood W 0 of Ω such that ϕ is defined in W 0 and \(\phi \in {C}^{1}(W_{0})\).

There is an open subset U y  ⊂ W 0 of \({\mathbb{R}}^{n}\) containing y such that \(\max _{U_{y}\cap \Omega }(u\,-\,\phi ) = (u-\phi )(y)\). Since \(u \in \mathrm{USC}(\Omega, \mathbb{R})\), for each x ∈ Ω ∖ { y} we may choose an open subset U x of \({\mathbb{R}}^{n}\) so that x ∈ U x , y ∉ U x and \(\sup _{U_{x}\cap \Omega }u < \infty \). Set \(a_{x} =\sup _{U_{x}\cap \Omega }u\) for every x ∈ Ω ∖ { y}.

We set \(\mathcal{O} =\{ U_{z}\, :\, z \in \Omega \}\) and \(W =\bigcup _{U\in \mathcal{O}}U\). Note that W is an open neighborhood of Ω. By Proposition A.1, there exists a collection \(\mathcal{F}\) of functions \(f \in {C}^{\infty }({\mathbb{R}}^{n})\) satisfying the conditions (i)–(iv) of the proposition. According to the condition (iv), for each \(f \in \mathcal{F}\) there is a point z ∈ Ω such that \(\mathrm{supp}f \subset U_{z}\). For each \(f \in \mathcal{F}\) we fix such a point z ∈ Ω and define the mapping \(p\, :\, \mathcal{F}\rightarrow \Omega \) by p(f) = z. We set

$$\displaystyle{\psi (x) =\sum _{f\in \mathcal{F},\ p(f)\not =y}a_{p(f)}f(x) +\sum _{f\in \mathcal{F},\ p(f)=y}\phi (x)f(x)\ \ \ \text{ for }x \in W.}$$

By the condition (ii), we see that ψ ∈ C 1(W). Fix any x ∈ Ω and \(f \in \mathcal{F}\), and observe that if f(x) > 0 and p(f) ≠ y, then we have \(x \in \mathrm{supp}f \subset U_{p(f)}\) and, therefore, \(a_{p(f)} =\sup _{U_{p(f)}\cap \Omega }u \geq u(x)\). Observe also that if f(x) > 0 and p(f) = y, then we have x ∈ U y and ϕ(x) ≥ u(x). Thus we see that for all x ∈ Ω,

$$\displaystyle{\psi (x) \geq \sum _{f\in \mathcal{F},\ p(f)\not =y}u(x)f(x) +\sum _{f\in \mathcal{F},\ p(f)=y}u(x)f(x) = u(x)\sum _{f\in \mathcal{F}}f(x) = u(x).}$$

Thanks to the condition (ii), we may choose a neighborhood V ⊂ W of y and a finite subset \(\{f_{j}\}_{j=1}^{N}\) of \(\mathcal{F}\) so that

$$\displaystyle{\sum _{j=1}^{N}f_{ j}(x) = 1\ \ \ \text{ for all }x \in V.}$$

If p(f j ) ≠ y for some j = 1, , N, then \(U_{p(f_{j})} \cap \{ y\} = \varnothing \) and hence \(y\not\in \mathrm{supp}f_{j}\). Therefore, by replacing V by a smaller one we may assume that p(f j ) = y for all j = 1, , N. Since f = 0 in V for all \(f \in \mathcal{F}\setminus \{ f_{1},\ldots,f_{N}\}\), we see that

$$\displaystyle{\psi (x) =\sum _{ j=1}^{N}\phi (x)f_{ j}(x) =\phi (x)\ \ \ \text{ for all }x \in V.}$$

It is now easy to see that u − ψ has a global maximum at y. □ 

1.2 A.2 A Quick Review of Convex Analysis

We discuss here basic properties of convex functions on \({\mathbb{R}}^{n}\).

By definition, a subset C of \({\mathbb{R}}^{n}\) is convex if and only if

$$\displaystyle{(1 - t)x + ty \in C\ \ \ \text{ for all }\ x,y \in C,\,0 < t < 1.}$$

For a given function \(f\, :\, U \subset {\mathbb{R}}^{n} \rightarrow [-\infty,\,\infty ]\), its epigraph epi(f) is defined as

$$\displaystyle{\mathrm{epi}(f) =\{ (x,y) \in U \times \mathbb{R}\, :\, y \geq f(x)\}.}$$

A function f : U → [ − , ] is said to be convex if epi(f) is a convex subset of \({\mathbb{R}}^{n+1}\).

We are henceforth concerned with functions defined on \({\mathbb{R}}^{n}\). When we are given a function f on U with U being a proper subset of \({\mathbb{R}}^{n}\), we may think of f as a function defined on \({\mathbb{R}}^{n}\) having value on the set \({\mathbb{R}}^{n} \setminus U\).

It is easily checked that a function \(f\, :\, {\mathbb{R}}^{n} \rightarrow [-\infty,\,\infty ]\) is convex if and only if for all \(x,y \in {\mathbb{R}}^{n}\), \(t,s \in \mathbb{R}\) and λ ∈ [0, 1],

$$\displaystyle{f((1-\lambda )x +\lambda y) \leq (1-\lambda )t +\lambda s\ \ \ \mathrm{if}\ \ t \geq f(x)\ \text{ and }\ s \geq f(y).}$$

From this, we see that a function \(f\, :\, {\mathbb{R}}^{n} \rightarrow (-\infty,\,\infty ]\) is convex if and only if for all \(x,y \in {\mathbb{R}}^{n}\) and λ ∈ [0, 1],

$$\displaystyle{f((1-\lambda )x +\lambda y) \leq (1-\lambda )f(x) +\lambda f(y).}$$

Here we use the convention for extended real numbers, i.e., for any \(x \in \mathbb{R}\), −  < x < , x ±  =  ± , x ⋅( ± ) =  ±  if x > 0, 0 ⋅( ± ) = 0, etc.

Any affine function f(x) = ax + b, where \(a \in {\mathbb{R}}^{n}\) and \(b \in \mathbb{R}\), is a convex function on \({\mathbb{R}}^{n}\). Moreover, if \(A \subset {\mathbb{R}}^{n}\) and \(B \subset \mathbb{R}\) are nonempty sets, then the function on \({\mathbb{R}}^{n}\) given by

$$\displaystyle{f(x) =\sup \{ a \cdot x + b\, :\, (a,b) \in A \times B\}}$$

is a convex function. Note that this function f is lower semicontinuous on \({\mathbb{R}}^{n}\). We restrict our attention to those functions which take values only in ( − , ].

Proposition B.1.

Let \(f\, :\, {\mathbb{R}}^{n} \rightarrow (-\infty,\,\infty ]\) be a convex function. Assume that p ∈ D f(y) for some \(y,p \in {\mathbb{R}}^{n}\) . Then

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

Proof.

By the definition of D  −  f(y), we have

$$\displaystyle{f(x) \geq f(y) + p \cdot (x - y) + o(\vert x - y\vert )\ \ \ \text{ as }\ x \rightarrow y.}$$

Hence, fixing \(x \in {\mathbb{R}}^{n}\), we get

$$\displaystyle{f(y) \leq f(tx + (1 - t)y) - tp \cdot (x - y) + o(t)\ \ \ \text{ as }\ t \rightarrow 0 +.}$$

Using the convexity of f, we rearrange the above inequality and divide by t > 0, to get

$$\displaystyle{f(y) \leq f(x) - p \cdot (x - y) + o(1)\ \ \ \text{ as }\ t \rightarrow 0 +.}$$

Sending t → 0 + yields

$$\displaystyle{f(x) \geq f(y) + p \cdot (x - y)\ \ \ \text{ for all }\,\,x \in {\mathbb{R}}^{n}.}$$

 □ 

Proposition B.2.

Let \(\mathcal{F}\) be a nonempty set of convex functions on \({\mathbb{R}}^{n}\) with values in (−∞, ∞]. Then \(\sup \mathcal{F}\) is a convex function on \({\mathbb{R}}^{n}\) having values in (−∞, ∞].

Proof.

It is clear that \((\sup \mathcal{F})(x) \in (-\infty,\,\infty ]\) for all \(x \in {\mathbb{R}}^{n}\). If \(f \in \mathcal{F}\), \(x,y \in {\mathbb{R}}^{n}\) and t ∈ [0, 1], then we have

$$\displaystyle{f((1 - t)x + ty) \leq (1 - t)f(x) + tf(y) \leq (1 - t)(\sup \mathcal{F})(x) + t(\sup \mathcal{F})(y)}$$

and hence

$$\displaystyle{(\sup \mathcal{F})((1 - t)x + ty) \leq (1 - t)(\sup \mathcal{F})(x) + t(\sup \mathcal{F})(y),}$$

which proves the convexity of \(\sup \mathcal{F}\). □ 

We call a function \(f\, :\, {\mathbb{R}}^{n} \rightarrow (-\infty,\,\infty ]\) proper convex if the following three conditions hold:

  1. (a)

    f is convex on \({\mathbb{R}}^{n}\).

  2. (b)

    \(f \in \mathrm{LSC}({\mathbb{R}}^{n})\).

  3. (c)

    f(x) ≢ .

Let \(f\, :\, {\mathbb{R}}^{n} \rightarrow [-\infty,\,\infty ]\). The conjugate convex function (or the Legendre–Fenchel transform) of f is the function \({f}^{\star }\, :\, {\mathbb{R}}^{n} \rightarrow [-\infty,\,\infty ]\) given by

$$\displaystyle{{f}^{\star }(x) =\sup _{ y\in {\mathbb{R}}^{n}}(x \cdot y - f(y)).}$$

Proposition B.3.

If f is a proper convex function, then so is f .

Lemma B.1.

If f is a proper convex function on \({\mathbb{R}}^{n}\) , then D f(y)≠∅ for some \(y \in {\mathbb{R}}^{n}\).

Proof.

We choose a point \(x_{0} \in {\mathbb{R}}^{n}\) so that \(f(x_{0}) \in \mathbb{R}\). Let \(k \in \mathbb{N}\), and define the function g k on \(\bar{B}_{1}(x_{0})\) by the formula g k (x) = f(x) + k | x − x 0 | 2. Since \(g_{k} \in \mathrm{LSC}(\overline{B}_{1}(x_{0}))\), and \(g_{k}(x_{0}) = g(x_{0}) \in \mathbb{R}\), the function g k has a finite minimum at a point \(x_{k} \in \overline{B}_{1}(x_{0})\). Note that if k is sufficiently large, then

$$\displaystyle{\min _{\partial B_{1}(x_{0})}g_{k} =\min _{\partial B_{1}(x_{0})}f + k > f(x_{0}).}$$

Fix such a large k, and observe that x k  ∈ B 1(x 0) and, therefore, \(-2k(x_{k} - x_{0}) \in {D}^{-}f(x_{k})\). □ 

Proof (Proposition B.3). 

The function \(x\mapsto x \cdot y - f(y)\) is an affine function for any \(y \in {\mathbb{R}}^{n}\). By Proposition B.2, the function f  ⋆  is convex on \({\mathbb{R}}^{n}\). Also, since the function \(x\mapsto x \cdot y - f(y)\) is continuous on \({\mathbb{R}}^{n}\) for any \(y \in {\mathbb{R}}^{n}\), as stated in Proposition 1.5, the function f  ⋆  is lower semicontinuous on \({\mathbb{R}}^{n}\).

Since f is proper convex on \({\mathbb{R}}^{n}\), there is a point \(x_{0} \in {\mathbb{R}}^{n}\) such that \(f(x_{0}) \in \mathbb{R}\). Hence, we have

$$\displaystyle{{f}^{\star }(y) \geq y \cdot x_{ 0} - f(x_{0}) > -\infty \ \ \ \text{ for all }\ y \in {\mathbb{R}}^{n}.}$$

By Lemma B.1, there exist points \(y,p \in {\mathbb{R}}^{n}\) such that p ∈ D  −  f(y). By Proposition B.1, we have

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

That is,

$$\displaystyle{p \cdot y - f(y) \geq p \cdot x - f(x)\ \ \ \text{ for all }\ x \in {\mathbb{R}}^{n},}$$

which implies that \({f}^{\star }(p) = p \cdot y - f(y) \in \mathbb{R}\). Thus, we conclude that \({f}^{\star }\, :\, {\mathbb{R}}^{n} \rightarrow (-\infty,\,\infty ]\), f  ⋆  is convex on \({\mathbb{R}}^{n}\), \({f}^{\star } \in \mathrm{LSC}({\mathbb{R}}^{n})\) and f  ⋆ (x) ≢ . □ 

The following duality (called convex duality or Legendre–Fenchel duality) holds.

Theorem B.1.

Let \(f\, :\, {\mathbb{R}}^{n} \rightarrow (-\infty,\,\infty ]\) be a proper convex function. Then

$$\displaystyle{{f}^{\star \star } = f.}$$

Proof.

By the definition of f  ⋆ , we have

$$\displaystyle{{f}^{\star }(x) \geq x \cdot y - f(y)\ \ \ \text{ for all }\ x,y \in {\mathbb{R}}^{n},}$$

which reads

$$\displaystyle{f(y) \geq y \cdot x - {f}^{\star }(x)\ \ \ \text{ for all }\ x,y \in {\mathbb{R}}^{n}.}$$

Hence,

$$\displaystyle{f(y) \geq {f}^{\star \star }(y)\ \ \ \text{ for all }\ y \in {\mathbb{R}}^{n}.}$$

Next, we show that

$$\displaystyle{{f}^{\star \star }(x) \geq f(x)\ \ \ \text{ for all }\ x \in {\mathbb{R}}^{n}.}$$

We fix any \(a \in {\mathbb{R}}^{n}\) and choose a point \(y \in {\mathbb{R}}^{n}\) so that \(f(y) \in \mathbb{R}\). We fix a number R > 0 so that | y − a |  < R. Let \(k \in \mathbb{N}\), and consider the function \(g_{k} \in \mathrm{LSC}(\overline{B}_{R}(a))\) defined by g k (x) = f(x) + k | x − a | 2. Let \(x_{k} \in \overline{B}_{R}(a)\) be a minimum point of the function g k . Noting that if k is sufficiently large, then

$$\displaystyle{g_{k}(x_{k}) \leq f(y) + k\vert y - a{\vert }^{2} <\min _{ \partial B_{R}(a)}f + k{R}^{2} =\min _{ \partial B_{R}(a)}g_{k},}$$

we see that x k  ∈ B R (a) for k sufficiently large. We henceforth assume that k is large enough so that x k  ∈ B R (a). We have

$$\displaystyle{{D}^{-}g_{ k}(x_{k}) = {D}^{-}f(x_{ k}) + 2k(x_{k} - a) \ni 0.}$$

Accordingly, if we set ξ k  =  − 2k(x k  − a), then we have \(\xi _{k} \in {D}^{-}f(x_{k})\). By Proposition B.1, we get

$$\displaystyle{f(x) \geq f(x_{k}) +\xi _{k} \cdot (x - x_{k})\ \ \ \text{ for all }\ x \in {\mathbb{R}}^{n},}$$

or, equivalently,

$$\displaystyle{\xi _{k} \cdot x_{k} - f(x_{k}) \geq \xi _{k} \cdot x - f(x)\ \ \ \text{ for all }\ x \in {\mathbb{R}}^{n}.}$$

Hence,

$$\displaystyle{\xi _{k} \cdot x_{k} - f(x_{k}) = {f}^{\star }(\xi _{ k}).}$$

Using this, we compute that

$$\displaystyle\begin{array}{rcl}{ f}^{\star \star }(a)& \geq & a \cdot \xi _{ k} - {f}^{\star }(\xi _{ k}) =\xi _{k} \cdot a -\xi _{k} \cdot x_{k} + f(x_{k}) {}\\ & =& 2k\vert x_{k} - a{\vert }^{2} + f(x_{ k}). {}\\ \end{array}$$

We divide our argument into the following cases, (a) and (b).

Case (a): \(\lim _{k\rightarrow \infty }k\vert x_{k} - a{\vert }^{2} = \infty \). In this case, if we set \(m =\min _{\bar{B}_{R}(a)}f\), then we have

$$\displaystyle{{f}^{\star \star }(a) \geq \liminf _{ k\rightarrow \infty }2k\vert x_{k} - a{\vert }^{2} + m = \infty,}$$

and, therefore, f  ⋆ ⋆ (a) ≥ f(a).

Case (b): \(\liminf _{k\rightarrow \infty }k\vert x_{k} - a{\vert }^{2} < \infty \). We may choose a subsequence \(\{x_{k_{j}}\}_{j\in \mathbb{N}}\) of {x k } so that \(\lim _{j\rightarrow \infty }x_{k_{j}} = a\). Then we have

$$\displaystyle{{f}^{\star \star }(a) \geq \liminf _{ j\rightarrow \infty }\left (2k_{j}\vert x_{k_{j}} - a{\vert }^{2} + f(x_{ k_{j}})\right ) \geq \liminf _{j\rightarrow \infty }f(x_{k_{j}}) \geq f(a).}$$

Thus, in both cases we have f  ⋆ ⋆ (a) ≥ f(a), which completes the proof. □ 

Theorem B.2.

Let \(f\, :\, {\mathbb{R}}^{n} \rightarrow (-\infty,\,\infty ]\) be proper convex and \(x,\xi \in {\mathbb{R}}^{n}\) . Then the following three conditions are equivalent each other.

  1. (i)

    ξ ∈ D f(x).

  2. (ii)

    \(x \in {D}^{-}{f}^{\star }(\xi )\).

  3. (iii)

    x ⋅ξ = f(x) + f (ξ).

Proof.

Assume first that (i) holds. By Proposition B.1, we have

$$\displaystyle{f(y) \geq f(x) +\xi \cdot (y - x)\ \ \ \text{ for all }\ y \in {\mathbb{R}}^{n},}$$

which reads

$$\displaystyle{\xi \cdot x - f(x) \geq \xi \cdot y - f(y)\ \ \ \text{ for all }y \in {\mathbb{R}}^{n}.}$$

Hence,

$$\displaystyle{\xi \cdot x - f(x) =\max _{y\in {\mathbb{R}}^{n}}(\xi \cdot y - f(y)) = {f}^{\star }(\xi ).}$$

Thus, (iii) is valid.

Next, we assume that (iii) holds. Then the function \(y\mapsto \xi \cdot y - f(y)\) attains a maximum at x. Therefore, ξ ∈ D  −  f(x). That is, (i) is valid.

Now, by the convex duality (Theorem B.1), (iii) reads

$$\displaystyle{x\cdot \xi = {f}^{\star \star }(x) + {f}^{\star }(\xi ).}$$

The equivalence between (i) and (iii), with f replaced by f  ⋆ , is exactly the equivalence between (ii) and (iii). The proof is complete. □ 

Finally, we give a Lipschitz regularity estimate for convex functions.

Theorem B.3.

Let \(f\, :\, {\mathbb{R}}^{n} \rightarrow (-\infty,\,\infty ]\) be a convex function. Assume that there are constants M > 0 and R > 0 such that

$$\displaystyle{\vert f(x)\vert \leq M\ \ \ \text{ for all }\ x \in B_{3R}.}$$

Then

$$\displaystyle{\vert f(x) - f(y)\vert \leq \frac{M} {R} \vert x - y\vert \ \ \ \text{ for all }\ x,y \in B_{R}.}$$

Proof.

Let x, y ∈ B R and note that | x − y |  < 2R. We may assume that x ≠ y. Setting ξ = (x − y) ∕ | x − y | and z = y + 2 and noting that z ∈ B 3R ,

$$\displaystyle{ x - y = \frac{\vert x - y\vert } {2R} (z - y), }$$

and

$$\displaystyle{ x = y + \frac{\vert x - y\vert } {2R} (z - y) = \frac{\vert x - y\vert } {2R} z + \left (1 -\frac{\vert x - y\vert } {2R} \right )y, }$$

we obtain

$$\displaystyle{f(x) \leq \frac{\vert x - y\vert } {2R} f(z) + \left (1 -\frac{\vert x - y\vert } {2R} \right )f(y),}$$

and

$$\displaystyle{f(x) - f(y) \leq \frac{\vert x - y\vert } {2R} (f(z) - f(y)) \leq \frac{\vert x - y\vert } {2R} (\vert f(z)\vert + \vert f(y)\vert ) \leq \frac{M\vert x - y\vert } {R}.}$$

In view of the symmetry in x and y, we see that

$$\displaystyle{\vert f(x) - f(y)\vert \leq \frac{M} {R} \vert x - y\vert \ \ \ \text{ for all }\ x,y \in B_{R}.}$$

 □ 

1.3 A.3 Global Lipschitz Regularity

We give here a proof of Lemmas 2.1 and 2.2.

Proof (Lemma 2.1). 

We first show that there is a constant C > 0, for each \(z \in \overline{\Omega }\) a ball B r (z) centered at z, and for each \(x,y \in B_{r}(z) \cap \overline{\Omega }\), a curve \(\eta \in \mathrm{AC}([0,T], {\mathbb{R}}^{n})\), with \(T \in \overline{\mathbb{R}}_{+}\), such that η(s) ∈ Ω for all s ∈ (0, T), \(\vert \dot{\eta }(s)\vert \leq 1\) for a.e. s ∈ (0, T) and T ≤ C | x − y | .

Let ρ be a defining function of Ω. We may assume that \(\|D\rho \|_{\infty,{\mathbb{R}}^{n}} \leq 1\) and | (x) | ≥ δ for all \(x \in {(\partial \Omega )}^{\delta } :=\{ y \in {\mathbb{R}}^{n}\, :\, \mathrm{dist}(y,\partial \Omega ) <\delta \}\) and some constant δ ∈ (0, 1).

Let z ∈ Ω. We can choose r > 0 so that B r (z) ⊂ Ω. Then, for each x, y ∈ B r (z), with x ≠ y, the line η(s) = x + s(y − x) ∕ | y − x | , with s ∈ [0,  | x − y | ], connects two points x and y and lies inside Ω. Note as well that \(\dot{\eta }(s) = (y - x)/\vert y - x\vert \in \partial B_{1}\) for all s ∈ [0,  | x − y | ].

Let \(z \in \partial \Omega \). Since | (z) | 2 ≥ δ 2, by continuity, we may choose r ∈ (0, δ 3 ∕ 4) so that (x) ⋅(z) ≥ δ 2 ∕ 2 for all \(x \in B_{{4\delta }^{-2}r}(z)\). Fix any \(x,y \in B_{r}(z) \cap \overline{\Omega }\). Consider the curve ξ(t) = x + t(y − x) − t(1 − t)6δ  − 2 | x − y | (z), with t ∈ [0, 1], which connects the points x and y. Note that

$$\displaystyle\begin{array}{rcl} \vert \xi (t) - z\vert & \leq & \,(1 - t)\vert x - z\vert + t\vert y - z\vert + 6t{(1 - t)\delta }^{-2}\vert x - y\vert \vert D\rho (z)\vert {}\\ & <& \,(1 + {3\delta }^{-2})r < {4\delta }^{-2}r {}\\ \end{array}$$

and 4δ  − 2 r < δ. Hence, we have \(\xi (t) \in B_{{4\delta }^{-2}r}(z) \cap {(\partial \Omega )}^{\delta }\) for all t ∈ [0, 1]. If t ∈ (0, 1 ∕ 2], then we have

$$\displaystyle\begin{array}{rcl} \rho (\xi (t))& \leq & \,\rho (x) + tD\rho (\theta \xi (t) + (1-\theta )x) \cdot (y - x - 6{(1 - t)\delta }^{-2}\vert x - y\vert D\rho (z)) {}\\ & \leq & \,t\vert x - y\vert (1 - 3(1 - t)) < 0 {}\\ \end{array}$$

for some θ ∈ (0, 1). Similarly, if t ∈ [1 ∕ 2, 1), we have

$$\displaystyle{\rho (\xi (t)) \leq \rho (y) + (1 - t)\vert x - y\vert (1 - 3t) < 0.}$$

Hence, ξ(t) ∈ Ω for all t ∈ (0, 1). Note that

$$\displaystyle{\vert \dot{\xi }(t)\vert \leq \vert y - x\vert (1 + {6\delta }^{-2}).}$$

If x = y, then we just set η(s) = x = y for s = 0 and the curve \(\eta \,:\, [0,\,0] \rightarrow {\mathbb{R}}^{n}\) has the required properties. Now let x ≠ y. We set t(x, y) = (1 + 6δ  − 2) | x − y | and η(s) = ξ(s ∕ t(x, y)) for s ∈ [0, t(x, y)]. Then the curve \(\eta : [0,\,t(x,y)] \rightarrow {\mathbb{R}}^{n}\) has the required properties with C = 1 + 6δ  − 2.

Thus, by the compactness of \(\overline{\Omega }\), we may choose a constant C > 0 and a finite covering \(\{{B}^{i}\}_{i=1}^{N}\) of \(\overline{\Omega }\) consisting of open balls with the properties: for each \(x,y \in \hat{ B}_{i} \cap \overline{\Omega }\), where \(\hat{B}_{i}\) denotes the concentric open ball of B i with radius twice that of B i , there exists a curve \(\eta \in \mathrm{AC}([0,\,t(x,y)], {\mathbb{R}}^{n})\) such that η(s) ∈ Ω for all s ∈ (0, t(x, y)), \(\vert \dot{\eta }(s)\vert \leq 1\) for a.e. s ∈ [0, t(x, y)] and t(x, y) ≤ C | x − y | .

Let r i be the radius of the ball B i and set r = minr i and R = ∑r i , where i ranges all over i = 1, , N.

Let \(x,y \in \overline{\Omega }\). If | x − y |  < r, then \(x,y \in \hat{ B}_{i}\) for some i and there is a curve \(\eta \in \mathrm{AC}([0,t(x,y)], {\mathbb{R}}^{n})\) such that η(s) ∈ Ω for all s ∈ (0, t(x, y)), \(\vert \dot{\eta }(s)\vert \leq 1\) for a.e. s ∈ [0, t(x, y)] and t(x, y) ≤ C | x − y | . Next, we assume that | x − y | ≥ r. By the connectedness of Ω, we infer that there is a sequence \(\{B_{i_{j}}\, :\, j = 1,\ldots,J\} \subset \{ B_{i}\, :\, i = 1,\ldots,N\}\) such that \(x \in B_{i_{1}}\), \(y \in B_{i_{J}}\), \(B_{i_{j}} \cap B_{i_{j+1}} \cap \Omega \not =\varnothing \) for all 1 ≤ j < J, and \(B_{i_{j}}\not =B_{i_{k}}\) if j ≠ k. It is clear that J ≤ N. If J = 1, then we may choose a curve η with the required properties as in the case where | x − y |  < r. If J > 1, then we may choose a curve \(\eta \in \mathrm{AC}([0,\,t(x,y)],\, {\mathbb{R}}^{n})\) joining x and y as follows. First, we choose a sequence {x j  : j = 1, , J − 1} of points in Ω so that \(x_{j} \in B_{i_{j}} \cap B_{i_{j+1}} \cap \Omega \) for all 1 ≤ j < J. Next, setting x 0 = x, x J  = y and t 0 = 0, since \(x_{j-1},x_{i_{j}} \in B_{j} \cap \overline{\Omega }\) for all 1 ≤ j ≤ J, we may select \(\eta _{j} \in \mathrm{AC}([t_{j-1},\,t_{j}],\, {\mathbb{R}}^{n})\), with 1 ≤ j ≤ J, inductively so that \(\eta _{j}(t_{j-1}) = x_{j-1}\), \(\eta _{j}(t_{j}) = x_{j}\), η j (s) ∈ Ω for all s ∈ (t j − 1, t j ) and \(t_{j} \leq t_{j-1} + C\vert x_{j} - x_{j-1}\vert \). Finally, we define \(\eta \in \mathrm{AC}([0,\,t(x,y)], {\mathbb{R}}^{n})\), with t(x, y) = t J , by setting η(s) = η i (s) for s ∈ [t j  − 1, t j ] and 1 ≤ j ≤ J. Noting that

$$\displaystyle{T \leq C\sum _{j=1}^{J}\vert x_{ j} - x_{j-1}\vert \leq C\sum _{j=1}^{J}r_{ i_{j}} \leq CR \leq CR{r}^{-1}\vert x - y\vert,}$$

we see that the curve \(\eta \in \mathrm{AC}([0,\,t(x,y)],\, {\mathbb{R}}^{n})\) has all the required properties with C replaced by CRr  − 1. □ 

Remark C.1.

(i) A standard argument, different from the above one, to prove the local Lipschitz continuity near the boundary points is to flatten the boundary by a local change of variables. (ii) One can easily modify the above proof to prove the proposition same as Lemma 2.1, except that Ω is a Lipschitz domain.

Proof (Lemma 2.2). 

Let C > 0 be the constant from Lemma 2.1. We show that | u(x) − u(y) | ≤ CM | x − y | for all x, y ∈ Ω.

To show this, we fix any x, y ∈ Ω such that x ≠ y. By Lemma 2.1, there is a curve \(\eta \in \mathrm{AC}([0,\,t(x,y)],\, {\mathbb{R}}^{n})\) such that  η(0) = x,  η(t(x, y)) = y,  t(x, y) ≤ C | x − y | ,  η(s) ∈ Ω  for all s ∈ [0, t(x, y)]  and  \(\vert \dot{\eta }(s)\vert \leq 1\)  for a.e. s ∈ [0, t(x, y)].

By the compactness of the image η([0, t(x, y)]) of interval [0, t(x, y)] by η, we may choose a finite sequence \(\{B_{i}\}_{i=1}^{N}\) of open balls contained in Ω which covers η([0, t(x, y)]). We may assume by rearranging the label i if needed that x ∈ B 1, y ∈ B N and \(B_{i} \cap B_{i+1}\not =\varnothing \) for all 1 ≤ i < N. We may choose a sequence \(0 = t_{0} < t_{1} < \cdots < t_{N} = t(x,y)\) of real numbers so that the line segment \([\eta (t_{i-1}),\,\eta (t_{i})]\) joining η(t i − 1) and η(t i ) lies in B i for any i = 1, , N.

Thanks to Proposition 1.14, we have

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

Using this, we compute that

$$\displaystyle\begin{array}{rcl} \vert u(y) - u(x)\vert & =& \,\vert u(\eta (t_{N})) - u(\eta (t_{0}))\vert \leq \sum _{i=1}^{N}\vert u(\eta (t_{ i})) - u(\eta (t_{i-1}))\vert {}\\ &\leq & \,M\sum _{i=1}^{N}\vert \eta (t_{ i}) -\eta (t_{i-1})\vert \leq M\sum _{i=1}^{N}\int _{ t_{i-1}}^{t_{i} }\vert \dot{\eta }(s)\vert \mathrm{d}s {}\\ & =& \,M\int _{t_{0}}^{t_{N} }\vert \dot{\eta }(s)\vert \mathrm{d}s \leq M(t_{N} - t_{0}) = \mathit{Mt}(x,y) \leq CM\vert x - y\vert. {}\\ \end{array}$$

This completes the proof. □ 

1.4 A.4 Localized Versions of Lemma 4.2

Theorem D.1.

Let U, V be open subsets of \({\mathbb{R}}^{n}\) with the properties: \(\overline{V } \subset U\) and \(V \cap \Omega \not =\varnothing \) . Let \(u \in C(U \cap \overline{\Omega })\) be a viscosity solution of

$$\displaystyle{ \left \{\begin{array}{llll} &H(x,\mathit{Du}(x)) \leq 0&&\mathrm{in}\ U \cap \Omega, \\ &\frac{\partial u} {\partial \gamma } (x) \leq g(x) &&\mathrm{on}\ U \cap \partial \Omega. \end{array} \right. }$$
(131)

Then, for each \(\varepsilon \in (0,\,1)\) , there exists a function \({u}^{\varepsilon } \in {C}^{1}(V \cap \overline{\Omega })\) such that

$$\displaystyle{ \left \{\begin{array}{llll} &H(x,{\mathit{Du}}^{\varepsilon }(x)) \leq \varepsilon &&\mathrm{in}\ V \cap \Omega, \\ &\frac{\partial {u}^{\varepsilon }} {\partial \gamma } (x) \leq g(x) &&\mathrm{on}\ V \cap \partial \Omega,\\ &\|{u}^{\varepsilon } - u\|_{ \infty,V \cap \Omega } \leq \varepsilon. \end{array} \right. }$$

Proof.

We choose functions \(\zeta,\,\eta \in {C}^{1}({\mathbb{R}}^{n})\) so that 0 ≤ ζ(x) ≤ η(x) ≤ 1 for all \(x \in {\mathbb{R}}^{n}\), ζ(x) = 1 for all x ∈ V, η(x) = 1 for all \(x \in \mathrm{supp}\zeta\) and \(\mathrm{supp}\eta \subset U\).

We define the function \(v \in C(\overline{\Omega })\) by setting v(x) = η(x)u(x) for \(x \in U \cap \overline{\Omega }\) and v(x) = 0 otherwise. By the coercivity of H, u is locally Lipschitz continuous in \(U \cap \overline{\Omega }\), and hence, v is Lipschitz continuous in \(\overline{\Omega }\). Let L > 0 be a Lipschitz bound of v in \(\overline{\Omega }\). Then v is a viscosity solution of

$$\displaystyle{ \left \{\begin{array}{llll} &\vert Dv(x)\vert \leq L&&\mathrm{in}\ \Omega, \\ &\frac{\partial v} {\partial \gamma } (x) \leq M &&\mathrm{in}\ \ \partial \Omega, \end{array} \right. }$$

where \(M := L\|\gamma \|_{\infty,\partial \Omega }\). In fact, we have a stronger assertion that for any \(x \in \overline{\Omega }\) and any p ∈ D  +  v(x),

$$\displaystyle{ \left \{\begin{array}{llll} &\vert p\vert \leq L &&\mathrm{if}\ \ x \in \Omega,\\ &\gamma (x) \cdot p \leq M&&\mathrm{if}\ \ x \in \partial \Omega. \end{array} \right. }$$
(132)

To check this, let \(\phi \in {C}^{1}(\overline{\Omega })\) and assume that v − ϕ attains a maximum at \(x \in \overline{\Omega }\). Observe that if x ∈ Ω, then | (x) | ≤ L and that if x ∈ ∂Ω, then

$$\displaystyle\begin{array}{rcl} 0& \leq & \,\liminf _{t\rightarrow 0+}\frac{(v-\phi )(x - t\gamma (x)) - (v-\phi )(x)} {-t} {}\\ & =& \,\liminf _{t\rightarrow 0+}\frac{v(x - t\gamma (x)) - v(x)} {-t} -\frac{\partial \phi } {\partial \gamma }(x), {}\\ \end{array}$$

which yields

$$\displaystyle{\gamma (x) \cdot D\phi (x) \leq L\vert \gamma (x)\vert \leq M.}$$

Thus, (132) is valid.

We set

$$\displaystyle\begin{array}{rcl} h(x)& =& \,\zeta (x)g(x) + (1 -\zeta (x))M\ \ \text{ for }\ x \in \partial \Omega, {}\\ G(x,p)& =& \,\zeta (x)H(x,p) + (1 -\zeta (x))(\vert p\vert - L)\ \ \text{ for }\ (x,p) \in \overline{\Omega } \times {\mathbb{R}}^{n}. {}\\ \end{array}$$

It is clear that h ∈ C(∂Ω) and G satisfies (A5)–(A7), with H replaced by G

In view of the coercivity of H, we may assume by reselecting L if necessary that for all \((x,p) \in \overline{\Omega } \times {\mathbb{R}}^{n}\), if | p |  > L, then H(x, p) > 0. We now show that v is a viscosity solution of

$$\displaystyle{ \left \{\begin{array}{llll} &G(x,Dv(x)) \leq 0&&\mathrm{in}\ \Omega, \\ &\frac{\partial v} {\partial \gamma } (x) \leq h(x) &&\mathrm{on}\ \partial \Omega. \end{array} \right. }$$
(133)

To do this, let \(\hat{x} \in \overline{\Omega }\) and \(\hat{p} \in {D}^{+}v(\hat{x})\). Consider the case where \(\zeta (\hat{x}) > 0\), which implies that \(\hat{x} \in U\). We have η(x) = 1 near the point \(\hat{x}\), which implies that \(\hat{p} \in {D}^{+}u(\hat{x})\). As u is a viscosity subsolution of (131), we have  \(H(\hat{x},\hat{p}) \leq 0\) if \(\hat{x} \in \Omega \) and \(\min \{H(\hat{x},\hat{p}),\,\gamma (\hat{x}) \cdot \hat{ p} - h(\hat{x})\} \leq 0\) if \(\hat{x} \in \partial \Omega \). Assume in addition that \(\hat{x} \in \partial \Omega \). By (132), we have \(\gamma (\hat{x}) \cdot \hat{ p} \leq M\). If \(\vert \hat{p}\vert > L\), we have both

$$\displaystyle{\gamma (\hat{x}) \cdot \hat{ p} \leq g(\hat{x})\ \ \text{ and }\ \ \gamma (\hat{x}) \cdot \hat{ p} \leq M.}$$

Hence, if \(\vert \hat{p}\vert > L\), then \(\gamma (\hat{x}) \cdot \hat{ p} \leq h(\hat{x})\). On the other hand, if \(\vert \hat{p}\vert \leq L\), we have two cases: in one case we have \(H(\hat{x},\hat{p}) \leq 0\) and hence, \(G(\hat{x},\hat{p}) \leq 0\). In the other case, we have \(\gamma (\hat{x}) \cdot \hat{ p} \leq g(\hat{x})\) and then \(\gamma (\hat{x}) \cdot \hat{ p} \leq h(\hat{x})\). These observations together show that

$$\displaystyle{\min \{G(\hat{x},\hat{p}),\,\gamma (\hat{x}) \cdot \hat{ p} - h(\hat{x})\} \leq 0.}$$

We next assume that \(\hat{x} \in \Omega \). In this case, we easily see that \(G(\hat{x},\hat{p}) \leq 0\).

Next, consider the case where \(\zeta (\hat{x}) = 0\), which implies that \(G(\hat{x},\hat{p}) = \vert \hat{p}\vert - L\) and \(h(\hat{x}) = M\). By (132), we immediately see that \(G(\hat{x},\hat{p}) \leq 0\) if \(\hat{x} \in \Omega \) and \(\min \{G(\hat{x},\hat{p}),\gamma (\hat{x}) \cdot \hat{ p} - h(\hat{x})\} \leq 0\) if \(\hat{x} \in \partial \Omega \). We thus conclude that v is a viscosity solution of (133).

We may invoke Theorem 4.2, to find a collection \(\{{v}^{\varepsilon }\}_{\varepsilon \in (0,1)} \subset {C}^{1}(\overline{\Omega })\) such that

$$\displaystyle{\left \{\begin{array}{llll} &G(x,D{v}^{\varepsilon }(x)) \leq \varepsilon &&\text{ for all }x \in \Omega, \\ &\frac{\partial {v}^{\varepsilon }} {\partial \gamma } (x) \leq h(x) &&\text{ for all }x \in \partial \Omega,\\ &\|{v}^{\varepsilon } - v\|_{ \infty,\Omega } \leq \varepsilon. \end{array} \right.}$$

But, this yields

$$\displaystyle{\left \{\begin{array}{llll} &H(x,{v}^{\varepsilon }(x)) \leq \varepsilon &&\text{ for all }\ x \in V \cap \Omega, \\ &\frac{\partial {v}^{\varepsilon }} {\partial \gamma } (x) \leq g(x) &&\text{ for all }\ x \in V \cap \partial \Omega,\\ &\|{v}^{\varepsilon } - u\|_{ \infty,V \cap \Omega } \leq \varepsilon. \end{array} \right.}$$

The functions \({v}^{\varepsilon }\) have all the required properties. □ 

The above theorem has a version for Hamilton–Jacobi equations of evolution type.

Theorem D.2.

Let U, V be bounded open subsets of \({\mathbb{R}}^{n} \times \mathbb{R}_{+}\) with the properties:   \(\overline{V } \subset U\) \(\overline{U} \subset {\mathbb{R}}^{n} \times \mathbb{R}_{+}\)  and   \(V \cap Q\not =\varnothing \) . Let \(u \in \mathrm{Lip}(U \cap Q)\) be a viscosity solution of

$$\displaystyle{ \left \{\begin{array}{llll} &u_{t}(x,t) + H(x,D_{x}u(x,t)) \leq 0&&\mathrm{in}\ U \cap (\Omega \times \mathbb{R}_{+}), \\ &\frac{\partial u} {\partial \gamma } (x,t) \leq g(x) &&\mathrm{on}\ U \cap (\partial \Omega \times \mathbb{R}_{+}). \end{array} \right. }$$

Then, for each \(\varepsilon \in (0,\,1)\) , there exists a function \({u}^{\varepsilon } \in {C}^{1}(V \cap Q)\) such that

$$\displaystyle{ \left \{\begin{array}{llll} &u_{t}^{\varepsilon }(x,t) + H(x,D_{x}{u}^{\varepsilon }(x,t)) \leq \varepsilon &&\mathrm{in}\ V \cap (\Omega \times \mathbb{R}_{+}), \\ &\frac{\partial {u}^{\varepsilon }} {\partial \gamma } (x,t) \leq g(x) &&\mathrm{on}\ V \cap (\partial \Omega \times \mathbb{R}_{+}),\\ &\|{u}^{\varepsilon } - u\|_{ \infty,V \cap Q} \leq \varepsilon. \end{array} \right. }$$
(134)

Proof.

Choose constants \(a,b \in \mathbb{R}_{+}\) so that \(U \subset {\mathbb{R}}^{n} \times (a,b)\) and let ρ be a defining function of Ω. We may assume that ρ is bounded in \({\mathbb{R}}^{n}\). We choose a function \(\zeta \in {C}^{1}(\mathbb{R})\) so that ζ(t) = 0 for all t ∈ [a, b], ζ′(t) > 0 for all t > b, ζ′(t) < 0 for all t < a and \(\min \{\zeta (a/2),\zeta (2b)\} >\|\rho \| _{\infty,\Omega }\).

We set

$$\displaystyle\begin{array}{rcl} \tilde{\rho }(x,t)& =& \,\rho (x) +\zeta (t)\ \ \text{ for }\ (x,t) \in {\mathbb{R}}^{n+1}, {}\\ \tilde{\Omega }& =& \,\{(x,t) \in {\mathbb{R}}^{n+1}\, :\,\tilde{\rho } (x,t) < 0\}. {}\\ \end{array}$$

It is easily seen that

$$\displaystyle{\tilde{\Omega } \subset \Omega \times (a/2,\,2b)\ \ \ \text{ and }\ \ \ \tilde{\Omega } \cap \left ({\mathbb{R}}^{n} \times [a,\,b]\right ) = \Omega \times [a,\,b].}$$

Let \((x,t) \in {\mathbb{R}}^{n+1}\) be such that \(\tilde{\rho }(x,t) = 0\). It is obvious that \((x,t) \in \overline{\Omega } \times [a/2,\,2b]\). If a ≤ t ≤ b, then ρ(x) = 0 and thus (x) ≠ 0. If either t > b or t < a, then | ζ′(t) |  > 0. Hence, we have \(D\tilde{\rho }(x,t)\not =0\). Thus, \(\tilde{\rho }\) is a defining function of \(\tilde{\Omega }\).

Let M > 0 and define \(\tilde{\gamma }\in C(\partial \tilde{\Omega }, {\mathbb{R}}^{n+1})\) by

$$\displaystyle{\tilde{\gamma }(x,t) = \left ((1 + M\rho (x))_{+}\gamma (x),\,\zeta ^{\prime}(t)\right ),}$$

where we may assume that γ is defined and continuous in \(\overline{\Omega }\). We note that for any \((x,t) \in \partial \tilde{\Omega }\),

$$\displaystyle{\tilde{\gamma }(x,t) \cdot D\tilde{\rho }(x,t) = (1 + M\rho (x))_{+}\gamma (x) \cdot D\rho (x) +\zeta ^{\prime}{(t)}^{2}.}$$

Note as well that (1 + (x)) +  = 1 for all x ∈ ∂Ω and

$$\displaystyle{\lim _{M\rightarrow \infty }(1 + M\rho (x))_{+} = 0\ \ \ \text{ locally uniformly in }\Omega.}$$

Thus we can fix M > 0 so that for all \((x,t) \in \partial \tilde{\Omega }\),

$$\displaystyle{\tilde{\gamma }(x,t) \cdot D\tilde{\rho }(x,t) = (1 + M\rho (x))_{+}\gamma (x) \cdot D\rho (x) +\zeta ^{\prime}{(t)}^{2} > 0.}$$

Noting that for each x ∈ Ω, the x-section \(\{t \in \mathbb{R}\, :\, (x,t) \in \tilde{ \Omega }\}\) of \(\tilde{\Omega }\) is an open interval (or, line segment), we deduce that \(\tilde{\Omega }\) is a connected set. We may assume that g is defined and continuous in \(\overline{\Omega }\). We define \(\tilde{g} \in C(\partial \tilde{\Omega })\) by \(\tilde{g}(x,t) = g(x)\). Thus, assumptions (A1)–(A4) hold with n + 1, \(\tilde{\Omega }\), \(\tilde{\gamma }\) and \(\tilde{g}\) in place of n, Ω, γ and g.

Let L > 0 be a Lipschitz bound of the function u in \(U \cap Q\). Set

$$\displaystyle{\tilde{H}(x,t,p,q) = H(x,p) + q + 2(\vert q\vert - L)_{+}\ \ \ \text{ for }(x,t,p,q) \in \overline{\tilde{\Omega }\,} \times {\mathbb{R}}^{n+1},}$$

and note that \(\tilde{H} \in C(\overline{\tilde{\Omega }\,} \times {\mathbb{R}}^{n+1})\) satisfies (A5)–(A7), with Ω replaced by \(\tilde{\Omega }\).

We now claim that u is a viscosity solution of

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

Indeed, since \(U \cap \tilde{ \Omega } = U \cap Q\) and \(U \cap \partial \tilde{\Omega } = U \cap \partial Q\), if \((x,t) \in U \cap \overline{\tilde{\Omega }}\) and (p, q) ∈ D  +  u(x, t), then we get | q | ≤ L by the cylindrical geometry of Q and, by the viscosity property of u,

$$\displaystyle{\left \{\begin{array}{llll} &q + H(x,p) + 2(\vert q\vert - L)_{+} \leq 0 &&\mathrm{if}\ \ (x,t) \in \tilde{ \Omega }, \\ &\min \{q + H(x,p) + 2(\vert q\vert - L)_{+},\gamma (x) \cdot p - g(x)\} \leq 0&&\mathrm{if}\ (x,t) \in \partial \tilde{\Omega }.\end{array} \right.}$$

We apply Theorem D.1, to find a collection \(\{{u}^{\varepsilon }\}_{\varepsilon \in (0,1)} \subset {C}^{1}(V \cap \overline{\tilde{\Omega }\,})\) such that

$$\displaystyle{\left \{\begin{array}{llll} &\tilde{H}(x,t,{\mathit{Du}}^{\varepsilon }(x,t)) \leq \varepsilon &&\mathrm{in}\ V \cap \tilde{ \Omega }, \\ &\tilde{\gamma }(x,t) \cdot {\mathit{Du}}^{\varepsilon }(x,t) \leq \tilde{ g}(x,t)&&\mathrm{on}\ U \cap \tilde{ \Omega }, \\ &\|{u}^{\varepsilon } - u\|_{\infty,V \cap \tilde{\Omega }} \leq \varepsilon.\end{array} \right.}$$

It is straightforward to see that the collection \(\{{u}^{\varepsilon }\}_{\varepsilon \in (0,1)} \subset {C}^{1}(V \cap Q)\) satisfies (134). □ 

1.5 A.5 A Proof of Lemma 5.4

This subsection is mostly devoted to the proof of Lemma 5.4, a version of the Dunford–Pettis theorem. We also give a proof of the weak-star compactness of bounded sequences in \({L}^{\infty }(J, {\mathbb{R}}^{m})\), where J = [a, b] is a finite interval in \(\mathbb{R}\).

Proof (Lemma 5.4). 

We define the functions \(F_{j} \in C(J, {\mathbb{R}}^{m})\) by

$$\displaystyle{F_{j}(x) =\int _{ a}^{x}f_{ j}(t)\mathrm{d}t.}$$

By the uniform integrability of {f j }, the sequence \(\{F_{j}\}_{j\in \mathbb{N}}\) is uniformly bounded and equi-continuous in J. Hence, the Ascoli–Arzela theorem ensures that it has a subsequence converging to a function F uniformly in J. We fix such a subsequence and denote it again by the same symbol {F j }. Because of the uniform integrability assumption, the sequence {F j } is equi-absolutely continuous in J. That is, for any \(\varepsilon > 0\) there exists δ > 0 such that

$$\displaystyle\begin{array}{rcl} & & a \leq a_{1} < b_{1} < a_{2} < b_{2} < \cdots < a_{n} < b_{n} \leq b,\ \quad \ \sum _{i=1}^{n}(b_{ i} - a_{i}) <\delta,\ {}\\ & & \quad \Rightarrow\ \sum _{i=1}^{n}\vert f_{ j}(b_{i}) - f_{j}(a_{i})\vert <\varepsilon \ \ \ \text{ for all }\ j \in \mathbb{N}. {}\\ \end{array}$$

An immediate consequence of this is that \(F \in \mathrm{AC}(J, {\mathbb{R}}^{m})\). Hence, for some \(f \in {L}^{1}(J, {\mathbb{R}}^{m})\), we have

$$\displaystyle{F(x) =\int _{ a}^{x}f(t)\,\mathrm{d}t\ \ \ \text{ for all }\ x \in J.}$$

Next, let ϕ ∈ C 1(J), and we show that

$$\displaystyle{ \lim _{j\rightarrow \infty }\int _{a}^{b}f_{ j}(x)\phi (x)\,\mathrm{d}x =\int _{ a}^{b}f(x)\phi (x)\,\mathrm{d}x. }$$
(135)

Integrating by parts, we observe that as j → ,

$$\displaystyle\begin{array}{rcl} \int _{a}^{b}f_{ j}(x)\phi (x)\,\mathrm{d}x& & \,=\big [F_{j}\phi \big]_{a}^{b} -\int _{ a}^{b}F_{ j}(x)\phi ^{\prime}(x)\,\mathrm{d}x {}\\ & & \,\rightarrow \big [F\phi \big]_{a}^{b} -\int _{ a}^{b}F(x)\phi ^{\prime}(x)\,\mathrm{d}x =\int _{ a}^{b}f(x)\phi (x)\,\mathrm{d}x. {}\\ \end{array}$$

Hence, (135) is valid.

Now, let ϕ ∈ L (J). We regard the functions f j , f, ϕ as functions defined in \(\mathbb{R}\) by setting f j (x) = f(x) = ϕ(x) = 0 for x < a or x > b. Let \(\{k_{\varepsilon }\}_{\varepsilon >0}\) be a collection of standard mollification kernels. We recall that

$$\displaystyle{ \lim _{\varepsilon \rightarrow 0}\|k_{\varepsilon } {\ast}\phi -\phi \|_{{L}^{1}(J)} = 0, }$$
(136)
$$\displaystyle{ \vert k_{\varepsilon } {\ast}\phi (x)\vert \leq \|\phi \|_{{L}^{\infty }(J)}\ \ \ \text{ for all }\ x \in J,\ \varepsilon > 0. }$$
(137)

Fix any δ > 0. By the uniform integrability assumption, we have

$$\displaystyle{M :=\sup _{j\in \mathbb{N}}\|f_{j} - f\|_{{L}^{1}(J)} < \infty.}$$

Let α > 0 and set

$$\displaystyle{E_{j} :=\{ x \in J\, :\, \vert (f_{j} - f)(x)\vert >\alpha \}.}$$

By the Chebychev inequality, we get

$$\displaystyle{\vert E_{j}\vert \leq \frac{M} {\alpha }.}$$

By the uniform integrability assumption, if α > 0 is sufficiently large, then

$$\displaystyle{ \int _{E_{j}}\vert (f_{j} - f)(x)\vert \,\mathrm{d}x <\delta. }$$
(138)

In what follows we fix α > 0 large enough so that (138) holds. We write f j  − f = g j  + b j , where  \(g_{j} = (f_{j} - f)(1 -\mathbf{1}_{E_{j}})\)  and  \(b_{j} = (f_{j} - f)\mathbf{1}_{E_{j}}\). Then,

$$\displaystyle{\vert g_{j}(x)\vert \leq \alpha \ \ \ \text{ for all }\ x \in J\quad \text{ and }\quad \|b_{j}\|_{{L}^{1}(J)} <\delta.}$$

Observe that

$$\displaystyle\begin{array}{rcl} I_{j} :& =& \,\int _{J}f_{j}(x)\phi (x)\,\mathrm{d}x -\int _{J}f(x)\phi (x)\,\mathrm{d}x {}\\ & =& \,\int _{J}(f_{j} - f)(x)\,k_{\varepsilon } {\ast}\phi (x)\,\mathrm{d}x +\int _{J}(f_{j} - f)(x)(\phi -k_{\varepsilon }{\ast}\phi )(x)\,\mathrm{d}x {}\\ \end{array}$$

and

$$\displaystyle\begin{array}{rcl} & & \Big\vert \int _{J}(f_{j} - f)(x)(\phi -k_{\varepsilon }{\ast}\phi )(x)\,\mathrm{d}x\Big\vert {}\\ & &\quad \leq \Big\vert \int _{J}g_{j}(x)(\phi -k_{\varepsilon }{\ast}\phi )(x)\,\mathrm{d}x\Big\vert +\Big \vert \int _{J}b_{j}(x)(\phi -k_{\varepsilon }{\ast}\phi )(x)\,\mathrm{d}x\Big\vert {}\\ & &\quad \leq \alpha \| k_{\varepsilon } {\ast}\phi -\phi \|_{{L}^{1}(J)} + 2\delta \|\phi \|_{{L}^{\infty }(J)}. {}\\ \end{array}$$

Hence, in view of (135) and (136), we get  \(\limsup _{j\rightarrow \infty }\vert I_{j}\vert \leq 2\delta \|\phi \|_{{L}^{\infty }(J)}.\) As δ > 0 is arbitrary, we get  \(\lim _{j\rightarrow \infty }I_{j} = 0,\) which completes the proof. □ 

As a corollary of Lemma 5.4, we deduce that the weak-star compactness of bounded sequences in \({L}^{\infty }(J, {\mathbb{R}}^{m})\):

Lemma E.1.

Let J = [a, b], with −∞ < a < b < ∞. Let \(\{f_{k}\}_{k\in \mathbb{N}}\) be a bounded sequence of functions in \({L}^{\infty }(J, {\mathbb{R}}^{m})\) . Then {f k } has a subsequence which converges weakly-star in \({L}^{\infty }(J, {\mathbb{R}}^{m})\).

Proof.

Set \(M =\sup _{k\in \mathbb{N}}\|f_{k}\|_{{L}^{\infty }(J)}\). Let E ⊂ J be a measurable set, and observe that

$$\displaystyle{\int _{E}\vert f_{k}(t)\vert \mathrm{d}t \leq M\vert E\vert \ \ \ \text{ for all }k \in \mathbb{N},}$$

which shows that the sequence {f k } is uniformly integrable in J. Thanks to Lemma 5.4, there exists a subsequence \(\{f_{k_{j}}\}_{j\in \mathbb{N}}\) of {f k } which converges to a function f weakly in L 1(J, R m).

Let \(i \in \mathbb{N}\) and set E i  = { t ∈ J :  | f(t) |  > M + 1 ∕ i} and \(g_{i}(t) = \mathbf{1}_{E_{i}}(t)f(t)/\vert f(t)\vert \) for t ∈ J. Since \(g_{i} \in {L}^{\infty }(J, {\mathbb{R}}^{m})\), we get

$$\displaystyle{\int _{J}f_{k_{j}}(t) \cdot g_{i}(t)\mathrm{d}t \rightarrow \int _{J}\vert f(t)\vert \mathbf{1}_{E_{i}}(t)\mathrm{d}t\ \ \text{ as }\ j \rightarrow \infty.}$$

Hence, using the Chebychev inequality, we obtain

$$\displaystyle{\Big(M + \frac{1} {i} \big)\vert E_{i}\vert \leq \int _{J}\vert f(t)\vert \mathbf{1}_{E_{i}}(t)\mathrm{d}t \leq \int _{J}M\mathbf{1}_{E_{i}}(t)\mathrm{d}t = M\vert E_{i}\vert,}$$

which ensures that | E i  |  = 0. Thus, we find that | f(t) | ≤ M a.e. in J.

Now, fix any \(\phi \in {L}^{1}(J, {\mathbb{R}}^{m})\). We select a sequence \(\{\phi _{i}\}_{i\in \mathbb{N}} \subset {L}^{\infty }(J, {\mathbb{R}}^{m})\) so that, as i → , ϕ i  → ϕ in \({L}^{1}(J, {\mathbb{R}}^{m})\). For each \(i \in \mathbb{N}\), we have

$$\displaystyle{\lim _{j\rightarrow \infty }\int _{J}f_{k_{j}}(t) \cdot \phi _{i}(t)\mathrm{d}t =\int _{J}f(t) \cdot \phi _{i}(t)\mathrm{d}t.}$$

On the other hand, we have

$$\displaystyle{\Big\vert \int _{J}f_{k_{j}}(t) \cdot \phi (t)\mathrm{d}t -\int _{J}f_{k_{j}}(t) \cdot \phi _{i}(t)\mathrm{d}t\Big\vert \leq M\|\phi -\phi _{i}\|_{{L}^{1}(J)}\ \ \ \text{ for all }\ j \in \mathbb{N}}$$

and

$$\displaystyle{\Big\vert \int _{J}f(t) \cdot \phi (t)\mathrm{d}t -\int _{J}f(t) \cdot \phi _{i}(t)\mathrm{d}t\Big\vert \leq M\|\phi -\phi _{i}\|_{{L}^{1}(J)}.}$$

These together yield

$$\displaystyle{\lim _{j\rightarrow \infty }\int _{J}f_{k_{j}}(t) \cdot \phi (t)\mathrm{d}t =\int _{J}f(t) \cdot \phi (t)\mathrm{d}t.}$$

 □ 

1.6 A.6 Rademacher’s Theorem

We give here a proof of Rademacher’s theorem.

Theorem F.1 (Rademacher). 

Let \(B = B_{1} \subset {\mathbb{R}}^{n}\) and \(f \in \mathrm{Lip}(B)\) . Then f is differentiable almost everywhere in B.

To prove the above theorem, we mainly follow the proof given in [1].

Proof.

We first show that f has a distributional gradient Df ∈ L (B).

Let L > 0 be a Lipschitz bound of the function f. Let i ∈ { 1, 2, , n} and e i denote the unit vector in \({\mathbb{R}}^{n}\) with unity as the i-th entry. Fix any ϕ ∈ C 0 1(B) and observe that

$$\displaystyle\begin{array}{rcl} \int _{B}f(x)\phi _{x_{i}}(x)\mathrm{d}x& =& \,\lim _{r\rightarrow 0+}\int _{B}f(x)\frac{\phi (x + re_{i}) -\phi (x)} {r} \mathrm{d}x {}\\ & =& \,\lim _{r\rightarrow 0+}\int _{B}\frac{f(x - re_{i}) - f(x)} {r} \phi (x)\mathrm{d}x {}\\ \end{array}$$

and

$$\displaystyle{\Big\vert \int _{B}f(x)\phi _{x_{i}}(x)\mathrm{d}x\Big\vert \leq L\int _{B}\vert \phi (x)\vert \mathrm{d}x \leq L\vert B{\vert }^{1/2}\|\phi \|_{{ L}^{2}(B)}.}$$

Thus, the map

$$\displaystyle{C_{0}^{1}(B) \ni \phi \mapsto -\int _{ B}f(x)\phi _{x_{i}}(x)\mathrm{d}x \in \mathbb{R}}$$

extends uniquely to a bounded linear functional G i on L 2(B). By the Riesz representation theorem, there is a function g i  ∈ L 2(B) such that

$$\displaystyle{G_{i}(\phi ) =\int _{B}g_{i}(x)\phi (x)\mathrm{d}x\ \ \ \text{ for all }\ \phi \in {L}^{2}(B).}$$

This shows that g = (g 1, , g n ) is the distributional gradient of f.

We plug the function ϕ ∈ L 2(B) given by \(\phi (x) = (g_{i}(x)/\vert g_{i}(x)\vert )\mathbf{1}_{E_{k}}(x)\), where \(k \in \mathbb{N}\) and E k  = { x ∈ B :  | g i (x) |  > L + 1 ∕ k}, into the inequality \(\vert G_{i}(\phi )\vert \leq L\|\phi \|_{{L}^{1}(B)}\), to obtain

$$\displaystyle{\int _{B}\vert g_{i}(x)\vert \mathbf{1}_{E_{k}}(x)\mathrm{d}x \leq L\int _{B}\mathbf{1}_{E_{k}}(x)\mathrm{d}x = L\vert E_{k}\vert,}$$

which yields

$$\displaystyle{(L + 1/k)\vert E_{k}\vert \leq L\vert E_{k}\vert.}$$

Hence, we get | E k  |  = 0 for all \(k \in \mathbb{N}\) and | {x ∈ B :  | g i (x) |  > L} |  = 0. That is, g i  ∈ L (B) and | g i (x) | ≤ L a.e. in B.

The Lebesgue differentiation theorem (see [57]) states that for a.e. x ∈ B, we have \(g(x) \in {\mathbb{R}}^{n}\) and

$$\displaystyle{ \lim _{r\rightarrow 0+} \frac{1} {{r}^{n}}\int _{B_{r}}\vert g(x + y) - g(x)\vert \mathrm{d}y = 0. }$$
(139)

Now, we fix such a point x ∈ B and show that f is differentiable at x. Fix an r > 0 so that B r (x) ⊂ B. For δ ∈ (0, r), consider the function \(h_{\delta } \in C(\overline{B})\) given by

$$\displaystyle{h_{\delta }(y) = \frac{f(x +\delta y) - f(x)} {\delta }.}$$

We claim that

$$\displaystyle{ \lim _{\delta \rightarrow 0}h_{\delta }(y) = g(x) \cdot y\ \ \ \text{ uniformly for }\ y \in \overline{B}. }$$
(140)

Note that h δ (0) = 0 and h δ is Lipschitz continuous with Lipschitz bound L. By the Ascoli–Arzela theorem, for any sequence {δ j } ⊂ (0, r) converging to zero, there exist a subsequence \(\{\delta _{j_{k}}\}_{k\in \mathbb{N}}\) of {δ j } and a function \(h_{0} \in C(\overline{B})\) such that

$$\displaystyle{\lim _{k\rightarrow \infty }h_{\delta _{j_{ k}}}(x) = h_{0}(y)\ \ \ \text{ uniformly for }\ y \in \overline{B}.}$$

In order to prove (140), we need only to show that h 0(y) = g(x) ⋅y for all y ∈ B.

Since h δ (0) = 0 for all δ ∈ (0, r), we have h 0(0) = 0. We observe from (139) that

$$\displaystyle{\int _{B}\vert g(x +\delta y) - g(x)\vert \mathrm{d}y =\int _{B_{\delta }}\vert g(x + y) - g(x){\vert \delta }^{-n}\mathrm{d}y\ \rightarrow \ 0\ \ \text{ as }\ \delta \rightarrow 0.}$$

Using this, we compute that for all ϕ ∈ C 0 1(B),

$$\displaystyle\begin{array}{rcl} \int _{B}h_{0}(y)\phi _{y_{i}}(y)\mathrm{d}y& =& \,\lim _{k\rightarrow \infty }\int _{B}h_{\delta _{j_{ k}}}(y)\phi _{y_{i}}(y)\mathrm{d}y {}\\ & =& \,-\lim _{k\rightarrow \infty }\int _{B}g_{i}(x +\delta _{j_{k}}y)\phi (y)\mathrm{d}y {}\\ & =& \,-\int _{B}g_{i}(x)\phi (y)\mathrm{d}y =\int _{B}g(x) \cdot y\phi _{y_{i}}(y)\mathrm{d}y. {}\\ \end{array}$$

This guarantees that h 0(y) − g(x) ⋅y is constant for all y ∈ B while h 0(0) = 0. Thus, we see that h 0(y) = g(x) ⋅y for all y ∈ B, which proves (140).

Finally, we note that (140) yields

$$\displaystyle{f(x + y) = f(x) + g(x) \cdot y + o(\vert y\vert )\ \ \ \text{ as }\ y \rightarrow 0.}$$

 □ 

Rights and permissions

Reprints and permissions

Copyright information

© 2013 Springer-Verlag Berlin Heidelberg

About this chapter

Cite this chapter

Ishii, H. (2013). A Short Introduction to Viscosity Solutions and the Large Time Behavior of Solutions of Hamilton–Jacobi Equations. In: Hamilton-Jacobi Equations: Approximations, Numerical Analysis and Applications. Lecture Notes in Mathematics(), vol 2074. Springer, Berlin, Heidelberg. https://doi.org/10.1007/978-3-642-36433-4_3

Download citation

Publish with us

Policies and ethics