Skip to main content
Log in

Local strict singular characteristics II: existence for stationary equations on \({\mathbb {R}}^2\)

  • Published:
Nonlinear Differential Equations and Applications NoDEA Aims and scope Submit manuscript

A Correction to this article was published on 07 September 2023

This article has been updated

Abstract

The notion of strict singular characteristics is important in the wellposedness issue of singular dynamics on the cut locus of the viscosity solutions. We provide an intuitive and rigorous proof of the existence of the strict singular characteristics of Hamilton–Jacobi equation \(H(x,Du(x),u(x))=0\) in two dimensional case. We also proved if \({\textbf{x}}\) is a strict singular characteristic, then we really have the right-differentiability of \({\textbf{x}}\) and the right-continuity of \(\dot{{\textbf{x}}}^+(t)\) for every t. Such a strict singular characteristic must give a selection \(p(t)\in D^+u({\textbf{x}}(t))\) such that \(p(t)=\arg \min _{p\in D^+u({\textbf{x}}(t))}H({\textbf{x}}(t),p,u({\textbf{x}}(t)))\).

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Change history

References

  1. Albano, P., Cannarsa, P.: Propagation of singularities for solutions of nonlinear first order partial differential equations. Arch. Ration. Mech. Anal. 162(1), 1–23 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  2. Albano, P., Cannarsa, P., Nguyen, K.T., Sinestrari, C.: Singular gradient flow of the distance function and homotopy equivalence. Math. Ann. 356(1), 23–43 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  3. Cannarsa, P., Chen, Q., Cheng, W.: Dynamic and asymptotic behavior of singularities of certain weak KAM solutions on the torus. J. Differ. Equ. 267(4), 2448–2470 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  4. Cannarsa, P., Cheng, W.: Generalized characteristics and Lax–Oleinik operators: global theory. Calc. Var. Partial Differ. Equ. 56(5), 125 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  5. Cannarsa, P., Cheng, W.: Local singular characteristics on \(\mathbb{R} ^2\). Boll. Unione Mat. Ital. 14(3), 483–504 (2021)

    Article  MathSciNet  MATH  Google Scholar 

  6. Cannarsa, P., Cheng, W.: Singularities of solutions of Hamilton–Jacobi equations. Milan J. Math. 89(1), 187–215 (2021)

    Article  MathSciNet  MATH  Google Scholar 

  7. Cannarsa, P., Cheng, W., Fathi, A.: On the topology of the set of singularities of a solution to the Hamilton–Jacobi equation. C. R. Math. Acad. Sci. Paris 355(2), 176–180 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  8. Cannarsa, P., Cheng, W., Fathi, A.: Singularities of solutions of time dependent Hamilton–Jacobi equations. Applications to Riemannian geometry. Publ. Math. Inst. Hautes Études Sci. 133(1), 327–366 (2021)

    Article  MathSciNet  MATH  Google Scholar 

  9. Cannarsa, P., Cheng, W., Jin, L., Wang, K., Yan, J.: Herglotz’ variational principle and Lax–Oleinik evolution. J. Math. Pures Appl. 9(141), 99–136 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  10. Cannarsa, P., Cheng, W., Wang, K., Yan, J.: Herglotz’ generalized variational principle and contact type Hamilton-Jacobi equations. In: Trends in Control Theory and Partial Differential Equations, volume 32 of Springer INdAM Ser., pp. 39–67. Springer, Cham (2019)

  11. Cannarsa, P., Sinestrari, C.: Semiconcave functions, Hamilton–Jacobi equations, and optimal control. In: Progress in Nonlinear Differential Equations and their Applications, vol. 58. Birkhäuser Boston Inc, Boston (2004)

  12. Cannarsa, P., Yifeng, Yu.: Singular dynamics for semiconcave functions. J. Eur. Math. Soc. (JEMS) 11(5), 999–1024 (2009)

  13. Cheng, W., Hong, J.: Local strict singular characteristics: Cauchy problem with smooth initial data. J. Differ. Equ. 328, 326–353 (2022)

    Article  MathSciNet  MATH  Google Scholar 

  14. Dafermos, C.M.: Generalized characteristics and the structure of solutions of hyperbolic conservation laws. Indiana Univ. Math. J. 26(6), 1097–1119 (1977)

    Article  MathSciNet  MATH  Google Scholar 

  15. Khanin, K., Sobolevski, A.: On dynamics of Lagrangian trajectories for Hamilton-Jacobi equations. Arch. Ration. Mech. Anal. 219(2), 861–885 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  16. Strömberg, T.: Propagation of singularities along broken characteristics. Nonlinear Anal. 85, 93–109 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  17. Strömberg, T., Ahmadzadeh, F.: Excess action and broken characteristics for Hamilton–Jacobi equations. Nonlinear Anal. 110, 113–129 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  18. Wang, K., Wang, L., Yan, J.: Variational principle for contact Hamiltonian systems and its applications. J. Math. Pures Appl. 9(123), 167–200 (2019)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

Wei Cheng is partly supported by National Natural Science Foundation of China (Grant No. 12231010). The authors would like to thank the anonymous referees for their careful reading and useful comments on the original version of this paper, which have helped us to improve the presentation significantly.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jiahui Hong.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

The original online version of this article was revised: The production team in the original publication of the article incorrectly processed the affiliation of the authors. Now, it has been corrected.

Appendix A. Rescaling of a lipschitz curve

Appendix A. Rescaling of a lipschitz curve

In this section, we suppose \(\tau _0>0\), \(\gamma :[0,\tau _0]\rightarrow {\mathbb {R}}^n\) is a Lipschitz curve with Lipschitz constant \(\kappa \). For \(t\in [0,\tau _0]\), we define

$$\begin{aligned} \tau _1(t)&=\inf \{\tau \in [0,t]:\gamma (r)=\gamma (t),\ \forall r\in [\tau ,t]\},\\ \tau _2(t)&=\sup \{\tau \in [t,\tau _0]:\gamma (r)=\gamma (t),\ \forall r\in [t,\tau ]\}. \end{aligned}$$

Obviously, we have \(0\leqslant \tau _1(t)\leqslant t\leqslant \tau _2(t)\leqslant \tau _0\). For \(0\leqslant t_1\leqslant t_2\leqslant \tau _0\), we define the length of \(\gamma \Big \vert _{[t_1,t_2]}\) as follows

$$\begin{aligned} l(t_1,t_2)=\sup \left\{ \sum _{i=1}^{k}|\gamma (r_i)-\gamma (r_{i-1})|:k\in {\mathbb {N}},\ t_1=r_0\leqslant r_1\leqslant \cdots \leqslant r_k=t_2\right\} . \end{aligned}$$

We can easily obtain \(0\leqslant l(t_1,t_2)\leqslant \kappa (t_2-t_1)\).

Lemma A.1

(1):

For any \(t_1,t_2\in [0,\tau _0]\), we have

$$\begin{aligned}{}[\tau _1(t_1),\tau _2(t_1)]\cap [\tau _1(t_2),\tau _2(t_2)]\ne \emptyset \Longleftrightarrow [\tau _1(t_1),\tau _2(t_1)]=[\tau _1(t_2),\tau _2(t_2)] \end{aligned}$$
(2):

\(l(t_1,t_3)=l(t_1,t_2)+l(t_2,t_3)\), \(\forall 0\leqslant t_1\leqslant t_2\leqslant t_3\leqslant \tau _0\).

(3):

For any \(t_1,t_2\in [0,\tau _0]\), we have

$$\begin{aligned} l(0,t_1)=l(0,t_2)\Longleftrightarrow [\tau _1(t_1),\tau _2(t_1)]=[\tau _1(t_2),\tau _2(t_2)]. \end{aligned}$$
(4):

If \(t,t_i\in [0,\tau _0]\) satisfies \(l(0,t_i)>l(0,t)\), \(i\in {\mathbb {N}}\) and \(\lim _{i\rightarrow \infty }l(0,t_i)=l(0,t)\), then

$$\begin{aligned} t_i>\tau _2(t),\ \forall i\in {\mathbb {N}},\qquad \lim _{i\rightarrow \infty }t_i=\tau _2(t). \end{aligned}$$

Proof

The proofs of (1) and (2) are immediate. To prove (3), without loss of generality, we assume \(t_1\leqslant t_2\). Then

$$\begin{aligned} l(0,t_1)=l(0,t_2)&\Longleftrightarrow l(t_1,t_2)=0 \Longleftrightarrow \gamma (t)=\gamma (t_1),\ \forall t\in [t_1,t_2]\\&\Longleftrightarrow t_2\in [\tau _1(t_1),\tau _2(t_1)] \Longleftrightarrow [\tau _1(t_1),\tau _2(t_1)]=[\tau _1(t_2),\tau _2(t_2)], \end{aligned}$$

where the first equivalence is from (2) and the last is from (1).

Finally, it follows from (3) that \(l(0,\tau _2(t))=l(0,t)<l(0,t_i)\), \(i\in {\mathbb {N}}\), which implies \(t_i>\tau _2(t)\), \(i\in {\mathbb {N}}\). If there exists \({\tilde{t}}>\tau _2(t)\) and a subsequence \(\{t_{i_k}\}\) of \(\{t_i\}\) such that \(t_{i_k}\geqslant {\tilde{t}}\), \(\forall k\in {\mathbb {N}}\), then by (3) we have \(l(t,{\tilde{t}})=l(0,{\tilde{t}})-l(0,t)>0\) and

$$\begin{aligned} l(0,t_{i_k})\geqslant l(0,{\tilde{t}})=l(0,t)+l(t,{\tilde{t}}),\qquad \forall k\in {\mathbb {N}}, \end{aligned}$$

which leads to a contradiction to \(\lim _{i\rightarrow \infty }l(0,t_i)=l(0,t)\). Therefore, we have \(\lim _{i\rightarrow \infty }t_i=\tau _2(t)\). This completes the proof of (4). \(\square \)

Now, we define a curve \(\gamma _l:[0,l(0,\tau _0)]\rightarrow {\mathbb {R}}^n\), \(\gamma _l(s)=\gamma (t)\), where \(t\in [0,\tau _0]\) satisfies \(l(0,t)=s\). Due to Lemma A.1 (3), the curve \(\gamma _l\) is well defined.

Lemma A.2

\(\gamma _l\) is a Lipschitz curve with constant 1 and satisfies

$$\begin{aligned} |{\dot{\gamma }}_l(s)|=1,\qquad a.e.\ s\in [0,l(0,\tau _0)]. \end{aligned}$$
(A.1)

Proof

For any \(0\leqslant s_1\leqslant s_2\leqslant l(0,\tau _0)\), there holds

$$\begin{aligned} |\gamma _l(s_2)-\gamma _l(s_1)|=|\gamma (t_2)-\gamma (t_1)|\leqslant l(t_1,t_2)=l(0,t_2)-l(0,t_1)=s_2-s_1, \end{aligned}$$

where \(t_i\in [0,\tau _0]\) satisfies \(l(0,t_i)=s_i\), \(i=1,2\). Thus, \(\gamma _l\) is a Lipschitz curve with constant 1 and

$$\begin{aligned} |{\dot{\gamma }}_l(s)|\leqslant 1,\qquad a.e.\ s\in [0,l(0,\tau _0)]. \end{aligned}$$
(A.2)

On the other hand, for any \(0=t_0\leqslant t_1\leqslant \cdots \leqslant t_k=\tau _0\), we have

$$\begin{aligned} \sum _{i=1}^{k}|\gamma (t_i)-\gamma (t_{i-1})|&=\sum _{i=1}^{k}|\gamma _l(l(0,t_i))-\gamma _l(l(0,t_{i-1}))|=\sum _{i=1}^{k}|\int _{l(0,t_{i-1})}^{l(0,t_i)}{\dot{\gamma }}_l(s)\ ds|\\&\leqslant \sum _{i=1}^{k}\int _{l(0,t_{i-1})}^{l(0,t_i)}|{\dot{\gamma }}_l(s)|\ ds=\int _{0}^{l(0,\tau _0)}|{\dot{\gamma }}_l(s)|\ ds. \end{aligned}$$

It follows that \(l(0,\tau _0)\leqslant \int _{0}^{l(0,\tau _0)}|{\dot{\gamma }}_l(s)|\ ds\), that is, \(\int _{0}^{l(0,\tau _0)}1-|{\dot{\gamma }}_l(s)|\ ds\leqslant 0\). Combing the inequality with (A.2), (A.1) follows. \(\square \)

Lemma A.3

Suppose \(\gamma :[0,\tau _0]\rightarrow {\mathbb {R}}^n\) is a Lipschitz curve, \(f:{\mathbb {R}}^n\rightarrow {\mathbb {R}}^n\) is a map.

(1):

If for any \(t\in [0,\tau _0]\) with \(\tau _2(t)<\tau _0\), there holds \(\lim _{r\rightarrow \tau _2(t)^+}f(\gamma (r))=f(\gamma (t))\), then we have

$$\begin{aligned} \lim _{r\rightarrow s+}f(\gamma _l(r))=f(\gamma _l(s)),\qquad \forall s\in [0,l(0,\tau _0)). \end{aligned}$$
(2):

Under the assumption of (1), if for any \(t\in [0,\tau _0]\) with \(\tau _2(t)<\tau _0\), there holds

$$\begin{aligned} \lim _{r\rightarrow \tau _2(t)^+}\frac{\gamma (r)-\gamma (t)}{|\gamma (r)-\gamma (t)|}=f(\gamma (t)), \end{aligned}$$

we have

$$\begin{aligned} {\dot{\gamma }}_l^+(s)=f(\gamma _l(s)),\qquad \forall s\in [0,l(0,\tau _0)). \end{aligned}$$

Proof

Set \(s\in [0,l(0,\tau _0))\). For any \(r_i\rightarrow s^+\), \(i\rightarrow \infty \), choose \(t,t_i\in [0,\tau _0]\), \(i\in {\mathbb {N}}\) such that \(l(0,t)=s\), \(l(0,t_i)=r_i\), \(i\in {\mathbb {N}}\). Lemma A.1 (4) implies \(t_i\rightarrow \tau _2(t)^+\), \(i\rightarrow \infty \). Thus,

$$\begin{aligned} \lim _{i\rightarrow \infty }f(\gamma _l(r_i))=\lim _{i\rightarrow \infty }f(\gamma (t_i))=f(\gamma (t))=f(\gamma _l(s)). \end{aligned}$$

It follows that \(\lim _{r\rightarrow s+}f(\gamma _l(r))=f(\gamma _l(s))\), and (1) holds.

Now, we turn to the proof of (2). First, for almost all \(s\in [0,l(0,\tau _0))\), we have

$$\begin{aligned} {\dot{\gamma }}_l(s)=\lim _{r\rightarrow s+}\frac{\gamma _l(r)-\gamma _l(s)}{r-s}=\lim _{r\rightarrow s+}\frac{\gamma (t_r)-\gamma (t)}{|\gamma (t_r)-\gamma (t)|}\cdot \frac{|\gamma _l(r)-\gamma _l(s)|}{r-s}, \end{aligned}$$

where \(t_r\) and t satisfy \(l(0,t_r)=r\) and \(l(0,t)=s\). Lemma A.2 implies

$$\begin{aligned} \lim _{r\rightarrow s+}\frac{|\gamma _l(r)-\gamma _l(s)|}{r-s}=|{\dot{\gamma }}_l(s)|=1. \end{aligned}$$

It follows from Lemma A.1 (4) that

$$\begin{aligned} \lim _{r\rightarrow s+}\frac{\gamma (t_r)-\gamma (t)}{|\gamma (t_r)-\gamma (t)|}=\lim _{r\rightarrow \tau _2(t)^+}\frac{\gamma (r)-\gamma (t)}{|\gamma (r)-\gamma (t)|}=f(\gamma (t))=f(\gamma _l(s)). \end{aligned}$$

Therefore, \({\dot{\gamma }}_l(s)=f(\gamma _l(s))\). Now, for any \(s\in [0,l(0,\tau _0))\), we have

$$\begin{aligned} {\dot{\gamma }}^+_l(s)&=\lim _{r\rightarrow s+}\frac{\gamma _l(r)-\gamma _l(s)}{r-s}=\lim _{r\rightarrow s+}\frac{1}{r-s}\int _{s}^{r}{\dot{\gamma }}_l(\tau )\ d\tau \\&=\lim _{r\rightarrow s+}\frac{1}{r-s}\int _{s}^{r}f(\gamma _l(\tau ))\ d\tau =f(\gamma _l(s)), \end{aligned}$$

where the last equality is from (1). \(\square \)

We say a function \(g:{\mathbb {R}}\rightarrow {\mathbb {R}}\) is b-increasing with \(b>0\) if

$$\begin{aligned} g(t_2) \geqslant g(t_1)+b(t_{2}-t_{1}),\qquad \forall t_1 \leqslant t_2. \end{aligned}$$

Lemma A.4

Suppose \(a>0\), \(0<b_1\leqslant b_2<+\infty \), \(f:[0,a]\rightarrow [b_1,b_2]\) is right-continuous. Then there exists a Lipschitz function \(x:[0,\frac{a}{b_2}]\rightarrow {\mathbb {R}}\) such that it is a solution of

$$\begin{aligned} {\left\{ \begin{array}{ll} {\dot{x}}^+(t)=f(x(t)),\qquad t\in [0,\frac{a}{b_2})\\ x(0)=0 \end{array}\right. } \end{aligned}$$

Proof

For \(m\in {\mathbb {N}}\), we define \(x_m:[0,\frac{a}{b_2}]\rightarrow {\mathbb {R}}\) as follows:

$$\begin{aligned} x_m(t)= {\left\{ \begin{array}{ll} f(0)t,\qquad t\in [0,\frac{1}{m}\frac{a}{b_2}]\\ x_m(\frac{i-1}{m}\frac{a}{b_2})+f(x_m(\frac{i-1}{m}\frac{a}{b_2}))(t-\frac{i-1}{m}\frac{a}{b_2}),\quad t\in [\frac{i-1}{m}\frac{a}{b_2},\frac{i}{m}\frac{a}{b_2}],2\leqslant i\leqslant m. \end{array}\right. } \end{aligned}$$

It is easy to verify that \(x_m\) are all \(b_1\)-increasing, \(b_2\)-Lipschitz functions on \([0,\frac{a}{b_2}]\), and \(x_m(0)=0\). By Alzela-Ascoli theorem, there exists a subsequence \(\{x_{m_k}\}\) and an \(x:[0,\frac{a}{b_2}]\rightarrow {\mathbb {R}}\) such that \(x_{m_k}\) converges uniformly to x on \([0,\frac{a}{b_2}]\). x is also \(b_1\)-increasing, \(b_2\)-Lipschitz on \([0,\frac{a}{b_2}]\), and \(x(0)=0\). So we have

$$\begin{aligned} 0\leqslant x(t)<a,\qquad \forall t\in [0,\frac{a}{b_2}). \end{aligned}$$

Fix any \(t\in [0,\frac{a}{b_2})\). For any \(\varepsilon >0\), by the right-continuity of f, there exists \(\delta >0\) such that

$$\begin{aligned} |f(y)-f(x(t))|<\varepsilon ,\qquad \forall x(t)\leqslant y\leqslant x(t)+\delta . \end{aligned}$$

Since \(x_{m_k}\) converges uniformly to x, there exists \(n_0\in {\mathbb {N}}\) such that \(\Vert x_{m_k}-x\Vert <\frac{1}{2}\delta \) for all \(k\geqslant n_0\). Thus, we have

$$\begin{aligned} x_{m_k}(s)&<x(s)+\frac{1}{2}\delta \leqslant x(t)+(s-t)b_2+\frac{1}{2}\delta \leqslant x(t)+\frac{\delta }{2b_2}b_2+\frac{1}{2}\delta \\&=x(t)+\delta ,\qquad \forall k\geqslant n_0,t\leqslant s\leqslant t+\frac{\delta }{2b_2}. \end{aligned}$$

Furthermore, choosing any \(t<{\tilde{t}}<s\leqslant t+\frac{\delta }{2b_2}\), by the uniform convergence of \(x_{m_k}\), there exists \(n_1\in {\mathbb {N}}\) such that \(\Vert x_{m_k}-x\Vert <\frac{b_1}{2}({\tilde{t}}-t)\) for all \(k\geqslant n_1\). Therefore, when \(k\geqslant \max \{n_0,n_1\}\), \(\frac{t+{\tilde{t}}}{2}\leqslant \tau \leqslant s\), we have

$$\begin{aligned} x_{m_k}(\tau )&\geqslant x_{m_k}\left( \frac{t+{\tilde{t}}}{2}\right) >x\left( \frac{t+{\tilde{t}}}{2}\right) -\frac{b_1}{2}({\tilde{t}}-t)\\&\geqslant x(t)+\frac{b_1}{2}({\tilde{t}}-t)-\frac{b_1}{2}({\tilde{t}}-t)=x(t), \end{aligned}$$

and \(x_{m_k}(\tau )\leqslant x(t)+\delta \). This implies \(|f(x_{m_k}(\tau ))-f(x(t))|<\varepsilon \). Combing this with the definition of \(x_m\), it is not hard to obtain that

$$\begin{aligned} \frac{x(s)-x({\tilde{t}})}{s-{\tilde{t}}}=\lim _{k\rightarrow \infty }\frac{x_{m_k}(s)-x_{m_k}({\tilde{t}})}{s-{\tilde{t}}}\in (f(x(t))-\varepsilon ,f(x(t))+\varepsilon ). \end{aligned}$$

Since \({\tilde{t}}\) is arbitrary, it follows that

$$\begin{aligned} \frac{x(s)-x(t)}{s-t}=\lim _{{\tilde{t}}\rightarrow t^+}\frac{x(s)-x({\tilde{t}})}{s-{\tilde{t}}}\in (f(x(t))-\varepsilon ,f(x(t))+\varepsilon ). \end{aligned}$$

Finally, we have

$$\begin{aligned} {\dot{x}}^+(t)=\lim _{s\rightarrow t^+}\frac{x(s)-x(t)}{s-t}=f(x(t)). \end{aligned}$$

This completes the proof. \(\square \)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cheng, W., Hong, J. Local strict singular characteristics II: existence for stationary equations on \({\mathbb {R}}^2\). Nonlinear Differ. Equ. Appl. 30, 64 (2023). https://doi.org/10.1007/s00030-023-00878-4

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00030-023-00878-4

Keywords

Mathematics Subject Classification

Navigation