Skip to main content
Log in

Existence of an Effective Burning Velocity in a Cellular Flow for the Curvature G-Equation Proved Using a Game Analysis

  • Published:
The Journal of Geometric Analysis Aims and scope Submit manuscript

Abstract

G-equation is a popular level set model in turbulent combustion, and becomes an advective mean curvature type evolution equation when curvature of a moving flame in a fluid flow is considered:

$$\begin{aligned} G_t + \left( 1-d\, \text {div}\left( \frac{DG}{|DG|}\right) \right) _+|DG|+V(x)\cdot DG=0. \end{aligned}$$

Here \(d>0\) is the Markstein number and the positive part \(()_+\) is imposed to avoid a non-physical negative laminar flame speed. For simplicity of presentation, we focus mainly on the case when \(V:{{\mathbb {R}}}^2\rightarrow {{\mathbb {R}}}^2\) is the two dimensional cellular flow with Hamiltonian \(H = \sin x_1 \, \sin x_2\) and amplitude A. Our main result is that for any unit vector \(p\in {{\mathbb {R}}}^2\), there exists a positive number \({\overline{H}}(p)\) such that if \(G(x,0)=p\cdot x\), then

$$\begin{aligned} \left| G(x,t)-p\cdot x+{\overline{H}}(p)t\right| \le C \quad \text {in}\, {{\mathbb {R}}}^2\times [0,\infty ) \end{aligned}$$

for a constant C depending only on on the Markstein number d and the cellular flow amplitude A. The number \({\overline{H}}(p)\) corresponds to the effective burning velocity in the physics literature. The non-coercivity encountered here is one of the major difficulties for homogenization of the mean curvature-type equations. To overcome it, we introduce a new approach that combines PDE methods with a dynamical analysis of the Kohn-Serfaty deterministic game characterization of the curvature G-equation utilizing the streamline structure of cellular flows. Extension to general two-dimensional incompressible flows is also discussed. In three dimensional incompressible flows, the existence of \({\overline{H}}(p)\) might fail when the flow intensity exceeds a bifurcation value even for simple shear flows Mitake et al. (Bifurcation of homogenization and nonhomogenization of the curvature G-equation with shear flows (2023)).

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.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15

Similar content being viewed by others

Data Availability

This paper does not generate any data.

References

  1. Alvarez, O., Bardi, M.: Singular perturbations of degenerate parabolic PDEs: a general convergence result. Arch. Rational Mech. Anal. 170, 17–61 (2003)

    Article  ADS  MathSciNet  Google Scholar 

  2. Armstrong, S., Cardaliaguet, P.: Stochastic homogenization of quasilinear Hamilton-Jacobi equations and geometric motions. J. Eur. Math. Soc. (JEMS) 20(4), 797–864 (2018)

    Article  MathSciNet  Google Scholar 

  3. Arnold, V.I.: Topological and ergodic properties of closed 1-forms with incommensurable periods. Funct. Anal. Its Appl. 25, 81–90 (1991)

    Article  MathSciNet  Google Scholar 

  4. Caffarelli, L.A.: The homogenization of surfaces and boundaries. Bll Braz. Math. Soc. New Series 44(4), 755–775 (2013)

    Article  MathSciNet  Google Scholar 

  5. Caffarelli, L.A., Monneau, R.: Counterexample in three dimension and homogenization of geometric motions in two dimension. Arch. Ration. Mech. Anal. 212, 503–574 (2014)

    Article  MathSciNet  Google Scholar 

  6. Cardaliaguet, P., Lions, P.L., Souganidis, P.E.: A discussion about the homogenization of moving interfaces. J. Math. Pures Appl. 91(4), 339–363 (2009)

    Article  MathSciNet  Google Scholar 

  7. Cardaliaguet, P., Nolen, J., Souganidis, P.: Homogenization and enhancement for the G-equation in periodic media. Arch. Ration. Mech. Anal. 199(2), 527–561 (2011)

    Article  MathSciNet  Google Scholar 

  8. Cesaroni, A., Novaga, M.: Long-time behavior of the mean curvature flow with periodic forcing. Commun. Partial Differ. Eqs. 38(5), 780–801 (2013)

    Article  MathSciNet  Google Scholar 

  9. Chaudhuri, S., Wu, F., Law, C.K.: Scaling of turbulent flame speed for expanding flames with Markstein diffusion considerations. Phys. Rev. E 88, 033005 (2013)

    Article  ADS  Google Scholar 

  10. Chen, Y.G., Giga, Y., Goto, S.: Uniqueness and existence of viscosity solutions of generalized mean curvature flow equations. J. Differ. Geo. 33, 749–786 (1991)

    MathSciNet  Google Scholar 

  11. Childress, S., Gilbert, A.D.: Stretch, Twist, Fold: The Fast Dynamo. Lecture Notes in Physics Monographs, 37, Springer (1995)

  12. Craciun, B., Bhattacharya, K.: Effective motion of a curvature-sensitive interface through a heterogeneous medium. Interf. Free Bound. 6(2), 151–173 (2004)

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  14. Dirr, N., Karali, G., Yip, N.K.: Pulsating wave for mean curvature flow in inhomogeneous medium. Eur. J. Appl. Math. 19(6), 661–699 (2008)

    Article  MathSciNet  CAS  Google Scholar 

  15. Evans, L.C.: The perturbed test function method for viscosity solutions of nonlinear PDE. Proce. R. Soc. Edinb. Sect. A: Math. 111(3–4), 359–375 (1989)

    Article  MathSciNet  Google Scholar 

  16. Evans, L.C., Spruck, J.: Motion of level sets by mean curvature. J. Differ. Geom. 33, 635–681 (1991)

    Article  MathSciNet  Google Scholar 

  17. Feldman, W.M.: Mean curvature flow with positive random forcing in 2-d, arXiv:1911.00488 (2019)

  18. Gao, H., Kim, I.: Head and tail speeds of mean curvature flow with forcing. Arch. Ration. Mech. Anal. 235(1), 287–354 (2020)

    Article  MathSciNet  Google Scholar 

  19. Giga, Y., Mitake, H., Tran, H.V.: On asymptotic speed of solutions to level-set mean curvature flow equation with driving and source terms. Siam J. Math. Anal. 48(5), 3515–3546 (2016)

    Article  MathSciNet  Google Scholar 

  20. Hamamuki, N., Liu, Q.: A game-theoretic approach to dynamic boundary problems for level-set curvature flow equations and applications. SN Partial Differ. Eqs. Appl. 2, 30 (2021)

    Article  MathSciNet  Google Scholar 

  21. Jing, W., Tran, H.V., Yu, Y.: Effective fronts of polygon shapes in two dimensions, arXiv:2112.10747 [math.AP] (2021)

  22. Kerstein, A.R., Ashurst, W.T., Williams, F.A.: Field equation for interface propagation in an unsteady homogeneous flow field. Phys. Rev. A 37, 2728 (1988)

    Article  ADS  CAS  Google Scholar 

  23. Kohn, R.V., Serfaty, S.: Second-order PDE’s and deterministic games. Proceedings of ICIAM, pp 239–249 (2007)

  24. Kohn, R.V., Serfaty, S.: A deterministic-control-based approach to motion by mean curvature. Comm. Pure. Appl. Math 59, 344–407 (2006)

    Article  MathSciNet  Google Scholar 

  25. Langer, J.: Instabilities and pattern formation in crystal growth. Rev. Mod. Phys. 52, 1 (1980)

    Article  ADS  CAS  Google Scholar 

  26. Lions, P.L., Souganidis, P.E.: Homogenization of degenerate second-order PDE in periodic and almost periodic environments and applications. Ann. Inst. H. Poincaré, Anal. Non Linéaire 22(5), 667–677 (2005)

    Article  ADS  MathSciNet  Google Scholar 

  27. Liu, Q.: Waiting time effect for motion by positive second derivatives and applications. Nonlinear Differ. Equ. Appl. 21, 589–620 (2014)

    Article  MathSciNet  Google Scholar 

  28. Liu, Y., Xin, J., Yu, Y.: Asymptotics for turbulent flame speeds of the viscous G-equation enhanced by cellular and shear flows. Arch. Rational. Mechanics. Anal 199(2), 461–492 (2011)

    Article  ADS  MathSciNet  Google Scholar 

  29. Lyu, J., Xin, J., Yu, Y.: Curvature Effect in Shear Flow: Slowdown of Turbulent Flame Speeds with Markstein Number. Commun. Math. Phys. 359, 515–533 (2018)

    Article  ADS  MathSciNet  Google Scholar 

  30. Markstein, G.H.: Experimental and theoretical studies of flame front stability. J. Aero. Sci. 18, 199–209 (1951)

    Article  Google Scholar 

  31. Mitake, H., Mooney, C., Tran, H.V., Xin, J., Yu, Y.: Bifurcation of homogenization and nonhomogenization of the curvature G-equation with shear flows, preprint, arXiv:2303.16304 [math.AP] (2023)

  32. Osher, S., Fedkiw, R.: Level Set Methods and Dynamic Implicit Surfaces. Springer-Verlag, New York (2002)

    Google Scholar 

  33. Osher, S., Sethian, J.: Fronts propagating with curvature-dependent speed: Algorithms based on Hamilton-Jacobi formulations. J. Comput. Phys. 79(1), 12–49 (1988)

    Article  ADS  MathSciNet  Google Scholar 

  34. Peters, N.: Turbulent Combustion. Cambridge University Press, Cambridge (2000)

    Book  Google Scholar 

  35. Ronney, P.: Some Open Issues in Premixed Turbulent Combustion, In: J. D. Buckmaster, T. Takeno, (Eds.). Modeling in Combustion Science. Lecture Notes In Physics, Vol. 449, Springer-Verlag, Berlin, pp. 3–22 (1995)

  36. Sethian, J.: Curvature and the evolution of fronts. Comm. Math. Phys. 101(4), 487–499 (1985)

    Article  ADS  MathSciNet  Google Scholar 

  37. Spencer, J.: Balancing games. J. Combinatorial Th. Ser B 23, 68–74 (1977)

    Article  MathSciNet  Google Scholar 

  38. Tran, H.V., Yu, Y.: Differentiability of effective fronts in the continuous setting in two dimensions, arXiv:2203.13807 [math.AP], (2022)

  39. Williams, F.: Turbulent combustion. In: Buckmaster, J. (ed.) The Mathematics of Combustion, pp. 97–131. SIAM, Philadelphia (1985)

    Chapter  Google Scholar 

  40. Xin, J., Yu, Y.: Periodic Homogenization of Inviscid G-equation for Incompressible Flows. Comm. Math. Sci. 8(4), 1067–1078 (2010)

    Article  MathSciNet  Google Scholar 

  41. Xin, J., Yu, Y.: Sharp asymptotic growth laws of turbulent flame speeds in cellular flows by inviscid Hamilton-Jacobi models. Ann. Insti. H. Poincar. Anal. Non Linéaire 30(6), 1049–1068 (2013)

    Article  ADS  MathSciNet  Google Scholar 

  42. Xin, J., Yu, Y.: Front Quenching in G-equation Model Induced by Straining of Cellular Flow. Arch. Rational. Mech. Anal. 214, 1–34 (2014)

    Article  ADS  MathSciNet  Google Scholar 

  43. Xin, J., Yu, Y., Zlatoš, A.: Periodic orbits of the ABC flow with \(A=B=C=1\). SIAM J. Math. Anal. 48(6), 4087–4093 (2016)

    Article  MathSciNet  Google Scholar 

  44. Zhu, J., Ronney, P.: Simulation of Front Propagation at Large Non-dimensional Flow Disturbance Intensities. Combust. Sci. Technol. 100(1), 183–201 (1994)

    Article  CAS  Google Scholar 

Download references

Acknowledgements

The authors would like to thank Inwon Kim and Hung V. Tran for helpful comments in improving the presentation of the paper. We are grateful to Robert V. Kohn for fruitful discussions of game representations and for providing references. Last but not the least, we thank the anonymous referees for all the constructive comments.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jack Xin.

Additional information

Publisher's Note

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

The work was partly supported by NSF grants DMS-1952644/DMS-2309520 (JX) and DMS-2000191 (YY).

Appendix

Appendix

It is well known that the front propagation under the viscosity solution framework is consistent with the classical meaning when the smooth solutions exist. See section 6 in [16] for instance. The following conclusion is a special case in our context, which is needed to derive a reachability property. For the reader’s convenience, we present its proof here in 2D, which is sufficient for our purpose.

Throughout this section, we only assume that \(V\in W^{1,\infty }({{\mathbb {R}}}^2)\) and

$$\begin{aligned} V(x)\cdot (1,0)=0 \quad \text {for}\, x\in \{0\}\times [0,1]. \end{aligned}$$
(5.1)

Let \(S=(0,1)^2\) and

$$\begin{aligned} g_S(x)={\left\{ \begin{array}{ll} -{2\over \pi }\arctan (d(x,\partial S)) \quad \text {for}\, x\in S\\ {2\over \pi }\arctan (d(x,\partial S)) \quad \text {otherwise}. \end{array}\right. } \end{aligned}$$

Here \(d(x,\partial S)\) is the distance from x to \(\partial S\).

Lemma 5.1

Suppose that \(G\in C({{\mathbb {R}}}^2\times [0,\infty ))\) is the unique viscosity solution to equation (1.2) subject to

$$\begin{aligned} G(x,0)=g_S(x). \end{aligned}$$

Then for a given \(\delta \in (0,{1\over 2})\), there exists \(t_{\delta }>0\) depending only on d, V and \(\delta \) such that

$$\begin{aligned} G((0,\theta ), t)<0 \quad \text {for}\, (\theta , t)\in [\delta , 1-\delta ]\times (0, t_\delta ]. \end{aligned}$$

Proof

Intuitively, this conclusion is obvious since the speed along the normal direction \(\vec {n}=(-1,0)\) at the point \((0, \theta )\) is

$$\begin{aligned} v_{\vec {n}}=1-d\kappa +V(x)\cdot \vec {n}=1. \end{aligned}$$

To make this rigorous, we need to build smooth supersolutions and employ comparison principle. It suffices to prove this at a fixed \(\theta \in [\delta ,1-\delta ]\).

Step 1: Choose an ellipse. Consider the ellipse

$$\begin{aligned} E_{\theta }(t): \frac{(x_1-a_0 - \nu )^2}{a^2(t)} + \frac{(x_2-\theta )^2}{b^2(t)} = 1. \end{aligned}$$

Here \(\nu \in (0,a_0)\) is added for technical convenience and will be sent to zero later (See Fig. 16).

Fig. 16
figure 16

Propagation of the ellipse

Let \(b_0={\delta \over 2}\). Then choose \(a_0\in (0, {b_0\over 4})\) small enough such that

$$\begin{aligned} |V(x)\cdot (1,0)|<{1\over 8} \quad \text {if}\, x\in [-4a_0, 4a_0]\times [0,1] \end{aligned}$$

and

$$\begin{aligned} \frac{d64a_0}{b_{0}^2}+4M_0\sqrt{3}\frac{a_0}{b_{0}}<{1\over 8}. \end{aligned}$$

Here \(M_0=\max _{x\in {{\mathbb {R}}}^2}|V(x)|\).

Then we define \(a(t) = a_0 + \frac{1}{2}t\) and \(b(t) = b_{0} - Lt\) for \(L>0\) satisfying

$$\begin{aligned} {1\over 2}-{3La_0\over 4b_0}<1-{db_0\over {a_0}^{2}}-M_0. \end{aligned}$$

Hereafter we require \(0 \le t \le t_\delta \) for \(t_\delta = \min \left\{ 2a_0, \frac{b_{0}}{2\,L}\right\} \), which implies that

$$\begin{aligned} (a(t), b(t))\in [a_0,\ 2a_0]\times \left[ \frac{b_0}{2},\ b_0\right] \end{aligned}$$

and

$$\begin{aligned} E_\theta (t)\subset R_0=[-4a_0, 4a_0]\times [0,1] \end{aligned}$$

for \(t\in [0,t_\delta ]\). Also, \(4a_0<b_0\) leads to

$$\begin{aligned} a(t)< b(t) \quad \text {for all}\, t\in [0,t_\delta ]. \end{aligned}$$

For convenience, we drop the dependence of a and b on t. Hereafter \(t\in [0,t_\delta ]\) unless specified otherwise. Let \(\vec {n}\) be the outward unit normal vector along \(E_{\theta }(t)\) that has the following parameterization: for \(\phi \in [0, 2\pi ]\)

$$\begin{aligned} {\left\{ \begin{array}{ll} x_1=a_0+\nu +a\cos \phi \\ x_2=\theta +b\sin \phi . \end{array}\right. } \end{aligned}$$

Then

$$\begin{aligned} {V} \cdot \vec {n} = {V} \cdot \left( \frac{b\cos \phi }{\sqrt{a^2\sin ^2\phi + b^2\cos ^2\phi }}, \frac{a\sin \phi }{\sqrt{a^2\sin ^2\phi + b^2\cos ^2\phi }}\right) . \end{aligned}$$

If \(|\sin \phi |<{\sqrt{3}\over 2}\),

$$\begin{aligned} \begin{array}{ll} |{V} \cdot \vec {n}| &{}\le |V(x)\cdot (1,0)| + M_0 \cdot \frac{a\sin \phi }{\sqrt{a^2\sin ^2\phi + b^2\cos ^2\phi }}\\ &{}\le |V(x)\cdot (1,0)| + M_0 \cdot \frac{a}{b} \cdot |\tan \phi |\\ &{} \le |V(x)\cdot (1,0)| + M_0 \cdot \frac{2a_0}{b_0/2} \cdot \sqrt{3} \\ &{}= |V(x)\cdot (1,0)|+ 4\sqrt{3}M_0 \cdot \frac{a_0}{b_0}< {1\over 8}+{1\over 8}={1\over 4}. \end{array} \end{aligned}$$
(5.2)

Step 2: Evolution of an elliptic boundary. Let us recall some basic facts. Given a \(C^1\) function f(xt) and the family of level curves

$$\begin{aligned} C(t)=\{x\in {{\mathbb {R}}}^2\ | f(x,t)=0\}, \end{aligned}$$

if \(D_xf\not =0\), the propagation speed of C(t) along the outward normal direction \(\vec {n}={D_xf\over |D_xf|}\) is given by

$$\begin{aligned} v_{\vec n}=-{f_t\over |D_xf|}, \end{aligned}$$

which can be easily derived through the chain rule. Moreover, the corresponding mean curvature along C(t) is

$$\begin{aligned} \kappa =\textrm{div}_x(\vec {n}). \end{aligned}$$

Now let us verify that for \(t\in [0,t_\delta ]\), the propagation of \(E_{\theta }(t)\) obeys the following inequality:

$$\begin{aligned} v_{\vec n}< 1 - d\kappa + {V} \cdot \vec {n}, \end{aligned}$$
(5.3)

which will be used to construct a supersolution. Fix any \(a, b > 0\) and an ellipse

$$\begin{aligned} \frac{x_1^2}{a^2} + \frac{x_2^2}{b^2} = 1, \end{aligned}$$

direct compuations show that its curvature at point \(P(a\cos \phi , b\sin \phi )\), \(0 \le \phi < 2\pi \), is

$$\begin{aligned} \kappa = \frac{ab}{\left( a^2\sin ^2\phi + b^2\cos ^2\phi \right) ^{\frac{3}{2}}} \end{aligned}$$

and the normal velocity at \(P(a\cos \phi , b\sin \phi )\) is

$$\begin{aligned} v_{\vec {n}} = \frac{a^{\prime }b\cos ^2\phi + ab^{\prime }\sin ^2\phi }{\sqrt{a^2\sin ^2\phi + b^2\cos ^2\phi }}. \end{aligned}$$

Case 1. If \(|\sin \phi |\le {\sqrt{3}\over 2}\), then

$$\begin{aligned} d\kappa = \frac{dab}{\left( a^2\sin ^2\phi + b^2\cos ^2\phi \right) ^{\frac{3}{2}}} \le \frac{dab}{\left( b^2\cos ^2\phi \right) ^{\frac{3}{2}}} \le \frac{dab}{(b/2)^3}\le \frac{d64a_0}{b_{0}^2} <{1\over 8}. \end{aligned}$$

Note that

$$\begin{aligned} v_{\vec n}= \frac{\frac{b}{2}\cos ^2\phi - La \sin ^2\phi }{\sqrt{a^2 \sin ^2\phi + b^2\cos ^2\phi }} \le {{b\over 2}\cos ^2\phi \over \sqrt{b^2\cos ^2\phi }}\le \frac{1}{2}. \end{aligned}$$

Since for \(t\in [0,t_\delta ]\),

$$\begin{aligned} E_{\theta }(t)\subset [-4a_0, 4a_0]\times [0,1], \end{aligned}$$

(5.2) implies that \(v_{\vec {n}} \le {1\over 2}< 1-{1\over 8}-{1\over 4}< 1 -d \kappa + {V} \cdot \vec {n}\).

Case 2. If \(|\sin \phi |\ge {\sqrt{3}\over 2}\), then

$$\begin{aligned} v_{\vec n}= & {} \frac{\frac{b}{2}\cos ^2\phi - La \sin ^2\phi }{\sqrt{a^2 \sin ^2\phi + b^2\cos ^2\phi }} \le {{b\over 2}\cos ^2\phi \over \sqrt{b^2\cos ^2\phi }}-{La\sin ^2\phi \over b}\\\le & {} {1\over 2}-{3La\over 4b}\le {1\over 2}-{3La_0\over 4b_0}. \end{aligned}$$

Meanwhile,

$$\begin{aligned} d\kappa = \frac{dab}{\left( a^2\sin ^2\phi + b^2\cos ^2\phi \right) ^{\frac{3}{2}}}\le {db\over a^2}\le {db_0\over {a_0}^{2}}. \end{aligned}$$

Therefore, due to the choice of L, we have \(v_{\vec n}< 1 - d\kappa + {V} \cdot \vec {n}\).

Combining case 1 and case 2, we see that (5.3) holds, i.e., the evolution of \(E_{\theta }(t)\) satisfies

$$\begin{aligned} v_{\vec n} < 1 -d \kappa + {V} \cdot \vec {n},&\text {for}&0 \le t \le t_\delta . \end{aligned}$$

Step 3: Comparison. Let

$$\begin{aligned} \Psi (x,t)=\frac{(x_1-a_0 - \nu )^2}{a^2(t)} + \frac{(x_2-\theta )^2}{b^2(t)} -1. \end{aligned}$$

Owing to (5.3), we may choose \(\mu _0\in (0, {1\over 2})\) such that (5.3) also holds along the curve \(\{x\in {{\mathbb {R}}}^2|\ \Psi (x,t)=\mu \}\) for all \(\mu \in [-\mu _0,\mu _0]\) and \(t\in [0,t_\delta ]\). Equivalently, for

$$\begin{aligned} D_{\theta }= & {} \left\{ (x,t)\in {{\mathbb {R}}}^2\times [0,t_\delta ]\ |-\mu _0\le \Psi (x,t) \le \mu _0\right\} ,\nonumber \\{} & {} \Psi _t + \left( 1-d\, \, \textrm{div}\left( {D\Psi \over |D\Psi |}\right) \right) |D\Psi |+V(x)\cdot D\Psi \ge 0 \quad \text {on}\, D_{\theta }. \end{aligned}$$
(5.4)

Let \(\{h_k\}_{k\ge 1}\in C^{\infty }({{\mathbb {R}}})\) be a sequence of functions such that

$$\begin{aligned} 0<h_{k}^{'}\le 1 \quad \text {in}\, I_k=\left( -\mu _0+{1\over k},\ \mu _0-{1\over k}\right) ,\quad h_{k}^{'}=0 \quad \text {in}\, {{\mathbb {R}}}\backslash I_k \end{aligned}$$

and \(\lim _{k\rightarrow +\infty }h_k(s)=h(s)\) uniformly in \({{\mathbb {R}}}\), where

$$\begin{aligned} h(s)={\left\{ \begin{array}{ll} \mu _0 \quad \text {for}\, s\ge \mu _0\\ s\quad \text {for}\, s\in [-\mu _0, \mu _0]\\ -\mu _0 \quad \text {for}\, s\le -\mu _0. \end{array}\right. } \end{aligned}$$

Apparently, \(\Psi _k(x,t)=h_k(\Psi (x,t))\) satisfies

$$\begin{aligned} {\partial \Psi _{k}\over \partial t} + \left( 1-d\, \, \textrm{div}\left( {D\Psi _k\over |D\Psi _k|}\right) \right) |D\Psi _k|+V(x)\cdot D\Psi _k\ge 0 \quad \text {on}\, {{\mathbb {R}}}^2\times (0,t_\delta ). \end{aligned}$$
(5.5)

By stability, we have that \(G_1(x,t)=h(\Psi (x,t))\in W^{1,\infty }({{\mathbb {R}}}^2\times [0,\infty ))\) is a viscosity supersolution of

$$\begin{aligned} {\partial G_{1}\over \partial t }+ \left( 1-d\, \, \textrm{div}\left( {DG_1\over |DG_1|}\right) \right) |DG_1|+V(x)\cdot DG_1\ge 0 \quad \text {on}\, {{\mathbb {R}}}^2\times (0,t_\delta ). \end{aligned}$$
(5.6)

Since \((a)_+\ge a\), \(G_1=G_1(x,t)\) is also a viscosity supersolution of equation (1.2) on \({{\mathbb {R}}}^2\times (0,t_\delta )\).

Because for fixed \(\nu >0\)

$$\begin{aligned} \{G_1(x,0)\le 0\}=\{\Psi (x,0)\le 0\}\subset S, \end{aligned}$$

we can choose a function \(\xi \in C^{\infty }({{\mathbb {R}}})\) such that \(\dot{\xi }>0\), \(\xi (0)=0\), \(\sup _{s\in {{\mathbb {R}}}}|\xi (s)|<\infty \) and

$$\begin{aligned} g(x)\le \xi (G_1(x,0)). \end{aligned}$$

Since \(\xi (G_1(x,0))\) is also a viscosity supersolution of equation (1.2) on \({{\mathbb {R}}}^2\times (0,t_\delta )\), thanks to Theorem 2.1, we have that

$$\begin{aligned} G(x,t)\le \xi (G_1(x,t)) \quad \text {for}\, (x,t)\in {{\mathbb {R}}}^2\times [0,t_\delta ]. \end{aligned}$$

In particular, this implies that

$$\begin{aligned} \{x\in {{\mathbb {R}}}^2|\ G_1(x,t)<0\}=\{x\in {{\mathbb {R}}}^2|\ \xi (G_1(x,t))<0\}\subset \{x\in {{\mathbb {R}}}^2|\ G(x,t)<0\}. \end{aligned}$$

Note that \(\{x\in {{\mathbb {R}}}^2|\ G_1(x,t)<0\}=\{x\in {{\mathbb {R}}}^2|\ \Psi (x,t)<0\}\). Sending \(\nu \rightarrow 0\), we have that for \(t\in [0,t_\delta ]\),

$$\begin{aligned} \left\{ x=(x_1,x_2)\in {{\mathbb {R}}}^2|\ \frac{(x_1-a_0)^2}{a^2(t)} + \frac{(x_2-\theta )^2}{b^2(t)}<1\right\} \subset \{x\in {{\mathbb {R}}}^2|\ G(x,t)<0\}. \end{aligned}$$

Then

$$\begin{aligned} G((0,\theta ),t)<0 \quad \text {for}\, t\in (0, t_\delta ], \end{aligned}$$

which finishes the proof. Note that \(t_\delta \) only depends on \(\delta \) and V. \(\square \)

Let \(\Omega \subset {{\mathbb {R}}}^2\) be an open convex set. Denote by \(G_{\Omega }(x,t)\) the unique viscosity solution to equation (1.2) subject to \(G_\Omega (x,0)=g_{\Omega }(x)\) where

$$\begin{aligned} g_{\Omega }(x)={\left\{ \begin{array}{ll} -{2\over \pi }\arctan (d(x, \partial \Omega )) \quad \text {for}\, x\in \Omega \\ {2\over \pi }\arctan (d(x,\partial \Omega ) )\quad \text {otherwise}. \end{array}\right. } \end{aligned}$$

Given two sets \(E_1\) and \(E_2\), their Hausdorff distance

$$\begin{aligned} d_H(E_1, E_2)=\max \{ \max _{x\in E_1}d(x, E_2), \ \max _{x\in E_2}d(x, E_1)\}. \end{aligned}$$

Also, for \(\alpha >0\) and \(\delta \in (0, {1\over 2})\), we write

$$\begin{aligned} W_{\alpha , \delta }=[-\alpha , \alpha ]\times [\delta , 1-\delta ]. \end{aligned}$$

Lemma 5.2

Let \(S=(0,1)^2\). For given \(\delta \in (0,{1\over 2})\) and \(n\in {{\mathbb {N}}}\), there exists \(\sigma _{\delta , n}>0\) such that if

$$\begin{aligned} d_H(S, \Omega )\le \sigma _{\delta ,n} \quad \textrm{and} \quad \alpha \le \sigma _{\delta ,n}, \end{aligned}$$

then

$$\begin{aligned} G_\Omega (x,t)<0 \quad \text {for}\, (x,t)\in W_{\alpha ,\delta }\times \left[ {t_{\delta }\over n},\ t_{\delta }\right] . \end{aligned}$$

Here \(t_\delta \) is from the previous Lemma 5.1.

Proof

We argue by contradiction. If not, then there exist a sequence of convex open sets \(\{\Omega _m\}_{m\ge 1}\) such that

$$\begin{aligned} d_H(S, \Omega _m)\le {1\over m} \end{aligned}$$

and for some \((x_m,t_m)\in W_{{1\over m},\delta }\times \left[ {t_{\delta }\over n},\ t_{\delta }\right] \)

$$\begin{aligned} G_{\Omega _m}(x_m,t_m)\ge 0. \end{aligned}$$
Fig. 17
figure 17

Reach \(\Omega \) from \(W_{\alpha ,\delta }\)

Since

$$\begin{aligned} \lim _{m\rightarrow +\infty }g_{\Omega _m}(x)=g_S(x) \quad \text {uniformly on}\, {{\mathbb {R}}}^2, \end{aligned}$$

due to Remark 2.1, the uniqueness of viscosity solutions, we have that

$$\begin{aligned} \lim _{m\rightarrow +\infty }G_{\Omega _m}(x,t)=G(x,t) \quad \text {locally uniformly on}\, {{\mathbb {R}}}^2\times [0,\infty ). \end{aligned}$$

Here G is from Lemma 5.1. The proof is similar to that of (2.4). Also, up to a subsequence if necessary, we may assume that \(\lim _{m\rightarrow +\infty }(x_m,t_m)=((0,\theta ), {\bar{t}})\) for \((\theta , {\bar{t}})\in [\delta , 1-\delta ]\times \left[ {t_{\delta }\over n},\ t_{\delta }\right] \). Then we have that

$$\begin{aligned} G((0,\theta ), {\bar{t}})=\lim _{m\rightarrow +\infty }G_{\Omega _m}(x_m,t_m)\ge 0. \end{aligned}$$

This is a contradiction. \(\square \)

As an immediate corollary, we have the following reachability.

Lemma 5.3

Consider the game in section 2.2. Under the assumption of Lemma 5.2, every point on \(W_{\alpha ,\delta }\) can reach \(\Omega \) within time \(t_{\delta }\over n\). Also, it is easy to see that if we replace S by an arbitrary rectangle, all previous results are still true in the corresponding forms (See Fig. 17).

Remark 5.1

Due to the hidden stochastic nature of the game trajectory, it is not clear to us how to use pure game dynamics to prove the above reachability conclusion. An interesting analog in random walk (or Brownian motion) is to use strong maximum principle of the Laplace equation to show that a particle has a positive probability to exit from any small window of the boundary.

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

Gao, H., Long, Z., Xin, J. et al. Existence of an Effective Burning Velocity in a Cellular Flow for the Curvature G-Equation Proved Using a Game Analysis. J Geom Anal 34, 81 (2024). https://doi.org/10.1007/s12220-023-01523-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s12220-023-01523-3

Keywords

Mathematics Subject Classification

Navigation