Proc. Indian Acad. Sci. (Math. Sci.) Vol. 121, No. 3, August 2011, pp. 349–367. c Indian Academy of Sciences Value functions for certain class of Hamilton Jacobi equations ANUP BISWAS1, RAJIB DUTTA1 and PROSENJIT ROY2 1 TIFR Centre for Applicable Mathematics, Sharadanagar, Chikkabommasandra, Bangalore 560 065, India 2 Institut für Mathematik, Universität Zürich, Winterthurerstrasse 190, 8057 Zürich, Switzerland E-mail: [email protected]; [email protected]; [email protected] MS received 1 September 2010; revised 11 October 2010 Abstract. We consider a class of Hamilton Jacobi equations (in short, HJE) of type ut + eu 1 |u x1 |2 + · · · + |u xn−1 |2 + |u xn |m = 0, 2 m in Rn × R+ and m > 1, with bounded, Lipschitz continuous initial data. We give a Hopf-Lax type representation for the value function and also characterize the set of minimizing paths. It is shown that the minimizing paths in the representation of value function need not be straight lines. Then we consider HJE with Hamiltonian decreasing in u of type u t + H1 u x1 , . . . , u xi + e−u H2 u xi+1 , . . . , u xn = 0 where H1 , H2 are convex, homogeneous of degree n, m > 1 respectively and the initial data is bounded, Lipschitz continuous. We prove that there exists a unique viscosity solution for this HJE in Lipschitz continuous class. We also give a representation formula for the value function. Keywords. function. Hamilton Jacobi equation; viscosity solution; minimizing paths; value 1. Introduction Let H : R × Rn → R be a continuous function and g : Rn → R be a bounded, Lipschitz continuous function. Consider the Hamilton Jacobi equation (HJE) u t + H (u, Du) = 0 for (x, t) ∈ Rn × R+ , u(x, 0) = g(x), (1.1) ∂u where Du = ∂ x1 , . . . , ∂∂u xn . This problem has been studied extensively in [2,4,5] and the references therein. Some of the open issues related to problem (1.1) are the following: (i) Does there exist a viscosity solution for HJE (1.1) and is the solution unique? (ii) Is it possible to give a Hopf-Lax type representation for the value function (solution)? Are the minimizing paths always straight lines? 349 350 Anup Biswas, Rajib Dutta and Prosenjit Roy The question (ii) was also asked in [3]. It has been shown in [5] that apart from the usual hypothesis in H if the H is increasing in u then (1.1) has a unique viscosity solution. But when the Hamiltonian is decreasing in u then a few results are known. Even in the general set up when H (u, p) depends on u, getting a Hopf-Lax type representation for u is quite difficult. For n = 1, an explicit Hopf-Lax type representation is given in [1]. But the dimensionality was crucially used there and its implementation for higher dimension is quite difficult. Here we consider a certain class of HJE ut + eu 1 |u x1 |2 + · · · + |u xn−1 |2 + |u xn |m = 0 2 m and give a Hopf-Lax type representation for u(x, t). We also characterize the set of minimizers for the representation and show that the minimizing paths could be of exponential types. Apart from minimizer characterizations, it is also possible to give the Hopf-Lax type formula for solution u(x, t) for a more general Hamiltonian (see Remark 2.2). Even though it is not known, in general, whether the viscosity solution exists or not when the Hamiltonian is decreasing in u, we give a class of Hamiltonian decreasing in u and show that there exists a unique viscosity solution for the corresponding HJE. For the convenience of the reader we recall the definition of viscosity solution, DEFINITION 1.1 Let U be a bounded, continuous function in Rn × R+ . 1. U is said to be a subsolution of (1.1) if for any (x0 , t0 ) ∈ Rn × R+ , ν ∈ C 1 (Rn × R+ ) such that (x0 , t0 ) is a local maximum for U − ν, with U (x0 , t0 ) = ν(x0 , t0 ), then at (x0 , t0 ), νt + H (ν, Dν) ≤ 0 and limt→0 U (x, t) ≤ g(x). 2. U is said to be a supersolution of (1.1) if for any (x0 , t0 ) ∈ Rn ×R+ , ν ∈ C 1 (Rn ×R+ ) such that (x0 , t0 ) is a local minimum for U − ν, with U (x0 , t0 ) = ν(x0 , t0 ), then at (x0 , t0 ), νt + H (ν, Dν) ≥ 0 and limt→0 U (x, t) ≥ g(x). U is said to be a viscosity solution if it is both sub and supersolution. The paper is organized as follows: In §2 we consider HJE with Hamiltonian increasing in u. We prove that there exists a unique viscosity solution for the HJE. Then we characterize the minimizing paths in the representation of u(x, t) in §3. In §4 we consider HJE with Hamiltonian decreasing in u and prove that there exists a unique viscosity solution for the corresponding HJE. In §5 we provide an example to strengthen the arguments of §3. 2. Value function for HJE increasing in u Let g : Rn → R be a bounded, Lipschitz continuous function. Consider the HJE u t + H (u, Du) = 0 for (x, t) ∈ Rn × R+ , u(x, 0) = g(x), (2.1) Value functions for HJE ∂u ∂u ∂ x1 , . . . , ∂ xn where Du = H (u, p) = and n−1 1 2 351 | pi |2 + i=1 eu | pn |m for (u, p) ∈ R × Rn . m We assume that m > 1 so that H (u, ·) becomes a convex function in Rn . For x ∈ Rn and t ≥ 0, we define tx = {γ : [0, t] → Rn : γ is absolutely continuous and γ (t) = x}. Define L(x1 , . . . , xn−1 ) = 1 |x1 |2 + · · · + |xn−1 |2 for (x1 , . . . , xn−1 ) ∈ Rn−1 . 2 Let γ̄ (l) = (γ1 (l), . . . , γn−1 (l)). For (x, t) ∈ Rn × R+ we define 1 t ˙ u(x, t) = infx (m − 1) ln e m−1 {g(γ (0))+ 0 L(γ̄ (θ))dθ} t t m 1 1 t ˙ + |γ̇n | m−1 e m−1 θ L(γ̄ (l))dl dθ . m 0 (2.2) Our main result is the following. Theorem 2.1. u(x, t) defined in (2.2) is a Lipschitz continuous function and is a viscosity solution of (2.1). Proof. We prove the theorem by subsequent lemmas. Lemma 2.1. Let M1 ≤ g ≤ M2 for all x ∈ Rn . Then M1 ≤ u(x, t) ≤ M2 for all (x, t) ∈ Rn × R+ . Also there exists a path φ ∈ tx such that t ˙ 1 u(x, t) = (m − 1) ln e m−1 {g(φ(0))+ 0 L(φ̄(θ))dθ } t ˙ m 1 1 t + |φ̇n | m−1 e m−1 θ L(φ̄(l))dl dθ . m 0 Proof. Taking γ ≡ x in (2.2) we have u(x, t) ≤ g(x) ≤ M2 . Again u(x, t) ≥ infx g(γ (0)) ≥ M1 . t This proves the first part. Let {γk } be the minimizing sequence for (2.2). Since t 0 and 0 n−1 1 xi − (γk )i (0) 2 ˙ L(γ̄k (θ ))dθ ≥ t 2 t i=1 t x − (γ ) (0) m m k n n m−1 |(γ̇k )n (θ )| m−1 dθ ≥ t , t 352 Anup Biswas, Rajib Dutta and Prosenjit Roy we note that {γk (0)} remains in a compact set. Therefore we can get a subsequence, denoted by the same index {γk (0)}, and a path φ ∈ tx such that γk → φ uniformly in [0, t], γ̄˙k → φ̄˙ weakly in L 2 ([0, t], Rn−1 ), m (γ̇k )n → φ̇n weakly in L m−1 ([0, t], R). Again we note that t m 1 1 t ˙ |γ̇n | m−1 e m−1 θ L(γ̄ (l))dl dθ m 0 is convex in (·, . . . , ·, γn ) and (γ1 , . . . , γn−1 , ·). Since lim inf{a(n) + b(n)} lim inf a(n) + lim inf b(n), it is easy to see that φ is a minimizing path. Lemma 2.2 (Dynamic Programming Principle). For 0 ≤ s < t, we have t−s 1 L(γ̄˙ (θ))dθ} m−1 {u(γ (0),s)+ 0 (m − 1) ln e u(x, t) = inf x t−s t−s m 1 1 t−s L(γ̄˙ (l))dl θ m−1 m−1 + |γ̇n | e dθ . m 0 ≥ (2.3) Proof. Let v(x, t) denote the right-hand side of (2.3). From Lemma 2.1, we can find φ ∈ tx such that t ˙ 1 u(x, t) = (m − 1) ln e m−1 {g(φ(0))+ 0 L(φ̄(θ))dθ} t ˙ m 1 1 t L(φ̄(l))dl θ m−1 m−1 + |φ̇n | e dθ . m 0 Therefore 1 t ˙ 1 s ˙ m m−1 s 1 m−1 θ ˙ L(φ̄(l))dl t ˙ L(φ̄(l))dl 0 L(φ̄(θ))dθ} eu(x,t)/(m−1) = e m−1 {g(φ(0))+ t ˙ m 1 1 t + |φ̇n | m−1 e m−1 θ L(φ̄(l))dl dθ m 0 t ˙ 0 L(φ̄(θ))dθ+ s L(φ̄(θ))dθ} = e m−1 {g(φ(0))+ s t ˙ m 1 1 + |φ̇n | m−1 e m−1 θ L(φ̄(l))dl dθ m 0 t ˙ m 1 1 t |φ̇n | m−1 e m−1 θ L(φ̄(l))dl dθ + m s s ˙ 1 − e m−1 {g(φ(0))+ 0 L(φ̄(θ))dθ} 1 + m 1 + m s |φ̇n | e 0 t s m 1 |φ̇n | m−1 e m−1 θ 1 dθ × e m−1 dθ. t s ˙ L(φ̄(θ))dθ Value functions for HJE 353 Define ψ(θ ) = φ(θ + s) for θ ∈ [0, t − s]. Then t−s 1 ˙ L(ψ̄(θ))dθ} 0 eu(x,t)/(m−1) ≥ e m−1 {u(ψ(0),s)+ t−s t−s ˙ m 1 1 + |ψ˙n (θ )| m−1 e m−1 θ L(ψ̄(l))dl dθ m 0 ≥ ev(x,t)/(m−1) . x such that Therefore u(x, t) ≥ v(x, t). Consider > 0. Then there exists φ ∈ t−s e v(x,t)+ m−1 t−s 1 ˙ ≥ e m−1 {u(φ(0),s)+ 0 L(φ̄(θ))dθ} t−s ˙ m 1 1 t−s + |φ˙n (θ )| m−1 e m−1 θ L(φ̄(l))dl dθ. m 0 (2.4) φ(0) Again using Lemma 2.1 we have an absolutely continuous path ψ ∈ s e u(φ(0),s) m−1 = e such that s ˙ 1 m−1 {g(ψ(0))+ 0 L(ψ̄(θ))dθ} + 1 m s m 1 |ψ˙n (θ )| m−1 e m−1 s θ ˙ L(ψ̄(l))dl dθ. (2.5) 0 Now define a path χ ∈ tx by ψ(θ ), for θ ∈ [0, s], χ (θ ) = φ(θ − s), for θ ∈ (s, t]. Then using (2.4), (2.5) and the same steps as above we have e v(x,t)+ m−1 u(x,t) ≥ e m−1 , > 0 being arbitrary we have v(x, t) ≥ u(x, t). This completes the proof. Lemma 2.3. The function u defined in (2.2) is Lipschitz in Rn × R+ and limt→0 u(x, t) = g(x). Proof. Let x1 , x2 ∈ Rn , and from Lemma 2.1 there exists φi ∈ txi (i = 1, 2) such that e u(xi ,t) m−1 t 1 ˙ = e m−1 {g(φi (0))+ 0 L(φ̄i (θ))dθ} t ˙ m 1 1 t + |φ̇i n (θ )| m−1 e m−1 θ L(φ̄i (l))dl dθ. m 0 1 t ˙ Since u(x, t) ≤ M2 we note that e m−1 0 L(φ̄i (θ))dθ ≤ e x constant of e m−1 in [M1 , M2 ]. Define ψi ∈ txi by t φ̇2 dθ for θ ∈ [0, t] ψ1 (θ ) = x1 − θ and ψ2 (θ ) = x2 − θ t φ̇1 dθ for θ ∈ [0, t]. M2 +|M1 | m−1 . Let K 1 be the Lipschitz 354 Anup Biswas, Rajib Dutta and Prosenjit Roy Therefore ψ1 (0) − φ2 (0) = x1 − x2 and ψ2 (0) − φ1 (0) = x2 − x1 . Hence t 1 eu(x1 ,t)/(m−1) − eu(x2 ,t)/(m−1) ≤ e m−1 {g(ψ2 (0))+ 1 L(ψ̄˙ 2 (θ))dθ} 0 t − e m−1 {g(φ1 (0))+ 1 = e m−1 ≤ e t 0 L(φ̄˙ 1 (θ))dθ M2 +|M1 | m−1 0 L(φ̄˙ 1 (θ))dθ} g(ψ2 (0)) g(φ1 (0)) e m−1 − e m−1 K 1 K 2 |x1 − x2 |, where K 2 is the Lipschitz constant of g(·). Similarly we can show that eu(x2 ,t)/(m−1) − eu(x1 ,t)/(m−1) ≤ e M2 /(m−1) K 1 K 2 |x1 − x2 |. Therefore there exists a constant K such that |eu(x1 ,t)/(m−1) − eu(x2 ,t)/(m−1) | ≤ K |x1 − x2 |. (2.6) From (2.3) we have u(x, t) ≤ u(x, s) for 0 ≤ s < t. From Lemma 2.1 there exists x such that φ ∈ t−s u(x,t) t−s 1 ˙ e m−1 = e m−1 {u(φ(0),s)+ 0 L(φ̄(θ))dθ} t−s ˙ m 1 1 t−s + |φ̇n (θ )| m−1 e m−1 θ L(φ̄(l))dl dθ. m 0 Therefore from Lemma 2.2 we have eu(x,t)/(m−1) − eu(x,s)/(m−1) 1 t−s ˙ u(x,s) = e m−1 {u(φ(0),s)+ 0 L(φ̄(θ))dθ} − e m−1 t−s ˙ m 1 1 t−s + |φ̇n (θ )| m−1 e m−1 θ L(φ̄(l))dl dθ m 0 t−s u(φ(0),s) u(x,s) ˙ 1 ≥ e m−1 0 L(φ̄(θ))dθ e m−1 − e m−1 t−s u(x,s) ˙ 1 + e m−1 e m−1 0 L(φ̄(θ))dθ − 1 m 1 t−s + |φ̇n (θ )| m−1 dθ m 0 t−s K4 ˙ ))dθ ≥ −K 3 |x − φ(0)| + L(φ̄(θ m−1 0 m 1 t−s + |φ̇n (θ )| m−1 dθ (for some constant K 3 , K 4 ) m 0 φ̄(t − s) − φ̄(0) K 4 (t − s) ≥ −K 3 |x − φ(0)| + L m−1 t −s m t − s φn (t − s) − φn (0) m−1 + m t −s m K4 1 L( p̄) − | pn | m−1 ≥ −(t − s) sup sup p, η − m−1 m η≤K 3 p = −K 5 |t − s| (for some constant K 5 ). Value functions for HJE 355 This proves that for any s, t ∈ R+ and x ∈ Rn , |eu(x,t)/(m−1) − eu(x,s)/(m−1) | ≤ K 5 |t − s|. (2.7) Now let (x1 , s), (x2 , t) ∈ Rn × R+ . Then |eu(x1 ,s)/(m−1) − eu(x2 ,t)/(m−1) | ≤ |eu(x1 ,s)/(m−1) − eu(x2 ,s)/(m−1) | + |eu(x2 ,s)/(m−1) − eu(x2 ,t)/(m−1) | ≤ max{K , K 5 }(|x1 − x2 | + |t − s|). u Since the range of u(·, ·) is in [M1 , M2 ] and e m−1 is not 0, we can have a constant M such that |u(x1 , s) − u(x2 , t)| ≤ M (|x1 − x2 | + |t − s|) . (2.8) Lemma 2.4. u is a viscosity solution of HJE (2.1). Proof. Let (x0 , t0 ) ∈ Rn × R+ and ν be a C 1 -function such that u(x0 , t0 ) = ν(x0 , t0 ) and u − ν has a local maximum at (x0 , t0 ). Let B be a ball centered at (x0 , t0 ) such that u(x, t) ≤ ν(x, t) for (x, t) ∈ B. Let 0 ≤ t < t0 , λi > 0, h = t0 − t and x = x0 − h · λ with λ = (λ1 , . . . , λn ). Define def ¯ tx = {γ ∈ tx : γ is a straight line}. Therefore for small h we have e ν(x0 ,t0 ) m−1 = e u(x0 ,t0 ) m−1 h 1 e m−1 {u(γ (0),t0 −h)+ ≤ inf x 0 L(γ̄˙ (θ))dθ} ¯ h 0 1 + m ⎡ h |γ̇n | m m−1 e h 1 m−1 θ L(γ̄˙ (l))dl dθ 0 1 = inf ⎣e m−1 {u(y,t0 −h)+h L( x0 −y h )} y∈Rn ⎛ ⎞⎤ x0 −y m 1 m − 1 (x0 )n − yn m−1 ⎝ e m−1 h L( h ) − 1 ⎠⎦ + m h L( x0 −y ) h ≤ e 1 m−1 {u(x,t0 −h)+h L(λ̄)} m m−1 + |λn | m−1 m e 1 m−1 {ν(x,t0 −h)+h L(λ̄)} m m−1 |λn | m−1 + m ≤ 1 1 e m−1 h L(λ̄) − 1 L(λ̄) e m−1 h L(λ̄) − 1 L(λ̄) . 356 Anup Biswas, Rajib Dutta and Prosenjit Roy Therefore if ν is C 2 using Taylor’s expansion we have from above e ν(x0 ,t0 ) m−1 ≤ e ν(x0 ,t0 ) m−1 n ν(x0 ,t0 ) 1 + −λi hνxi (x0 , t0 ) − hνt (x0 , t0 ) + h L(λ̄) e m−1 m−1 i=1 1 m e m−1 h L(λ̄) − 1 m−1 2 |λn | m−1 + O(h ) + . m L(λ̄) Hence dividing both sides by h and letting h ↓ 0 we have n ν(x0 ,t0 ) 1 −λi νxi (x0 , t0 ) − νt (x0 , t0 ) + L(λ̄) e m−1 0 ≤ m−1 i=1 m 1 m−1 |λn | m−1 . + m−1 m Similarly we could choose λi ≤ 0 to get the above estimate. Hence λ being arbitrary we have ν(x0 ,t0 ) m − 1 m − m−1 m−1 ≤ 0, |λn | νt (x0 , t0 ) + sup λ, Dν(x0 , t0 ) − L(λ̄) − e m λ i.e 1 eν(x0 ,t0 ) |νxi |2 + |νxn |m ≤ 0. 2 m n−1 νt (x0 , t0 ) + i=1 Since the result can be deduced for ν in C 1 , this proves that u is a subsolution. Now we prove that u is a supersolution. Let (x0 , t0 ) ∈ Rn × R+ be a point and ν be a function on Rn × R+ such that ν(x0 , to ) = u(x0 , t0 ) and u − ν has a local minimum at (x0 , t0 ). We have to show that C1 1 eν(x0 ,t0 ) |νxn |m ≥ 0. |νxi |2 + 2 m n−1 νt (x0 , t0 ) + (2.9) i=1 If this is not true then there exists θ > 0 such that 1 eν(x,t) |νxn |m ≤ −θ ∀ (x, t) ∈ B1 |νxi |2 + 2 m n−1 νt (x, t) + (2.10) i=1 and ν(x, t) ≤ u(x, t) ∀ (x, t) ∈ B1 , (2.11) where B1 is a small closed ball around (x0 , t0 ). From (2.3) we note that for small h > 0 x there exists a path γ ∈ h 0 such that γ (l) ∈ B1 ∀ l ∈ [0, h] Value functions for HJE 357 and e u(x0 ,t0 ) m−1 = h 1 e m−1 {u(γ (0),t0 −h)+ + 1 m h 0 m L(γ̄˙ (θ))dθ} 1 |γ̇n | m−1 e m−1 h θ L(γ̄˙ (l))dl dθ . (2.12) 0 Define φ(s) = γ (s − t0 + h) for s ∈ [t0 − h, t0 ]. Then e ν(φ(t0 ),t0 ) m−1 = t0 t0 −h = 1 − e m−1 t0 t0 −h t ˙ {ν(φ(t0 −h),t0 −h)+ t 0−h L(φ̄(θ))dθ} 0 ˙ d 1 {ν(φ(s),s)+ t0 L(φ̄(θ))dθ} s e m−1 ds t ˙ 1 1 0 ˙ e m−1 {ν(φ(s),s)+ s L(φ̄(θ))dθ} [νt + Dν, Dφ − L(φ̄)]ds. m−1 From (2.10) we have ˙ νt (φ(s), s) + Dν(φ(s), s), Dφ(s) − L(φ̄(s)) − m m − 1 − ν(φ(s),s) e m−1 |φ̇n (s)| m−1 ≤ −θ, m i.e. ν(φ(s),s) 1 ˙ e m−1 [νt (φ(s), s) + Dν(φ(s), s), Dφ(s) − L(φ̄(s))] m−1 ≤− ν(φ(s),s) m θ 1 e m−1 + |φ̇n (s)| m−1 , m−1 m i.e. e ν(φ(t0 ),t0 ) m−1 1 − e m−1 ≤− t0 t0 −h + 1 m t ˙ {ν(φ(t0 −h),t0 −h)+ t 0−h L(φ̄(θ))dθ} 0 t ˙ 1 θ 0 e m−1 {ν(φ(s),s)+ s L(φ̄(θ))dθ} m−1 t0 t0 −h m 1 |φ̇n (s)| m−1 e m−1 t s 0 ˙ L(φ̄(l))dl ds. Therefore using (2.11) and (2.12) we have t0 t ˙ 1 θ 0 0≤− e m−1 {ν(φ(s),s)+ s L(φ̄(θ))dθ} . m − 1 t0 −h This is a contradiction. Therefore (2.9) is true. Hence u is a supersolution. 358 Anup Biswas, Rajib Dutta and Prosenjit Roy Remark 2.1. We note that the Hamiltonian in (2.1) is increasing in u and so the viscosity solution of (2.1) is unique. Hence u is the unique solution of (2.1) and given by (2.2). Remark 2.2. Let 1 < i < n and H1 : Ri → R+ , H2 : Rn−i → R+ be two convex homogeneous functions of order n, m > 1 respectively that vanish only at 0 and have homogeneous Legendre transformations. It is easy to note that we can give a Hopf-Lax type representation for the viscosity solution of HJE of type u t + H1 (u x1 , . . . , u xi ) + eu H2 (u xi+1 , . . . , u xn ) = 0, u(x, 0) = g(x). Let L 1 , L 2 be the Legendre transformations of H1 , H2 respectively. Then t 1 u(x, t) = infx (m − 1) ln e m−1 {g(γ (0))+ 0 L 1 (χ̇1 (θ))dθ} t + 1 m−1 t 1 L 2 (χ̇2 (θ ))e m−1 t θ L 1 (χ̇1 (l))dl dθ , 0 where χ1 (·) denotes the path corresponding to the first i coordinates of γ (·) and χ2 (·) denotes the path corresponding to the rest of the coordinates of γ (·). 3. Characterization for minimizing paths For γ ∈ tx define def J (γ ) = t 1 e m−1 {g(γ (0))+ 1 + m t |γ̇n | 0 L(γ̄˙ (θ))dθ} m m−1 e t 1 m−1 θ L(γ̄˙ (l))dl dθ . (3.1) 0 The minimizer of J defines u and from Lemma 2.1 we note that there exists a minimizer Z (·) of J in tx . In this section we deduce some characterizations for the minimizing paths that would reduce the set of controls in (2.2). Let α(·) be a smooth curve from [0, t] to R with the property that α(0) = 0 = α(t). Let τ > 0 and we define Yτ (l) = (Z 1 (l), . . . , Z n−1 (l), Z n (l) + τ α(l)). Then t ˙ 1 J (Yτ ) = e m−1 {g(Yτ (0))+ 0 L(Ȳτ (θ))dθ} 1 + m t | Ż n + τ α̇| m m−1 e t 1 m−1 θ L(Ȳ˙τ (l))dl dθ 0 First we note that if Z i (0) = xi for some i, 1 ≤ i ≤ n, then Z i (l) = xi for all l ∈ [0, t] since it is a minimizer of J . So it is better to assume that Z i (0) = xi just by removing all those i for which this does not hold. Z being a minimizer, we have dJ (Yτ ) = 0. dτ τ =0 Value functions for HJE 359 Therefore from the above we have t t ˙ 1 1 | Ż n | m−1 sign( Ż n (θ ))α̇(θ )e m−1 θ L( Z̄ (l))dl dθ = 0, 0 α being arbitrary and we have 1 1 | Ż n | m−1 sign( Ż n (θ ))e m−1 t θ L( Z̄˙ (l))dl = constant = A (3.2) almost everywhere in [0, t]. We can assume Ż n to be continuous in [0, t] and hence (3.2) holds everywhere in [0, t]. Again sign of Ż n depends on sign of A. If sign of A is negative Ż n < 0 is everywhere which readily implies that Z n is decreasing in [0, t]. Since Z n (0) = xn , we can assume that Ż n (l) = 0 for all [0, t]. Also we note that Ż n is monotone in [0, t]. Now taking logarithm on both sides of (3.2) and differentiating we get d2 Z n (θ ) = Ż n (θ )L( Z̄˙ (θ )), dθ 2 a.e. in [0, t]. (3.3) Fix i, 1 ≤ i < n, such that Z i (0) = xi . Then as before perturbing in the i-th direction we have (mG + Z n (l) − Z n (0)) Ż i (l) = m Bi , where G = e g(Z (0)) m−1 Ż n (0) a.e. in [0, t], (3.4) and Bi is some constant. This implies that Ż i is continuous in [0, t]. Now we note that if Bi = 0, Z n being a monotone function we have Ż i = 0 a.e. in [0, t] what contradicts the assumption that Z i (0) = xi . This implies that Bi = 0 and hence (mG + Z n (l)− Z n (0)) = 0 and Ż i (l) = 0 for all l ∈ [0, t]. Since (mG + Z n (l)− Z n (0)) is a monotone function, Ż i is also monotone and hence differentiable. After differentiating (3.4), we get Ż n (l) = − m Bi d2 Z i (l) , a.e. in [0, t]. ( Ż i )2 dl 2 (3.5) Therefore from (3.3) and (3.5) we get n−1 d2 Z n (l) m d2 Z i (l) = − Bi 2 dl 2 dl 2 a.e. in [0, t]. i=1 After integration we have Ż n (l) − Ż n (0) = − n−1 m Bi ( Ż i (l) − Ż i (0)), 2 i=1 i.e. n−1 i=1 Bi Ż i (l) = n−1 i=1 Bi Ż i (0) − 2 [ Ż n (l) − Ż n (0)]. m (3.6) 360 Anup Biswas, Rajib Dutta and Prosenjit Roy Now using (3.6) we have from (3.4), n−1 n−1 2 Bi Ż i (0) − ( Ż n (l) − Ż n (0)) (mG + Z n (l) − Z n (0)) = m Bi2 , m i=1 i.e. i=1 n−1 n−1 m m2 2 Bi Ż i (0) + Ż n (0) − Ż n (l) (mG + Z n (l) − Z n (0)) = Bi . 2 2 i=1 i=1 Define n−1 m Bi Ż i (0) + Ż n (0), C = 2 def def D = i=1 n−1 2 m 2 (3.7) Bi2 . (3.8) i=1 Therefore integrating the above equality we have D D mG + Z n (l) − Z n (0) − = mG + xn − Z n (0) − C C C × e− D {C(t−l)−xn +Z n (l)} . (3.9) We can rewrite (3.9) as follows: D D = Z n (l) − mG + xn − Z n (0) − −mG + Z n (0) + C C C × e− D {C(t−l)−xn +Z n (l)} = F(Z n (l)) (say). We show that F is invertible. First we note from (3.4) that Bi = G Ż i (0) ∀i. Therefore m C mG − 1 = 2 D > 0. n−1 2 i=1 G Ż i (0) + Ż n (0) m n−1 2 i=1 G Ż i (0) 2 −1 Therefore for the fix l we have C C C F (z) = 1 + mG − 1 + (xn − Z n (0)) e− D {C(t−l)−xn +z} > 0. D D This implies that F is invertible in z and the inverse depends on l, Z (0), D Z (0). Let us denote this inverse by F l,Z (0),D Z (0) and hence from (3.9) we have D . (3.10) Z n (l) = F l,Z (0),D Z (0) −mG + Z n (0) + C Value functions for HJE 361 So we define a class of function as follows: def Vtx = {F p,q : R+ × R → R : continuous paths}, (3.11) where the following conditions are satisfied. (a) l ∈ R+ , p, q ∈ Rn and (q1 , . . . , qn−1 ) = 0. (b) qn = 0 and if C, D are defined by (3.7), (3.8) respectively, by replacing Z (0), D Z (0) g( p) m−1 with p, q respectively. Let G( p, q) = e qn . Then for l ∈ [0, t] the following are satisfied: D D p,q mG( p, q) + F − pn − l, −mG( p, q) + pn + C C D = mG( p, q) + xn − pn − C C D × e− D {C(t−l)−xn +F (l,−mG( p,q)+ pn + C )} . (c) F p,q 0, −mG( p, q) + pn + CD = pn and F p,q t, −mG( p, q) + pn + (d) For all l ∈ [0, t], F p,q l, −mG( p, q) + pn + CD = pn − mG( p, q). p,q D C = xn . If (q1 , . . . , qn−1 ) = 0 then Vtx denotes all straight lines γ with γ (t) = xn . For (q1 , . . . , qn−1 ) = 0 we define V̄tx = {(x1 , . . . , xn−1 , γ ) : γ ∈ Vtx }. For qn = 0 we define V̄tx = {(Z 1 , . . . , Z n−1 , xn )}, where Z i denotes straight lines with property Z i (t) = xi . For q other than the above two cases we define V̄tx = {(Z 1 , . . . , Z n ) : Z n ∈ Vtx }, where Z i is defined as follows: t mG( p, q)qi ds, Z i (l) = xi − (mG( p, q) + Z n (s) − pn ) l if qi = 0 otherwise Z i (l) = xi . The minimizer of J is contained in V̄tx . We also note that the only straight lines that V̄tx contains is given by the first two cases. At the end of the paper we will provide an example to show that there are points (x, t) where J does not attend minimum on the straight lines. 4. Value function for HJE decreasing in u In this section we consider a Hamiltonian of the form e−u 1 | pi |2 + | p n |m , H (u, p) = 2 m n−1 i=1 for (u, p) ∈ R × Rn , (4.1) 362 Anup Biswas, Rajib Dutta and Prosenjit Roy and the HJE u t + H (u, Du) = 0 for (x, t) ∈ Rn × R+ , u(x, 0) = g(x). (4.2) Let tx be as before. For (x, t) ∈ Rn × R+ , we define t −1 ˙ def u(x, t) = infx −(m − 1) ln e m−1 {g(γ (0))+ 0 L(γ̄ (θ)dθ)} t t m 1 1 t − m−1 L(γ̇ (l))dl θ m−1 − |γ̇n (θ )| e dθ . (4.3) m 0 (We have taken the convention − ln(a) = ∞ for a ≤ 0.) Theorem 4.1. The function u defined by (4.3) is Lipschitz continuous and a viscosity solution of (4.2). Proof. Proof of the theorem is same as the proof of Theorem 2.1. For convenience we give some sketch. Lemmas 2.1 and 2.2 could be shown along the same lines. For Lemma 2. we show only the Lipschitz continuity in t. First we note that u(x, t) ≤ u(x, s) for 0 ≤ s ≤ t. Again if φ is a minimizing path for u(x, t), then t−s −1 e m−1 {u(φ(0),s)+ ˙ L(φ̄(θ)dθ)} 0 ≥e −u(x,t) m−1 −M2 ≥ e m−1 . Therefore −1 e m−1 t−s 0 ˙ L(φ̄(θ)dθ) ≥e M1 −M2 m−1 . In a similar manner we can show that u(x2 ,t) − u(x1 ,t) ≤ K |x1 − x2 | ∀ x1 , x2 ∈ Rn , e m−1 − e− m−1 For some constant K , consider u(x,s) u(x,t) u(x,s) u(φ(0),s) e− m−1 − e− m−1 t−s u(φ(0),s) ˙ 1 + e− m−1 1 − e− m−1 0 L(φ̄)dθ t−s ˙ m 1 1 t−s + |φ̇n | m−1 e− m−1 θ L(φ̄)dl dθ m 0 ≥ −K |x − φ(0)| 1 t−s ˙ t−s ˙ −1 + e m−1 {u(φ(0),s)+ 0 L(φ̄(θ)dθ)} e m−1 0 L(φ̄)dθ − 1 t−s ˙ m 1 1 t−s |φ̇n | m−1 e− m−1 θ L(φ̄)dl dθ + m 0 t−s −M2 ˙ )dθ ≥ −K |x − φ(0)| + e m−1 L(φ̄(θ 0 t−s M1 −M2 1 m m−1 m−1 |φ̇n | . +e m 0 e− m−1 − e− m−1 = Value functions for HJE 363 Rest of the proof is similar to Lemma 2.3. Proof of subsolution and supersolution would be similar to Lemma 2.4. We can also characterize the minimizer for (4.3) in the same way as §3. But our goal is to show that u defined by (4.3) is the unique viscosity solution for (4.2). 4.1 Uniqueness First we consider the HJE n−1 1 ut + (m − 1)m m−1 2(m − 1 − m 1 m−1 u) |u xi |2 + |u xn |m = 0 for (x, t) ∈ Rn × R+ , i=1 u(x, 0) = h(x). (4.4) 1 m−1 The above HJE is valid whenever m u = m − 1. Define s t − m−1 e F(t) = ds ∀ t ∈ R. 1 0 m m−1 (4.5) Therefore F(·) is an increasing C 1 function. First we show that if u is a viscosity solution of (4.2) then F(u) is a viscosity solution of (4.4) with initial condition F(g). Let (x0 , t0 ) ∈ Rn × R+ and ν be a C 1 function such that F(u) − ν has a local maximum at (x0 , t0 ) with F(u(x0 , t0 )) = ν(x0 , t0 ). Then u − F −1 (ν) has a local maximum at (x0 , t0 ). Since u is subsolution of (4.2), we have (F −1 (ν(x0 , t0 )))t + H (F −1 (ν(x0 , t0 ), DF −1 (ν(x0 , t0 )) ≥ 0. A direct calculation shows that n−1 1 νt + (m − 1)m m−1 2(m − 1 − m 1 m−1 ν) |νxi |2 + |νxn |m ≥ 0, i=1 at the point (x0 , t0 ). We note that F (F −1 (s)) = 1 m−1−m m−1 s 1 (m−1)m m−1 > 0 if s is in the image of F. This proves that F(u) is a subsolution for (4.4). Similarly we can show that F(u) is a supersolution for (4.4). We now prove that viscosity solution of (4.4) is unique whenever the value function is Lipschitz, bounded with supremum strictly less than m−1 1 . Let u and v be two viscosity m m−1 solutions (4.4) with the same initial data. Then there exists σ > 0 such that sup (u − v) ≥ σ > 0. (4.6) R n ×R + Now we consider a function : (Rn )2 × (R̄+ )2 → R defined as follows x − y2 + |t − s|2 2α 2 − λ(t + s) − (x2 + y2 ), (x, y, t, s) = u(x, t) − v(y, s) − (4.7) 364 Anup Biswas, Rajib Dutta and Prosenjit Roy where α, λ, > 0. Note that, u and v being bounded the supremum of is attended inside a compact set, say at (x̄, ȳ, t¯, s̄). Again for λ ∈ (0, λ0 ], ∈ (0, 0 ], for λ0 , 0 small, we can have σ (4.8) (x̄, ȳ, t¯, s̄) ≥ sup (x, x, t, t) ≥ , 2 R n ×R + and this implies that u(x̄, t¯) > v( ȳ, s̄). Since (x̄, ȳ, t¯, s̄) ≥ (0, 0, 0, 0), we have λ(t¯ + s̄) + x̄ − ȳ2 + |t¯ − s̄|2 + (x̄2 + ȳ2 ) ≤ C, 2α 2 (4.9) for some constant C depending on the bounds of u and v. This implies that 1 3 (x̄ + ȳ) = 4 4 (x̄ + ȳ) 1 ≤ 3 2 2 (x̄ + ȳ)2 + 2 2 1 ≤ C1 2 . (4.10) Let h(x) be the initial condition for u, v. Therefore from (4.8) we have σ ≤ u(x̄, t¯) − v( ȳ, s̄) 2 = u(x̄, t¯) − h(x̄) + h(x̄) − v(x̄, t¯) + v(x̄, t¯) − v( ȳ, s̄) ≤ K 1 t¯ + K 2 t¯ + K 2 (x̄ − ȳ + |t − s|), where K 1 , K 2 are the Lipschitz constants of u, v respectively. Using (4.9) we note that for small α > 0, t¯ > 0. Similarly we have s̄ > 0. Also note that the choice of α is independent of , λ. Now we observe that u − ν has a maximum at (x̄, t¯) for def ν(x, t) = v( ȳ, s̄) + x − ȳ2 + |t − s̄|2 + λ(t + s̄) + (x2 + ȳ2 ) + K , 2α 2 where K is a constant that ensures u(x̄, t¯) = ν(x̄, t¯). Since u is a viscosity solution we have from (4.4), λ+ t¯ − s̄ x̄ − ȳ + H1 (u(x̄, t¯), + 2 x̄) ≤ 0, α2 α2 (4.11) where 1 H1 (u, p) = (m − 1)m m−1 1 2(m − 1 − m m−1 u) n−1 × | pi |2 + | pn |m , for (u, p) ∈ m−1 1 , ∞ × Rn . m m−1 i=1 We also observe that v − ν has a minimum at ( ȳ, s̄) for def ν(y, s) = u(x̄, t¯) − x̄ − y2 + |t¯ − s|2 − λ(t¯ + s) − (x̄2 + y2 ) + K , 2α 2 Value functions for HJE 365 where K is a constant that ensures v( ȳ, s̄) = ν( ȳ, s̄). Since v is a viscosity solution of (4.4), we have t¯ − s̄ x̄ − ȳ −λ + 2 + H1 v( ȳ, s̄), − 2 ȳ ≥ 0. (4.12) α α2 Now subtracting (4.12) from (4.11) we have x̄ − ȳ x̄ − ȳ + 2 x̄ − H1 v( ȳ, s̄), − 2 ȳ ≤ 0. 2λ + H1 u(x̄, t¯), α2 α2 (4.13) As ↓ 0, we note that x̄, ȳ → 0 from (4.10). Also from (4.9), we note that x̄ − ȳ → z ∈ Rn , may be for a subsequence, as ↓ 0. u, v being bounded, as ↓ 0 let u(x̄, t¯) → ū and v( ȳ, s̄) → v̄ where ū ≥ v̄. Therefore letting ↓ 0 in (4.13) we have z z 2λ + H1 ū, 2 − H1 v̄, 2 ≤ 0. α α But this is a contradiction as the left-hand side is > 0 due to monotonicity of H1 in u. Therefore our assumption (4.6) is wrong and hence u ≤ v. Similarly we can prove that v ≤ u. This also proves that (4.2) has a unique viscosity solution. Therefore we have the following theorem: Theorem 4.2. The Hamilton-Jacobi equation (4.2) has a unique bounded, Lipschitz continuous viscosity solution and this is given by (4.3). Remark 4.1. Let H1 , H2 be the same as in Remark 2.2. Then u t + H1 (u x1 , . . . , u xi ) + e−u H2 (u xi+1 , . . . , u xn ) = 0 has a unique viscosity solution. The Hopf-Lax type representation is given by −1 t u(x, t) = infx −(m − 1) ln e m−1 {g(γ (0))+ 0 L 1 (χ̇1 (θ))dθ} t 1 − m−1 t −1 L 2 (χ̇2 (θ ))e m−1 t θ L 1 (χ̇1 (l))dl dθ . 0 Remark 4.2. Let H1 , H2 be as above. Then it is easy to check that for α = 0 the unique viscosity solution of u t + H1 (u x1 , . . . , u xi ) + eαu H2 (u xi+1 , . . . , u xn ) = 0 has representation u(x, t) = infx t (m − 1) α {g(γ (0))+ t L 1 (χ̇1 (θ))dθ} 0 ln e m−1 α t t α α + L 2 (χ̇2 (θ ))e m−1 θ L 1 (χ̇1 (l))dl dθ . m−1 0 366 Anup Biswas, Rajib Dutta and Prosenjit Roy 5. Example Consider the HJE ut + |u x1 |2 2 + eu 2 2 |u x2 | u(x, 0) = 0 in R2 × R+ ⎫ ⎬ = g1 (x1 ) + g2 (x2 ) ⎭ , (5.1) where g1 : R → R is defined by ⎧ if |x| ≥ 1, ⎨ 0, g1 (x) = 1 + x, if − 1 ≤ x ≤ 0, ⎩ 1 − x, if 0 ≤ x ≤ 1 and g2 : R → R is defined by ⎧ ⎨ 0, g2 (x) = ln(1 + x), ⎩ ln 2, Define and if x ≤ 0, if 0 ≤ x ≤ 1, if 1 ≤ x. ' ( t x1 − y1 2 u 1 (x, t) = inf g1 (y1 ) + + g2 (x2 ) 2 t y1 ∈R (5.2) ' 2 ( x − y t 2 2 . u 2 (x, t) = inf ln eg1 (x1 )+g2 (y2 ) + 2 t y2 ∈R (5.3) If the minimizer of J (defined for HJE (5.1)) is attended on straight lines for all points (x, t) ∈ R2 × R+ , then u(x, t) = min{u 1 (x, t), u 2 (x, t)} would be the viscosity solution of (5.1). We note that u 1 (x, t) = g2 (x2 ) if |x| ≥ 1. Also for 0 ≤ x ≤ 1, by direct calculation we have for t ≤ 2, ⎧ t ⎪ ⎪ ⎨ 1 − x1 − 2 + g2 (x2 ), if 0 ≤ x1 ≤ t − 1, u 1 (x, t) = 2 ⎪ ⎪ ⎩ (1 − x1 ) + g2 (x2 ), if t − 1 ≤ x1 ≤ 1. 2t Similarly for x2 ≤ 0, u 2 (x, t) = g1 (x1 ). Consider t > 0 small such that te2 < 1. By direct calculation we see that for 0 ≤ x2 ≤ 1, ⎧ ⎪ x22 ⎪ g (x ) ⎪ 1 1 + , if 0 ≤ x2 ≤ eg1 (x1 ) t, ⎨ ln e 2t u 2 (x, t) = ⎪ t g1 (x1 ) ⎪ ⎪ , if eg1 (x1 ) t ≤ x2 ≤ 1. ⎩ g1 (x1 ) + ln 1 + x2 − e 2 Now in {(x1 , x2 ) | 0 ≤ x1 ≤ 1, 0 ≤ x2 ≤ 1}, g1 , g2 = 0 and so it is easy to show that none of u 1 , u 2 is a viscosity solution. So u cannot be a viscosity solution if it is defined Value functions for HJE 367 as a minimum of u 1 , u 2 . This proves that the minimizer of J may not be attained on straight lines. Acknowledgements The authors would like to thank Prof. Adimurthi and Prof. G D Verrappa Gowda for their help and fruitful discussions. The authors would also like to thank the referee for his useful comments. This research is supported in part by the Microsoft Research India Fellowship. References [1] Adimurthi and Gowda G D V, Hopf Lax type formula for sub and supersolutions, Adv. Differential Equations 5 (2000) 97–119 [2] Bardi M and Capuzzo Dolcetta I, Optimal control and viscosity solutions of HamiltonJacobi-Bellman equations, Systems and Control: Foundations and Applications (1997) (Boston, MA: Birkhäuser Boston Inc.) [3] Barron E N, Jensen R and Liu W, Hopf-Lax type formula for u t + H (u, Du) = 0, J. Differential Equations 126 (1996) 48–61 [4] Evans L C, Partial Differential Equations, Berkeley Mathematics Lecture Notes 3A, 3B (1994) [5] Lions P L, Generalized Solution of Hamilton-Jacobi Equations, Research Notes in Mathematics (1982) (Pitman)
© Copyright 2025 Paperzz