On some connections between probability and differential equations Rafael Granero Belinchón [email protected] Master Thesis in Partial Differential Equations- Random and Deterministic Modelling. Advisor: Mr. Jesús Garcı́a Azorero Contents Introduction 5 1 First results 1.1 The brownian motion . . . . . . . . . . . . . 1.2 Existence and uniqueness theorem for SDE . 1.3 The Wiener measure . . . . . . . . . . . . . . 1.4 Markov process and semigroups of operators . 2 Elliptic equations 2.1 The laplacian . . . . . . . . . . . . . . . 2.2 Poisson equation and shape recognition 2.3 A general elliptic equation . . . . . . . . 2.4 Unbounded domains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 7 13 19 23 . . . . 29 30 35 37 39 3 Parabolic equations 41 3.1 A general parabolic equation . . . . . . . . . . . . . . . . . . 41 3.2 The Fisher equation . . . . . . . . . . . . . . . . . . . . . . . 46 3.3 Feynman and quantum mechanics . . . . . . . . . . . . . . . 48 4 Fluid dynamics 4.1 The 1-dimensional Burgers equation . . . . 4.2 The d−dimensional Burgers equations . . . 4.3 The incompressible Navier-Stokes equations 4.4 Proof of local existence for Navier-Stokes . . . . . . . . . . . . . . . . . . . . . 5 Differential games and equations 5.1 The operators . . . . . . . . . . . . . . . . . . . . . . 5.2 The games . . . . . . . . . . . . . . . . . . . . . . . . 5.2.1 ’Tug of war’ . . . . . . . . . . . . . . . . . . 5.2.2 Approximations by SDE to ∆∞ . . . . . . . . 5.2.3 Existence of game’s value for the ’Tug of war’ 5.2.4 ’Tug of war with noise’ . . . . . . . . . . . . 5.2.5 Spencer game . . . . . . . . . . . . . . . . . . 5.2.6 Other games . . . . . . . . . . . . . . . . . . 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56 57 59 64 68 . . . . . . . . 71 71 73 73 75 76 79 80 81 3 CONTENTS 6 Numerical experiments 82 7 Conclusion 85 A Some useful results A.1 A construction for the brownian motion A.2 The Kolmogorov’s regularity theorem . A.3 The Itô formula . . . . . . . . . . . . . . A.4 Existence and uniqueness for PDE . . . 87 87 88 91 95 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . B Itô integral C Matlab code C.1 Brownian motion paths . . . . . . . . . . . . C.2 Brownian bridge paths . . . . . . . . . . . . . C.3 Euler method for a SDE . . . . . . . . . . . . C.3.1 1D case . . . . . . . . . . . . . . . . . C.3.2 2D case . . . . . . . . . . . . . . . . . C.4 Monte-Carlo method for the laplacian . . . . C.5 Silhouette recognition . . . . . . . . . . . . . C.6 Monte-Carlo method for parabolic equations . C.7 Code to approximate the ∞−laplacian . . . . 96 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100 100 100 101 101 101 101 103 105 107 List of Figures 1.1 1.2 1.3 1.4 1.5 1.6 Two brownian paths. . . . . . . . . . . . . . . . A brownian path in the plane. . . . . . . . . . . Some solution of the Langevin equation paths. A solution of the 2D-Langevin equation path. . The cylinders. . . . . . . . . . . . . . . . . . . . Paths of a brownian bridge. . . . . . . . . . . . . . . . . . 9 10 13 14 20 24 2.1 2.2 Numerical experiment, silhouette and u. . . . . . . . . . . . . Results, Φ in the upper figure, Ψ in the lowe figure. . . . . . . 36 38 3.1 Traveling wave solution of (3.3). . . . . . . . . . . . . . . . . 46 4.1 4.2 4.3 4.4 4.5 Navier-Stokes solution at time 10. . . . . . . . . Stokes problem solution. . . . . . . . . . . . . . . Inviscid Burgers equation at different times. . . . Burgers equation with different dissipation rates. Flow. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56 57 58 59 61 5.1 5.2 An ∞−harmonic function. . . . . . . . . . . . . . . . . . . . . The posible positions for the ’Tug of war with noise’ game. . 72 80 6.1 6.2 6.3 The numerical solution. . . . . . . . . . . . . . . . . . . . . . Initial value. . . . . . . . . . . . . . . . . . . . . . . . . . . . . Numerical solution at time 4. . . . . . . . . . . . . . . . . . . 83 84 84 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Introduction It may appear that the partial differential equations and probability theory are very diverse fields of study. But when we study these fields in a deeper way we find multiple connections between them (representation formulas, new numerical methods...). We are going to display some of these relationships. We will mainly study them from a partial differential equations point of view. However, at the same time, we will state the probabilistic results, giving general ideas of the proofs and references for the technical details. This approach is going to permit us to resolve some problems in an easier way, or, at least, in a different way. In addition, from a numerical analysis point of view, it is useful because it will give us the opportunity to utilize the Monte-Carlo method to approach the solution of a PDE. Other applications derived from this calculations have been the functional integration, a key point in quantum theory, and a new method (we will talk about it later) for silhouette recognition ([GGSBB]). This text is formed by two parts. In the first part we will obtain the representation formulas as integrals in a certain functional space for the solution to diverse PDE’s. We will give another proof for the local (in time) existence of classical solution for Navier-Stokes equations in 3D. For this topic, we will follow closely the work of G. Iyer and P. Constantin contained in the Ph.D. thesis of the former and some papers of both of them ([CI],[Iy],[Iy2], [C]). We will also study some equation from the quantum mechanics, in particular we will go in depth into Feynman’s formulation of quantum mechanics ([FH],[Fe],[Fe2],[GJ],[S],[Z]). This method of considering the Itô diffusions can be understood as the method of characteristics, but random ones (see chapter 4). After we will take into account problems related to the infinity laplacian, with an approximation based on the game theory. For this we follow the work of Y.Peres, S.Sheffield, D.Wilson and O.Schramm ([PSSW]). Finally, the second part is dedicated to the appendix which contains technical results that complement the other chapters. We do not want to finish this introduction without explicitly mentioning the great mathematicians and physicist responsible for the development of this theory, like A. Einstein for his papers about th brownian motion ([E]), R. Feynman ([Fe],[Fe2] ,[FH]) and P. Dirac for thinking into ’integrate on 5 6 INTRODUCTION functions’ and for giving form to the third formulation of quantum theory. The rigour to these calculations was given by N. Wiener and M. Kac ([K],[K2]). The last author has a paper which inspires the title of this text. We also have to mention K. Itô, who gave us a result with his name, and S. Ulam, responsible for the Monte-Carlo method. It will be pleasent that this text will serve as humble homage. It is a courious fact to observe how small is the quantity of squared kilometers where these ideas were developed: Wiener, Feynman and Ulam they knew each other from Los Álamos, where they helped develop the atomic bomb. Kac was Feynman’s companion at the Cornell University. We also want to aknowledge various people for their contribution, specially to Mr. Jesús Garcı́a Azorero (Universidad Autónoma de Madrid), for his effort and care. Also to Mr. Rafael Orive Illera (Universidad Autónoma de Madrid), for his review of the draft version of this text, to Mr. Massimiliano Gubinelli (Université Paris-Dauphine) for his interesting explanations, to Mr. Bela Farago (Institut Laue-Langevin) for his hospitality and to Mr. Julio D. Rossi (Universidad de Buenos Aires) and to Mr. Fernando Charro (Universidad Autónoma de Madrid) for their explanations, very useful for the fifth chapter. Less formal but equally useful was the help that Ms. Eva Martı́nez Garcı́a, Mr. David Paredes Barato and Mr. Jesús Rabadán Toledo gave me. Chapter 1 First results In this chapter we give the results and definitions we will use after. We give some properties of brownian motion’s path and stochastic differential equations. In the appendix A there is a construction of brownian motion (for more properties of this object see [Du]). In this chapter we define and construct the Wiener measure and we study the relation between semigroups of operators and Markov processes. 1.1 The brownian motion The brownian motion is the random motion we can observe in certain microscopic particles in a fluid (for example, pollen suspended in water). Robert Brown observed this highly irregular motion in 1827. The motion of this particles is because their surface is randomly hitted with random force by the fluid molecules. The diffusion is a phenomenon based in the brownian motion. The first person who described mathematically the brownian motion was Thorvald N. Thiele in 1880, in a paper about the least squares. Louis Bachelier in his Ph.D. thesis in 1900 give a stochastic approximation to market’s fluctuations. However it was Einstein who, in 1905, studied this phenomenon and rederived and extended the previous results. In those years the atomic and molecular nature of matter was a controverted idea. Einstein and Marian Smoluchowski proved that if the cinetic fluid theory was correct then the water molecules would have random motions. Let a 2-dimensional grid (for space and time) {(ndx, mdt), m, n ∈ Z} with increments dx and dt. Let a particle which starts in time 0 in x = 0. The probability of a movement towards the right for this particle is 1/2. The probability of a movement towards the left is the same. Our particle automatically will move up (the vertical axis is for the time). As we said before, our model is a model of the position of a particle with random motion 7 8 CHAPTER 1. FIRST RESULTS caused by the random hits. Let p(n, m) be the probability, for this particle, of being in the position ndx in time mdt. Using conditional probabilities, we have 1 p(n, m + 1) = (p(n − 1, m) + p(n + 1, m)); 2 thus, 1 p(n, m + 1) − p(n, m) = (p(n − 1, m) − 2p(n, m) + p(n + 1, m)) 2 If we suppose dx2 =D>0 dt (1.1) we can write p(n, m + 1) − p(n, m) D (p(n − 1, m) − 2p(n, m) + p(n + 1, m)) = dt 2 dx2 The quotient condition we suppose in (1.1) is needed to obtain a parabolic equation, if we consider a different condition the resulting limit would not make any sense. Formally, assuming the limits that we take exist; doing dx, dt → 0 with (1.1) and writing ndx = x, mdt = t, our discrete probability converges to a density. p(n, m) → f (x, t) And we obtain that the density verifies the heat equation with diffusion parameter D/2. ∂t f (x, t) = D ∆f (x, t), f (x, 0) = δ0 (x) 2 (1.2) The hypothesis (1.1) is a key point and gives us the diffusion equation, as we expect because the model we consider. These calculations are formal. Indeed, the previous limit is not rigorous. However we can rigorize using the central limit theorem. This theorem shows us that the density has a normal distribution N (0, Dt). All these calculations are justified in [Ev]. Einstein in [E] studied this problem. Our formal arguments show us that there is a relationship between probability and PDE’s. We give a definition and some results about the brownian motion. Definition 1. Let (Ω, B, P ) a probability space, we say that a process W (ω, t)1 , W : Ω × [0, T ] → R is a brownian motion if the following conditions hold ~ (t). However, to remark We will use the following notation for the brownian motion W the idea of the brownian motion like a random variable taking values in a functional space, we will write W (ω) ∈ C([0, T ]) o ω(t). 1 9 1.1. THE BROWNIAN MOTION 1. W (ω, 0) = 0 y t 7→ W (ω, t) is continuous a.e 2. W (ω, t) − W (ω, s) ∼ N (0, t − s) ∀t ≥ s > 0 3. We have independent increments. 2 1.5 1 0.5 0 −0.5 −1 −1.5 0 200 400 600 800 1000 1200 Figure 1.1: Two brownian paths. Let t1 , t2 ..., tn times and B1 , ...Bn intervals. We can calculate the probability that a brownian path takes values in Bi at time ti . Indeed, let −|x − y|2 1 exp p(t, x, y) = √ 2t 2πt P (a1 < W (t1 ) < b1 , ...an < W (tn ) < bn ) = Z ... B1 Z Bn p(t1 , 0, x1 )p(t2 − t1 , x1 , x2 )...p(tn − tn−1 , xn−1 , xn )dxn ...dx1 (1.3) This calculation is a key point for the Wiener measure. Using a standard argument, if we have the formula for step functions, we can generalize the formula by approximation. We obtain E[f (W (t1 ), ..., W (tn ))] = Z Rn f (x1 , ..., xn )p(t1 , 0, x1 )p(t2 − t1 , x1 , x2 )...p(tn − tn−1 , xn−1 , xn )dxn ...dx1 (1.4) Remark 1 We will see that (1.4) can be understood as an ’integral in functions’. Indeed, if we fix T we can see the brownian motion as W (ω) : 10 CHAPTER 1. FIRST RESULTS Ω 7→ C([0, T ]), and then f will be a function with another function as argument. Using the definition we conclude E[W (t)] = 0, E[W 2 (t)] = t We can calculate the covariance in a similar way. If s < t then E[W (t)W (s)] = E[(W (s)+W (t)−W (s))W (s)] = s+E[(W (t)−W (s))W (s)] = s 2 1.5 1 0.5 0 −0.5 −1 −1.6 −1.4 −1.2 −1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 Figure 1.2: A brownian path in the plane. Thinking in the applications to differential equations, we are interested in the path properties. We give a theorem of Kolmogorov about the regularity of the paths. Theorem 1 (Kolmogorov). Let X be a stochastic process with continuous path a.e. and such that E[|X(t) − X(s)|β ] ≤ C(t − s)1+α , ∀t, s ≥ 0 then for all 0 < γ < α β and T > 0 there exists K(ω) such that |X(t) − X(s)| ≤ K|t − s|γ We check the hypothesis hold in the case of the brownian motion. Let t > 0, then Z 1 2m |x|2m exp(−|x|2 /2(t − s))dx E[|W (t) − W (s)| ] = p 2π(t − s) R Z (t − s)m √ = |y|2m exp(−|y|2 /2)dy 2π R = C|t − s|m 11 1.1. THE BROWNIAN MOTION where we do the natural change of variables, y=√ x . t−s (1.5) The change of variables is natural if we see the hypothesis (1.1). The hypothesis is fulfilled with β = 2m and α = m − 1. Thus γ < αβ = 1 1 1 2 − 2m for all m and we conclude γ < 2 . We prove the brownian motion has Hölder paths in [0, T ] with exponent γ < 1/2. This result is optimal in the sense that any other γ ≥ 21 holds. Indeed, if we have a Hölder estimation with γ = 1/2 then sup 0<s<t<T |W (t) − W (s)| ≤ C(ω) c.t.p. |t − s|1/2 (1.6) An inequality like the previous one is not possible, because of if we consider a partition 0 = t1 < t2 ... < tn = T , we have sup 0<s<t<T |W (ti+1 ) − W (ti )| |W (t) − W (s)| ≥ sup 1/2 (t − s) (ti+1 − ti )1/2 i We bound the original expression by independent and identically distributed (with standard normal ditributions) random variables (we fix the times), and then we can calculate explicitly the probability that the supremum is greater or equal than a parameter L. n |W (ti+1 ) − W (ti )| |W (t2 ) − W (t1 )| P sup → 1, if n → ∞ ≥ L = 1−P ≥L (ti+1 − ti )1/2 (t2 − t1 )1/2 i We want to point out this stochastic process has not bounded total variation. If we have bounded total variation, given a partition, we will have n X i=0 |W (ti+1 ) − W (ti )|2 = max(|W (ti+1 − W (ti )|) i n X i=0 |W (ti+1 − W (ti )| ≤ V (0, T ) max(|W (t) − W (s)|) i and the last expression tends to zero because of the continuity of the paths when we refine the partition. The contradiction is that the quadratic variation of the brownian motion is greater than zero, and then V (0, T ), the total variation, can not be bounded. We have proved the following result: Theorem 2. The brownian motion has Hölder continuous paths with exponent γ < 21 . This exponent is optimal. In particular the paths are not of bounded variation and they are nowhere differentiable a.e.. 12 CHAPTER 1. FIRST RESULTS A very important property of the stochastic processes we study is the Markov property, which tell us the process has not memory of the previous history. Or more precisely Definition 2. Let X(t) be a stochastic process. It is a Markov process if the following condition holds, given Fs the filtration generated by the process (the history, see appendix B), P [X(t) ∈ B|Fs ] = P [X(t) ∈ B|X(s)] a.e, ∀t > s This is, supposing we can define the process X starting in x, we have the process start again every time, without remembering where it was previously. This is that the process X(t + s) has the same distribution of the process X started in X(s). Theorem 3. The brownian motion, W , is a Markov process. Proof. We see that X(t) = W (t + s) − W (s), t ≥ 0 is an independent (of W (t), 0 ≤ t ≤ s) brownian motion (started in the origin) because of the third property in the definition of the brownian motion.2 And then W (t + s) is a brownian motion started in W (s). See [MP],[Du] for more properties. We continue with our problem, our model of a particle suspended in a fluid and impacted randomly. Our first model, the brownian motion, has a velocity defined nowhere and has not bounded total variation in any interval. This is a problem from a phisical point of view. We are going to present another model. We study the velocity and not the position of the particle. Let v(t) be the velocity of the particle.The forces are the friction, which is proportional to the velocity, and a random term which is the random hits. We will write the random term as dW dt and we will call it white noise. In some sense, if we are modelling the velocity, this function will be the ’differential’ of the brownian motion (see [Ev]). Using the second Newtonian law, and writing the friction as −av(t)dt, we have dv(t) = −av(t)dt + bdW, v(0) = v0 . (1.7) This is the Langevin equation. The position is dx(t) = v(t), x(0) = x0 . This is the Ornstein-Uhlenbeck equation. Formally, we can solve the Langevin equation as it would be an ODE and we can write its solution Z t −at e−a(t−s) dW (s). (1.8) v(t) = e v0 + b 0 2 Given two processes X(t), Y (t) we say they are independent if for all times (ti , si ) we have the random vector (X(t1 ), .., X(tn )) is independent of the random vector (Y (s1 ), ..., Y (sn )). 13 1.2. EXISTENCE AND UNIQUENESS THEOREM FOR SDE Rt We have to define the term 0 e−a(t−s) dW (s). This is not a Riemann integral, this is a stochastic integral in the Itô sense (see [Ev], [Du], [MP] and appendix B).3 10 8 v(t) 6 4 2 0 −2 0 1 2 3 4 5 t 6 7 8 9 10 Figure 1.3: Some solution of the Langevin equation paths. The ideas are that the object defined (1.7) (a stochastic differential equation) do not make any sense, this way of writing is only a formal one. Only it makes sense when we write it in a integral way (and when we define the Itô integral) v(t) = v0 + Z 0 t −av(s)ds + Z t bdW (s). (1.9) 0 These solutions are stochastic processes and then, they are random processes. We can understand the SDE like an ODE at each ω. In the following section we study when a problem like (1.7) is well possed and what properties have the solution paths. 1.2 Existence and uniqueness theorem for SDE ~ 0 be a random variable, and W ~ a independent of X ~ 0 random motion. Let X The σ−algebra we consider is the generated by the initial random variable and the brownian motion, i.e.4 ~ 0, W ~ (s) 0 ≤ s ≤ t}. F(t) = Σ{X 3 There is other important fashion to define this kind of integrals, the Stratonovich integral (see appendix B). 4 We write Σ{A} the σ−algebra generated by A. We save σ for the difussion matrix. 14 CHAPTER 1. FIRST RESULTS 2.5 2 1.5 1 0.5 0 −0.5 −0.5 0 0.5 1 1.5 2 Figure 1.4: A solution of the 2D-Langevin equation path. We consider two functions ~b : Rd × [0, T ] → Rd σ : Rd × [0, T ] → Md×m where Md×m is the d × m matrix space. The brownian paths are not smooth (in time) so we can not expect the paths of the solution of a SDE will be smooth (in time). Definition 3. A function f (s, ω) is progressively measurable if is measurable in the set [0, T ]×Ω with respect to B×F, the minimum σ−algebra in [0, T ]×Ω which contains the sets A × B with A in [0, T ] and B in Ω. Definition 4. For the processes f (s, ω) we define the following spaces 1 L ([0, T ]) = {f (s, ω), E Z T 0 |f (s)|ds < ∞}. For a general p we consider p L ([0, T ]) = {f (s, ω), E Z T 0 p |f (s)| ds < ∞}. ~ Definition 5. A stochastic process X(t) is a solution of the SDE ~ = ~b(X, ~ t)dt + σ(X, ~ t)dW ~ , X(0) ~ ~0 dX =X if the following conditions hold (1.10) 1.2. EXISTENCE AND UNIQUENESS THEOREM FOR SDE 15 ~ 1. X(t) es progressively measurable. ~ 2. ~b(X(t), t) ∈ L1 [0, T ]. ~ 3. σ(X(t), t) ∈ L2 [0, T ]. 4. ~ ~0 + X(t) =X Z t ~b(X(s), ~ s)ds + 0 Z t 0 ~ ~ c.t.p. ∀ 0 ≤ t ≤ T σ(X(s), s)dW (1.11) To consider the first order equations is not a restriction, because a n−order equation can be written as n first order equations. To prove existence and uniqueness we use the sucessive aproximations method, exactly the same as in the ODE case. The existence and uniqueness theorem is Theorem 4 (Existence and uniqueness). Let ~b and σ be Lipschitz functions in the spatial variable and for all times in the [0, T ] interval i.e. |~b(x, t) − ~b(x′ , t)| ≤ L1 |x − x′ |, ∀ 0 ≤ t ≤ T |σ(x, t) − σ(x′ , t)| ≤ L2 |x − x′ |, ∀ 0 ≤ t ≤ T ~ 0 be an independent of the brownian motion considered random variable Let X 2 in L [0, T ]. Then there is an unique process in L2 [0, T ] such that is a solution of (1.10).5 We need some further results before the proof of the main theorem. We give only the statement of the results, without proof (see [O] and [Ev]). Lemma 1 (Gronwall inequality). Let φ be a nonnegative function supported in 0 ≤ t ≤ T , and C0 , A be two constants. If we have Z t Aφ(s)ds ∀ 0 ≤ t ≤ T φ(t) ≤ C0 + 0 then φ(t) ≤ C0 exp(At). Theorem 5 (Martingale inequality). Let X be a martingale and 1 < p < ∞, then the following inequality holds p p p E(|X(t)|p ). E max |X(s)| ≤ 0≤s≤t p−1 5 In the sense that if there are two solutions they are equal a.e.. 16 CHAPTER 1. FIRST RESULTS Lemma 2 (Chevichev inequality). Let X be a random variable, then for all λ > 0 and p ∈ [1, ∞) the following inequality holds P (|X| ≥ λ) ≤ E(|X|p ) λp Lemma 3 (Borel-Cantelli). We write An i.o. for the set of elements which appear in An infinitely often. If ∞ X n=1 P (An ) < ∞ then P (An i.o.) = 0 We give now the proof of the existence and uniqueness theorem. ~ yX ~ ′ be two solutions. Then, subProof of theorem 4. (Uniqueness) Let X stracting them ~ −X ~ ′ (t) = X(t) Z t 0 ~b(X(t), ~ ~ ′ (t), t)dt + t) − ~b(X Z t ~ ~ ′ (t), t)dW ~. σ(X(t), t) − σ(X 0 Thus 2 Z t ′ 2 ′ ~ ~ ~ ~ ~ ~ E(|X(t) − X (t)| ) ≤ 2E b(X(t), t) − b(X (t), t)ds 0 Z t 2 ′ ~ ~ ~ + σ(X(t), t) − σ(X (t), t)dW 0 We use Cauchy-Schwarz and the Lipschitz conditions to bound each term. 2 2 Z t Z t ′ ′ ~ ~ ~ ~ ~ ~ ~ ~ E b(X(t), t) − b(X (t), t)ds ≤ TE b(X(t), t) − b(X (t), t) 0 0 Z t 2 ~ ~ ′ (t)|2 ) E(|X(t) −X ≤ L T 0 To bound the second quantity we use the Itô integral properties ([Ev],[Du]). 2 2 Z t Z t ′ ′ ~ ~ ~ ~ ~ ds σ(X(t), t) − σ(X (t), t)dW E σ( X(t), t) − σ( X (t), t) = E 0 0 Z t ~ ~ ′ (t)|2 )ds ≤ L2 E(|X(t) −X 0 1.2. EXISTENCE AND UNIQUENESS THEOREM FOR SDE 17 Considering the two previous inequalities Z t ′ 2 ~ ~ ′ (t)|2 )ds ~ ~ E(|X(t) −X E(|X(t) − X (t)| ) ≤ C 0 We use the Gronwall inequality with ~ ~ ′ (t)|2 ), φ(t) = E(|X(t) −X C0 = 0 ~ and X ~ ′ are the same a.e. for all time. and we conclude that X (Existence) We consider the approximations Z t Z t n+1 n ~ ~ n (s), s)dW ~ ~ ~ σ(X X (t) = X0 + b(X (s), s)ds + 0 0 We use the following result (it proof, based in mathematical induction, can be seen in [Ev]): Let the ’distance’ ~ n+1 (t) − X ~ n (t)|2 ) dn (t) = E(|X Then (M t)n+1 ∀ n = 1, ..., 0 ≤ t ≤ T (n + 1)! ~ 0 ). for some constant M = M (L, T, X We have, for the previous calculations, that Z T n+1 n 2 ~ n (t) − X ~ n−1 (t)|2 dt ~ ~ |X max |X (t) − X (t)| ≤ L T 2 0≤t≤T 0 Z t 2 n n−1 ~ ~ ~ + max 2 σ(X (s), s) − σ(X (s), s)dW dn (t) ≤ 0≤t≤T 0 We use the theorem 5 and the previous result and we conclude Z T n+1 n 2 ~ n (t) − X ~ n−1 (t)|2 dt ~ ~ |X E[ max |X (t) − X (t)|] ≤ L T 2 0≤t≤T 0 Z T 2 ~ n (t) − X ~ n−1 (t)|2 dt + 8L |X 0 (M T )n ≤ C n! Using Chevichev inequality and Borel-Cantelli lemma we conclude 1 n+1 n ~ ~ P max |X (t) − X (t)| > i.o. = 0 0≤t≤T 2 ~ n converges uniformly in [0, T ] to a process X. ~ Then in almost all ω, X n+1 ~ Passing to limits in X definition and in the integrals we conclude that the limit process is the (1.11) solution. See [Ev] and [O] for the proof of the L2 belonging. This proof is based in the X n+1 (t) definition and in the exponential series. 18 CHAPTER 1. FIRST RESULTS A stochastic differential equation is a generalization of an ordinary equation, so we expected similarities in the proof. However, as the brownian motion is nowhere differentiable, we can not expect that a solution of a SDE will be differentiable. The brownian smoothness is the maximum smoothness expected for a solution, i.e. Hölder-α with 0 < α < 1/2 in time. If the hypothesis of the theorem hold we have Hölder-β with 0 < β < 1 in space. To introduce the idea of a stochastic flow, which is the random version of the ODE deterministic flow, we use a new parameter, s, the initial time, ~ t (x) for the solution of and we write X s ~ ~ ~ ~, dX(t) = ~b(X(t), t)dt + σ(X(t), t)dW ~ X(s) =x (1.12) ~ ut (X ~ su (x)) = X ~ st (x) a.e. ∀ 0 ≤ s ≤ u ≤ t ≤ T, ∀x ∈ Rd X (1.13) We have the flow property The proof of this claim can be seen in the M.Gubinelli course notes in his homepage or in [Ku]. We need an inequality to use the Kolmogorov theorem to be able to study the regularity in the parameters. In [BF] we can see ~ st (x) − X ~ t′′ (x′ )|p ] ≤ C[|x − x′ |p + |s − s′ |p/2 + |t − t′ |p/2 ]. E[|X s (1.14) If we consider x = x′ and we want to know the Hölder exponent in time we use the Kolmogorov theorem (theorem 1) in the same fashion as we did in the previous section. We conclude that, if we see the solution a function in s (or in t) the Hölder exponent is γ < 1/2. To see that in space is similar. We consider s = s′ , t = t′ . We use the Kolmogorov theorem and we conclude that the exponent is γ < 1. Given the inequality (1.14), we proof the following statement Theorem 6 (Regularity). Given a stochastic differential equation with co~ st (x) be it solution. efficients holding the hypothesis of the theorem 4. Let X Then the following statements hold ~ t (x) is Hölder-γ with γ < 1/2. 1. s 7→ X s ~ st (x) is Hölder-γ with γ < 1/2. 2. t 7→ X ~ t (x) is Hölder-γ with γ < 1. 3. x 7→ X s If the functions ~b and σ are more regular in space then we have more regularity in space. Indeed, we have the following result Theorem 7. Let the coefficients of a SDE be C k,α in x functions, then the solution X0t (x) is C k,β in x with β < α. We win nothing in time because of the brownian motion is an obstacle. See [Ku] for the proof of this theorem. 19 1.3. THE WIENER MEASURE Theorem 8. Given a stochastic differential equation with coefficients holding the hypothesis of the theorem 4. Then there exists c constant such that two solutions with different initial values satisfy the following inequality ~ 1 (t) − X ~ 2 (t)|2 ] ≤ |x1 − x2 |2 ect . E[|X Proof. We apply the Itô formula (see appendix A) to the norm function, ~ 1 (t), X ~ 2 (t)) = ρ2 (X d X (X1i (t) − X2i (t))2 . i=1 Then, we apply Gronwall inequality. There are two types of stochastic equations, they are based in different stochastic integrals. They are the Itô equations, who are based in the Itô integral and they are the equations we use, and the Stratonovich. They are based in the Stratonovich integral. See the appendix B for more details. The stochastic equations can be understood as generalizations of Langevin equation for a particle suspended in a fluid and randomly hitted. They are diffusion6 . Thinking in the solutions as diffusions we can expect that they will be Markov process. See [O] for the proof. We will see that a Markov process gives us a semigroup of operators. But we need the Wiener measure. 1.3 The Wiener measure In this section we study the Wiener measure, what is needed to define the Markov semigroups in the next section. The Wiener measure is induced by the brownian motion, if we under~ (ω, t), but like stand the brownian motion not like a function W ~ : Ω 7→ C([0, T ], Rd ) W Then we think in the brownian motion like a random variable with values in a functional space. Indeed, let x be a point in Rd , then we consider the following spaces Cx ([0, T ], Rd ) = {f ∈ C([0, T ], Rd ), f (0) = x} Cxy ([0, T ], Rd ) = {f ∈ C([0, T ], Rd ), f (0) = x, f (T ) = y} (1.15) (1.16) The function ~b is the drift, σ is the diffusion term and the (1.10) solutions are Itô diffusions. 6 20 CHAPTER 1. FIRST RESULTS We can define a measure in the path space (1.15) if we consider a brown~ (t) = x+W ~ (t). ian motion who started in x. Or we can consider the process V For the second space, (1.16), the measure is called pinned because we impose the final point. To define both in a rigorous fashion we do the same.7 To simplify, we consider the one dimensional case (d = 1) y x = 0.8 We consider the following sets (we call them cylinders).9 Given times t1 , ...tn and Borel sets in R B1 , ..., Bn we define the sets 1 ,...,Bn ΠB t1 ,...,tn = {f ∈ C0 ([0, T ], R), f (ti ) ∈ Bi } (1.17) We have to give them a probability in a correct fashion, and it is now when the previous calculation (1.3) becomes useful because of we give them the probability using it. 0.6 0.4 B1 0.2 0 B2 B3 −0.2 −0.4 B4 −0.6 −0.8 0 50 100 150 200 250 300 t Figure 1.5: The cylinders. 7 A integration (in y) is the difference between them (see chapter 3) We write W0 = W. 9 There are some ways to define the Wiener measure. We consider the cylinders and the Kolmogorov extension theorem. See [GJ] for other proof. 8 350 21 1.3. THE WIENER MEASURE ,...,Bn W(ΠtB11,...,t ) n = Z ... B1 Z Bn −|xn −xn−1 | −|x1 | 1 √ e 2t1 ...e 2(tn −tn−1 ) dxn ...dx1 ( 2πt)n (1.18) We see that for some time our Borel set is the whole space then this time does not count, i.e. if in ti we have Bi = R then B ,...B ,B ,...,B n 1 i−1 i+1 1 ,...,Bn W(ΠB t1 ,...,tn ) = W(Πt1 ,...,ti−1 ,ti+1 ,...,tn ) This is a consequence of Chapman-Kolmogorov equations. If p(t, x, y) = √ 1 exp(−(x − y)2 /2t) 2πt then the Chapman-Kolmogorov equations can be written this way Z p(s, x, z)p(t, z, y)dz p(s + t, x, y) = (1.19) R i.e., the probability of going from x to y in s + t is equal to the probability of going from x to z in s and going from z to y in t after and if we consider all possible z. In order to use the Kolmogorov extension theorem we have to see that similar set have the same measure. We have to see that if A1 ,...,Am 1 ,...,Bn ΠB t1 ,...,tn = Πs1 ,...,sm then ,...,Bn ,...,Am ) W(ΠtB11,...,t ) = W(ΠsA11,...,s m n This problem can be reduced to the case when one set of times and Borel sets are contained in the other. Indeed, if the sets are the same we can consider the intersection of both and une of them, i.e. we have the following case B1 ,...,Bn,A1 ,...,Am 1 ,...,Bn ΠB t1 ,...,tn = Πt1 ,...,tn ,s1 ,...,sm And now we apply the property we talk before. If the sets are equal, in the new times sj in the right-hand side of the previous equation we have that the Borel sets, Aj , are the whole space. If it is not the case, there exists a continuous functions such that at each previous and following times ti takes value in the correct Borel set and in time sj takes value in Acj . So this is a contradiction because of the sets are not the same. When the Borel sets are the whole space we can conclude that the measure of both sets are equal using the Chapman-Kolmogorov equations. We use the Kolmogorov extension theorem, because of the consistency conditions becomes true. Moreover, the measure is countably aditive in the cylinders. Indeed, we can write N ∞ W(∪∞ i=1 Ci ) = W(∪i=1 Ci ) + W(∪i=N +1 Ci ) 22 CHAPTER 1. FIRST RESULTS and we can observe that it is correct for a finite number of sets. We conclude seeing that the second term tends to zero when N becomes greater because of the set converging to the empty set. Recall that we only have to take care about different sets at the same time. If we have different times our sets are not intersection-free. Thus for a finite number of sets the additive property is valid because of the properties of the integral. So we have a measure in the σ−algebra generated by cylinders. However it is not clear which σ−algebra is. Let the set A = {φ, f (s) < φ(s) < g(s)} for certain functions given, f, g. Let si be the rational numbers in the [0, T ] interval. Then we can write the set as A= ∞ \ [ {f (si ) + 1/n < φ(si ) < g(si ) − 1/n}. n=1 si So we have a countable union of intersections of cylinders. Sets like the previous one are in the σ−algebra, and so the Borel sets (with respect to the uniform norm) are. Indeed, given ψ it suffices to take f = ψ − ε and g = ψ + ε. Moreover, it can be shown that both σ−algebras are the same. We write ~ (ω, t) = ω(t). W Then this measure gives us an expectation Z f (ω)dW. E0 [f ] = (1.20) C0 ([0,T ],R) We consider a brownian motion which started at the origin, so we write the index zero. If we consider brownian motions which started at x we write Z f (ω)dWx . (1.21) Ex [f ] = Cx ([0,T ],R) This measure is supported on the continuous functions which at time zero were at the point x. If we fix some times, then we can change the functional integral by some integrals in Rd .10 This is a consequence of the definition of the cylinders’ measure. Z Y n ~ (t1 )), ..., fn (W ~ (tn ))] = fj (xj )p(tj −tj−1 , xj , xj−1 )dx1 dx2 , ...dxn E[f1 (W Rn j=1 With a time this is a well known formula. This is the heat semigroup. Thus if 1 H0 = − ∆ 2 10 We call path integral to the functional integral. 1.4. MARKOV PROCESS AND SEMIGROUPS OF OPERATORS −tH0 e f (x) = Z 23 ~ (t))] = E0 [f (x + W ~ (t))] (1.22) p(t, x, y)f (y)dy = Ex [f (W R This is the first representation formula we obtain. We can define the measure with respect to this operators. ~ (t1 )), ..., fn (W ~ (tn ))] = [e−t1 H0 f1 e−(t2 −t1 )H0 f2 ...e−(tn −tn−1 )H0 fn ](0) E0 [f1 (W We define the Wiener measure like the induced by the brownian motion, but we can do the same for other process. For example, the measure in (1.16) is not induced by the brownian motion, but by the brownian bridge ~ defined by this formula X, ~ (t)) ~ ~ (t) + t (y − W X(t) =W T We can define the brownian bridge like the solution of this SDE ~ dX(t) = ~ t − X(t) ~ dt + dW T −t with the same initial point as the brownian motion. Thus we can define the measure (with measure of the total space equal to p(T2 − T1 , x, y)) induced by the brownian bridge who in T1 is at x and in T2 is at y with certain operators. Z Cxy ([T1 ,T2 ],R) ~ 1 ), f2 (X(t ~ 2 ), ...fn (X(t ~ n )dW x,y f1 (X(t [T1 ,T2 ] = = [e−(T1 −t1 )H0 f1 e−(t2 −t1 )H0 f2 ...e−(tn −tn−1 )H0 fn e−(T2 −tn )H0 (·, y)](x) (1.23) The diffusions give us measures in the continuous functions space with respect to the σ−algebra defined by the cylinders11 . See [F] for more details. 1.4 Markov process and semigroups of operators Let U ⊂ Rd be a domain. And given a Markov process X we define the family of operator ~ t (x))] = Tt f (x) = Ex [f (X 0 11 Z f (y)P (t, x, dy) (1.24) Rd Actually all stochastic process almost everywhere continuous give us a measure with respect to the cylinders’ σ−algebra. We can generalize this idea to the well-posed (existence, uniqueness, continuity of the solution) ODE problems. In this case the measure is singular, the measure is supported in the solution of the ODE. 24 CHAPTER 1. FIRST RESULTS 2 1.5 1 0.5 0 −0.5 −1 −1.5 −2 0 200 400 600 800 1000 1200 Figure 1.6: Paths of a brownian bridge. where P (t, x, Γ) is the transition function, this function gives us the probability of arriving to Γ at time t for our process X started at x, i.e. Z t ~ p(t, x, y)dy P (X0 (x) ∈ Γ) = Γ and p(t, x, y) is the transition density. If f ≥ 0 then Tt f ≥ 0. Moreover if f ∈ L∞ we have Tt f ∈ L∞ and ||Tt f ||∞ ≤ ||f ||∞ . Because of the Markov property and the conditional expectation properties Tt is a semigroup12 (see [Ev]), ~ t+s (x))|Σ(X(z), ~ ~ t (X ~ s (x)))] = Tt f (X ~ s (x)). E[f (X 0 ≤ z ≤ s)] = E[f (X s 0 0 0 Taking expectations and appliying the properties of the conditional expectation we conclude Ts+t f (x) = Ts Tt f (x). So Tt is a contraction semigroup in L∞ . We define the domain as D(Tt ) = {f ∈ L∞ , ||T f − f || → 0, if t → 0}. 12 There are many formulations of the Markov property. We use before a probability ~ is a based definition, but now we use a expectation based one (see [Du],[F]). We say X Markov process if ~ + h)|Σ{X(s), ~ ~ + h)|X(t)]. ~ E[X(t 0 ≤ s ≤ t}] = E[X(t 1.4. MARKOV PROCESS AND SEMIGROUPS OF OPERATORS 25 This set is a vector space (by the linearity of the expectation and the triangle inequality). Moreover it is closed. Indeed, we have to observe that if fn ∈ D(Tt ) → f, Tt fn → Tt f then ||Tt f − f ||∞ ≤ ||Tt f − Tt fn ||∞ + ||Tt fn − fn ||∞ + ||fn − f ||∞ → 0. So f ∈ D(Tt ). We have that t 7→ Tt f is a continuous function. Indeed, ||Tt+h f − Tt f ||∞ ≤ ||Tt (Th f − f )||∞ ≤ ||Th f − f ||∞ → 0. We have shown the following result Theorem 9 (Semigroup). Given a Markov process, we have that 1. Tt is contraction semigroup in L∞ , with D(Tt ) = {f ∈ L∞ , ||Tt f − f || → 0, si t → 0} as domain. Moreover it is a closed vector space. 2. s 7→ Ts is a continuous map. We give some examples in one spatial dimension. Example 1: We consider the ODE, written to conserve the previous notation as dXt (x) = bdt. Then, there is not randomness and so we can forget the expectations. In this case the operator is Tt f (x) = Ex [f (X0t (x))] = f (x + bt). We see that it is a solution of the transport equation ut = bux , u(0, x) = f (x). In this case the semigroup’s generator is A=b ∂ . ∂x 26 CHAPTER 1. FIRST RESULTS Example 2: We consider now the stochastic equation dX = dW, X(0) = x with solution X0t (x) = x + W (t). We saw that Tt f (x) = Ex [f (x + W (t))] solves the heat equation with diffusion parameter equal to 1/2, 1 ut = uxx , 2 u(x, 0) = f (x) In this case the semigroup’s generator is A= 1 ∂2 . 2 ∂x2 Example 3: We can consider the stochastic equation dX = bdt + dW, X(0) = x and then the semigroup is associated to 1 ut = bux + uxx . 2 And the semigroup’s generator is A=b 1 ∂2 ∂ + . ∂x 2 ∂x2 In general we see that the semigroup’s generator, Th f (x) − f (x) h→0 h A = lim is the elliptic operator who appears in (A.3)13 Au = d d X ∂u ∂2u 1 X bi (x) ai,j (x) + . 2 ∂xi ∂xj ∂xi i,j=1 13 i=1 If we consider smooth enough functions (see [F] or [Dy]). 1.4. MARKOV PROCESS AND SEMIGROUPS OF OPERATORS 27 To show it we take expectations in the Itô formula applied to f (x) ∈ C 2 with bounded derivatives. So we have the equation14 d Tt f (x) = ATt f (x), dt T0 f (x) = f (x). For more details see [Du], [App]. Remark 2 We have that A has classical sense for the functions f ∈ Cb2 = C 2 ∩ L∞ . Moreover Cb2 ⊆ D(A) ⊂ D(Tt ) ⊂ L∞ . The domains depends on d, for example if d = 1 D(A) = Cb2 , but for d > 1 it is bigger. We recall that we have continuity in t, but for certain functions, f ∈ 2 Cb , we have differentiability in t. We want the differentiability in x, and this smoothness is because of the smoothness of the coefficients of the SDE considered. Definition 6. A semigroup is a Feller semigroup if f ∈ C(U ) ∩ L∞ (U ) = Cb (U ) implies Tt f (x) ∈ Cb (U ). We have the following result: Theorem 10. If the coefficients of the SDE are as in the theorem 4 and bounded, then the semigroup is a Feller semigroup. Proof. Let x, y be two initial values of the SDE with coefficients as in the theorem 4. Then ~ 0t (x))−f (X ~ 0t (y))|] ≤ ε+2||f ||∞ P (|X ~ 0t (x)−X ~ 0t (y)| > δ). |Tt (f (x)−Tt f (y))| ≤ E[|f (X We can write this expectation as the sum of two terms; if the arguments X0t (x), X0t (y) are close (the continuity of f gives us that the distance of the ~ t (y))| are less than ε) or if they are not. The ~ t (x)) − f (X two values |f (X 0 0 second term is the probability of that the arguments are not close enough. In this second case we can bound ~ 0t (x)) − f (X ~ 0t (y))| ≤ 2||f ||∞ . |f (X We can apply the Chivechev lemma with p = 2 and the theorem 8 to bound the last term with powers of |x − y| and time-dependent constant. This implies the continuity. Indeed, ~ t (x) − X ~ t (y)| > δ) ≤ P (|X 0 0 14 ct 1 ~ t (x) − X ~ t (y)|2 ] ≤ e |x − y|2 . E[| X 0 0 δ2 δ2 For all f ∈ D(A) = Cb2 ⊂ C 2 ∩ L∞ ⊂ D(Tt ) = Cb 28 CHAPTER 1. FIRST RESULTS So Tt f (x) is continuous in x, and then we have some regularity for the solution of the equation d Tt f (x) = ATt f (x), T0 f (x) = f (x) dt with f ∈ Cb and the coefficients of the SDE bounded and Lipschitz. If we suppose that the coefficients of the SDE are C 2,α and the function is f ∈ Cb2 then Tt f (x) ∈ C 2 (Rd ). This is the theorem 7. This result can be optimized. The Bismut-Elworthy-Li formula (see [Du], [Ku]) gives us the differential of Tt f (x) without the derivatives of f . To fix details, ~ t (x) Theorem 11 (Bismut-Elworthy-Li formula). Let f ∈ Cb , a diffusion X 0 2,α d with coefficients Cb and ~v , w ~ two directions in R .The derivative in the ~v direction of Tt f (x) is C|~v | ∇~v Tt f (x) ≤ √ ||f ||∞ . t For the second derivatives the formula is, ∇~v,w~ Tt f (x) ≤ C|~v ||w| ~ ||f ||∞ . t So we only need f ∈ Cb and coefficients of the SDE in Cb2,α for the classical meaning of the generator.15 We recall that p in the Itô diffusion case is the fundamental solution of the parabolic problem, as we expected from the heat equation. Remark 3 We need a Markov process to have a semigroup but not an Itô diffusion. We can have Lévy processes with jumps. This gives us non-local equations and fractional operators. 15 We write Cb2 for the space of functions with 2 bounded derivatives. We write Cb for the space C(U ) ∩ L∞ (U ). Chapter 2 Representation formulas for elliptic equations We start with a definition. We need this definition to study of the bounded domain case. We impose Dirichlet boundary conditions1 Definition 7. A stopping time with respect to a filtration, F(t), is a random variable τ : Ω 7→ [0, ∞] who fulfills {ω, τ (ω) ≤ t} ∈ F(t), ∀t ≥ 0. We have the following results: Proposition 1. Given τ1 and τ2 stopping times with respect to the same filtration. Then 1. {ω, τ1 < t} y {ω, τ1 = t} are in the filtration for all time t. 2. min(τ1 , τ2 ) y max(τ1 , τ2 ) are stopping times. See [Ev] for the proof. We are interested in the least time of hitting a given set. ~ Proposition 2. Let X(t) be the solution of (1.11) with the conditions of the theorem 4 hold. Let E a given closed or open and non-empty set in Rd . Then ~ τ = inf{t ≥ 0|X(t) ∈ E} is a stopping time. 1 Neumann boundary conditions are not studied in this text, but the idea is to define a diffusion reflected in the boundary (see [R]). 29 30 CHAPTER 2. ELLIPTIC EQUATIONS The relationship between the Itô integral and this random variables who are times is clear. The stopping times will be integration limits. See [Ev] for the proof of the following results. Proposition 3. If G ∈ L2 [0, T ] and 0 ≤ τ (ω) ≤ T is a stopping time then the integral Z T Z τ (ω) G1t≤τ (ω) dW GdW = 0 0 fulfills the following properties 1. E Z τ (ω) GdW 0 2. E Z τ (ω) GdW 0 2 =E =0 Z 0 τ (ω) 2 G dt This result is a consequence of the result with deterministic times (see appendix B). We have an Itô formula with stopping times (see appendix A). See [Ev] for the proof. These stopping times τ = τx are random variables with sample points in the continuous functions space. This is because the diffusion takes value in the continuous functions, and a continuous function (random function because we do not know which function we will obtain) have its own value τx . 2.1 The laplacian There are two approaches to the representation formulas. We can start knowing that a certain PDE has a classical solution and conclude that this solution is an expectation. On the other hand we can start with a function, who is an expectation, and show that this function is smooth enough to be a classical solution of an associated PDE. For the moment we consider only the first approach, so we have a classical solution of a PDE and we obtain a probabilistic representation formula. We saw that the generator A associated to the brownian motion is 1 −H0 = ∆. 2 by So we consider the family of equations in a smooth domain U ⊂ Rd given 2 1 − ∆u(x) = f (x) if x ∈ U, 2 2 u(x) = g(x) if x ∈ ∂U The smoothness condition of the boundary of U can be optimized. See [Du]. (2.1) 31 2.1. THE LAPLACIAN We consider the case f = 0 and g a continuous function. Then we look for a function for the harmonic functions. Our problem is 1 − ∆u(x) = 0 si x ∈ U, 2 u(x) = g(x) si x ∈ ∂U (2.2) We apply the Itó formula with integration limit the stopping time ~ (t) ∈ ∂U ) τx (ω) = inf(t|x + W to the stochastic process (where u is the solution of (2.1)) ~ t (x)) = u(x + W ~ (t)). u(X 0 Recall that A = 1/2∆ is the generator. So, Z Z τx τx ~ Auds + u(X0 (x)) − u(x) = 0 τx 0 ~. ∇u · σdW If we take the expectation we observe that the stochastic integration term disappear (see appendix A) Z τx ~ ~ Auds . Ex [u(X(τx ))] − Ex [u(X(0))] = Ex 0 But ~ x ))] = Ex [g(X(τ ~ x )], Ex [u(X(τ ~ Ex [u(X(0))] = u(x), Ex Z τx Auds = 0 0 and we obtain the following result Theorem 12 (Kakutani). Let U ⊂ Rd be a domain and g a continuous function defined in the boundary of U , then the function, u, solution of (2.2) hold the following condition ~ x ))]. u(x) = Ex [g(X(τ (2.3) We have to see that the brownian motioncan not be inside U during infinite time. To see that we suppose that U is contained in the semispace {x1 < a} for certain a. Then we have a τx < τ1,x where the subindex indicates the component and a τ1,x = inf(t|X1 = x1 + W1 (t) = a). 32 CHAPTER 2. ELLIPTIC EQUATIONS Thus a a a P (τ1,x ≤ t) = P (τ1,x ≤ t, x1 + W1 (t) < a) + P (τ1,x ≤ t, x1 + W1 (t) ≥ a) a = 2P (τ1,x ≤ t, x1 + W1 (t) ≥ a) = 2P (x1 + W1 (t) ≥ a) because of the definition of stopping time and the symmetry of the brownian motion. This integral can be calculated explicitly and taking the limit when t → ∞ we obtain that the probability is 1 because of being two times the integral of the standard normal distribution √ in the positive semiaxis. Indeed, if we do the change of variables x = y/ t, we obtain r Z ∞ 2 2 2P (x1 + W1 (t) ≥ a) = e−y /2 dy. π √a t Using this theorem we can obtain the mean value property. Theorem 13 (Mean value property). If u is harmonic then we have Z 1 u(y)dy. u(x) = |∂B(x, r)| ∂B(x,r) Proof. We use the previous theorem, the isotropy of the brownian motion and that the total measure has to be one. Other consequence is that in the case g(x) = 1Γ with Γ ⊂ ∂U , the function u(x) is the probability of hit Γ. Proposition 4. Let u be the solution of (2.2) with g(x) = 1Γ then we have ~ τx (x) ∈ Γ). u(x) = P (X 0 Now we are interested in the second approach we said before to the harmonic functions. We show that if we define ~ τx (x))] v(x) = Ex [g(X 0 we have a classical solution of the PDE considered. It is a well-known fact that the mean value property is true for a harmonic function. Moreover if the function who verifies the mean value property is continuous then the function is harmonic. In the chapter 1 we saw that the semigroup is continuous in x so if we show that the mean value property holds for the v we are done. To do this we have the previous result. In the case x ∈ ∂U we have τx = 0, so the boundary condition holds. We have proof the following result 33 2.1. THE LAPLACIAN Theorem 14. Let u solution of (2.1) with f = 0, the we have ~ x ))]. u(x) = Ex [g(X(τ (2.4) And conversely, let u be the function defined in the previous equation with a certain g in the boundary of a given domain, then u is a classical solution of (2.1). If we consider the case with g = 0 and f a Cbα function we have Theorem 15. If u is the classical solution of (2.1) with g = 0 and f ∈ Cbα (U ) then Z τx s ~ f (X0 (x))ds . u(x) = Ex 0 Proof. The proof is similar to the previous one. We have to observe that the boundary term in the Itô formula is zero. We claim that E[τx ] < ∞ (see the following section for the proof). This fact and that we have f ∈ L∞ (U ) ∩ C(U ) gives us the finiteness of the integral. We can combine the previous result to obtain a formula for the complete problem (2.1) Theorem 16. Let g be a continuous function and f ∈ Cbα (U ). Then a classical solution u of (2.1) verifies Z τx ~ 0s (x))ds + Ex [g(X(τ ~ x ))]. f (X u(x) = Ex 0 Proof. It is enough to put together the previous results. Now we are interested in the steady Schrödinger equation (in units such that the constants are unitary and we suppose that the function is real). 1 − ∆u(x) + c(x)u(x) = f if x ∈ U, 2 u(x) = 0, if x ∈ ∂U (2.5) The function c(x) (the potential) is positive and Lipschitz. The function f is Hölder and bounded. To impose a sign to c is to avoid an eigenvalues problem (so with nonuniqueness). We have the following result: Theorem 17 (Feynman-Kac). The solution of (2.5), with f and c satisfying the previous hypothesis is given by Z τx R ~ s (x))ds t − 0t c(X ~ 0 u(x) = Ex f (X0 (x))e dt . (2.6) 0 34 CHAPTER 2. ELLIPTIC EQUATIONS Proof. We have E[τx ] < ∞ (see the following section for the proof), and f and c are bounded funtions we have bounded the previous integrals. Let u be the solution of (2.5) and consider Rt ~ 0t (x))e− R0t (x) = u(X 0 ~ t (x))ds c(X 0 . We consider also the processes Z0t (x) = − and Z 0 t ~ 0s (x))ds c(X t Y0t (x) = eZ0 (x) . We want to apply the Itô formula. The differentials of these processes are ~ 0t (x))dt dZ = −c(X and ~ 0t (x))Y0t (x)dt. dY = −c(X Appliying the product rule (see appendix A) to R0t (x) we obtain Rt t (x))ds ~ c( X t − ~ (x)e 0 ~ t (x)))Y t (x) + u(X ~ t (x))dY. 0 d u(X = (du(X 0 0 0 0 We apply the Itô formula (see appendix A) to u and we obtain R 1 ~ t (x))ds t − 0t c(X ~ ~ 0t (x))dt + 0 d u(X0 (x)e = ∆u(X 2 d X ~ t (x)) ∂u(X t t 0 ~ (x))(−c(X ~ (x))dt) Y t (x). dWi + u(X 0 0 0 ∂xi i=1 Now we integrate in (0, τx ) and we take expectations. Thus R ~ τx (x))ds τx − 0τx c(X ~ 0 Ex u(X0 (x)e − Ex [u(x)] = Z τx 1 t t t t ~ ~ ~ ∆u(X0 (x)) − c(X0 (x))u(X0 (x)) Y0 (x)dt . = Ex 2 0 We conclude using the boundary conditions and the equation. Z τx R ~ s (x))ds t − 0t c(X ~ 0 dt f (X0 (x))e u(x) = Ex 0 2.2. POISSON EQUATION AND SHAPE RECOGNITION 35 This equation is diffusion with killing. We consider a brownian particle ~ t (x))h be the probability of disappearing in who can disappear. Let c(X 0 (t, t + h) interval. Then the probability of survival until time t is approximated by ~ t1 (x))h)(1 − c(X ~ t2 (x))h)...(1 − c(X ~ tn (x))h) (1 − c(X 0 0 0 where ti is a partition of the (0, t) interval with h as the step. As we take the limit h → 0 this probability converges to the exponential e− Rt 0 ~ s (x)ds) c(X 0 . And so u is the mean of f in the brownian paths with diffusion hitting the boundary of U . 2.2 Poisson equation and shape recognition This section is based in the paper [GGSBB], however, the arguments in this paper are with random walks. In the limit appears a constant, and they do not care about. This constant is 12 . The idea is that in the limit we consider a brownian motion with generator 1/2∆. We consider a silhouette with a simple closed curve as boundary. The ’classical’ fashion of obtaining properties is to assign, for each point x in the silhouette, a value that gives us the position with respect to the boundary. The popular way to do that is considering the distance. In this work we consider a different one. We consider brownian particles started at each point and we measure the expected time of hitting the boundary. As we mention before, in the paper [GGSBB] they argue with random walks and a discretized laplacian. ~ t (x) = x + W ~ (t), for each interior We consider the brownian motion X 0 point x. We consider the equation 1 ∆u = −1 2 with homogeneus Dirichlet boundary conditions. With the previous section notation we have the case f = 1 and g = 0. We obtain a 2 that is not in the paper, but the authors mention that a constant appears and they take this constant as the unity. Let τx = inf{t|X0τx (x) ∈ ∂S}, where S is the silhouette (domain). Theorem 18. For the classical solution of the previous equation we have u(x) = E[τx ]. (2.7) 36 CHAPTER 2. ELLIPTIC EQUATIONS Proof. We apply the Itô formula (we can because u is regular enough) to ~ min(τx ,n) (x)), and after taking expectations we obtain the process u(X 0 Z min(τx ,n) 1 ~ min(τx ,n) (x))] − Ex [u(x)] = Ex ~ 0s (x))ds . Ex [u(X ∆u( X 0 2 0 We use the boundary conditions and the equation to obtain Z min(τx ,n) min(τx ,n) ~ 1ds = −Ex [min(τx , n)]. Ex [u(X0 (x))] − u(x) = −Ex 0 We show before that P (τx < ∞) = 1, but this does not give the integrability. To see that we have the integrability of τx we have to use the properties of u. We have that u is bounded and so limn→∞ Ex [min(τx , n)] < ∞. Figure 2.1: Numerical experiment, silhouette and u. We know that u is as regular as the boundary allows and it is positive. The positivity property can be obtained only seeing the expectation. The level set of u gives us smooth approximations of the boundary. In the paper [GGSBB] the authors mention other properties of u (existence and uniqueness of solutions with other boundary conditions, the mean value property...). 2.3. A GENERAL ELLIPTIC EQUATION 37 We can use u to divide the silhouette in parts with thresholding. But with this method we can lose information in the periferic parts of our silhouette. To solve this problem (loss of information) we consider Φ(x) = u(x) + |∇u(x)|2 . The most important properties of Φ is that the high values indicate concavities (the gradient is big there) and that we can divide the silhouette without loss of information if we use this function. To detect concavities we can use Φ but there is a better way. We define the function ∇u . Ψ(x) = −∇ · |∇u| Remark 4 This operator is the 1−laplacian. We will study in a deeper way in the chapter 5. The high values (in absolute value) indicate the curves in the silhouette. So we can obtain the corners of the silhouette. The negative values of Ψ indicate concavities. As the value is more negative, more ’pointed’ is the concavity. And conversely, the high positive values indicate convexities. To find the ’skeleton’ of our silhouette we define the function Ψ̃ = −uΨ/|∇u|. And we use the threshold method. We want to discern between different silhouettes. To do that we consider the ’decision trees’ (see [GGSBB]). We give an example. We see (Figure 2.2) the high values of Φ close to the chimney of the house, this is the concavities zone. We also see that Ψ detect the concavities and the convexities (lower corners), and the ’skeleton’, the high value in the center of the house. Remark 5 All the programs we use have a time counter. When we run the programs in my PC (1.73 GHz) with 10−4 as the tolerance marks 20 seconds. But this is a small image, 50 × 150 pixels. 2.3 A general elliptic equation We can obtain representation formulas for other elliptic operators. We are going to study d d X ∂u ∂2u 1 X bi (x) ai,j (x) + + c(x)u = f si x ∈ U, (2.8) −à = − 2 ∂xi ∂xj ∂xi i,j=1 i=1 38 CHAPTER 2. ELLIPTIC EQUATIONS Figure 2.2: Results, Φ in the upper figure, Ψ in the lowe figure. Cb2,α u = g iff x ∈ ∂U where ai,j , bi , c are in (for α ∈ (0, 1]), c is positive and the characterPd istic form − i,j=1 ai,j λi λj is positive. We can apply the sema idea to this operator.3 In physics, this operator is a hamiltonian. c is the potential. Let g be a continuous and bounded funtion and f ∈ Cbα . Moreover we suppose (ai,j )(x) = σ(x)σ t (x) with σ(x) a Cb2,α matrix. σ there exists if the determinant of a never vanishes.4 It is possible that there is no unique σ and so there are more than one diffusions, however this is not a problem because the measures in (1.15) induced by the respectives diffusions are the same. Thus the expectation is well defined (see [F] for the proof). Given an equation as (2.8) the diffusion we consider is the solution of5 ~ 0t (x) = ~b(X ~ 0t (x))dt + σ(X ~ 0t (x))dW ~. dX Let τx be a stopping time as defined in the previous section. 3 We need to impose conditions to have a classical solution (see appendix A). This result is not optimal (see [F]). 5 ~ 0t (x) is the value of the solution of the SDE started in x at time Recall the notation, X 4 t. 39 2.4. UNBOUNDED DOMAINS If the previous regularity conditions for the coefficients of (2.8) hold then we have the following results (the proofs are similar to the proofs in the laplacian case): Theorem 19. Let g be a continuous function and f ∈ Cbα (U ). We suppose c = 0. Then the solution, u, of (2.8) is Z τx ~ s (x))ds + Ex [g(X(τ ~ x ))]. f (X u(x) = Ex 0 0 Theorem 20 (Feynman-Kac). The solution of (2.8), with g ∈ Cb , f ∈ Cbα (U ) and c bounded, positive and Lipschitz is given by Z τx Rτ Rt ~ x ))e− 0 x c(X~ 0s (x))ds . ~ 0t (x))e− 0 c(X~ 0s (x))ds dt + Ex g(X(τ f (X u(x) = Ex 0 (2.9) We can apply these methods to the ’iterated’ operators, for example the bilaplacian with the correct boundary conditions, ∆∆u = 0, u = g1 , ∆u = g2 . Indeed, we can write the equation as ∆v = 0, v = g2 ∆u = v, u = g1 and we apply the same previous idea. 2.4 Unbounded domains Now we are interested in elliptic operator in unbounded domains. For example we are interested in the Schrödinger equation, (2.5), in the whole space. We can not argue with stopping times, because there is not boundary to hit. However we can see the Schrödinger equation in a disc with radious R centered at the origin (we write this domain as UR ) and take the limit in R. We expect that lim UR = Rd and τx → ∞. So, if we consider the 3D case for the laplacian, −∆u = f if x ∈ R3 we know that the Green operator (see [Ev2]) is G(x, y) = 1 1 4π |x − y| 40 CHAPTER 2. ELLIPTIC EQUATIONS and that u is the convolution with G. If our idea is correct, our diffusion, with transition density p(t, x, y), has to fulfil that the time integral will be the Green operator. Indeed, we have Z ∞ p(t, x, y)dt = G(x, y). 0 To see that we do the following change of variables t= |x − y|2 2s in the integral. The same is valid for the elliptic operator (2.8). We obtained the well-known results with different methods, and sometimes (Schödinger or Navier-Stokes equations) give new intuition to the problem. These methods can be applied to give properties of this equations. For example, in [CR] they proof an Harnack inequality for (2.5). Summarizing, we have a representation formula of the classical solution (we know that this classical solution exists) for the elliptic problem (2.8) as an integral in (1.15) and being U a bounded or unbounded domain. The advantages of these methods is the new intuition and the new numerical methods. Chapter 3 Representation formulas for parabolic equations 3.1 A general parabolic equation In the first chapter we studied that a Markov process defines a contraction semigroup on L∞ . In the case where the Markov process is an Itô diffusion the generator of the semigroup is an elliptic operator, and so the function Tt f (x) solves a parabolic equation.1 Remark 6 We suppose f ∈ Cb2 (Rd )∩L1 (Rd ). Then, using interpolation, we have f ∈ Lp for 1 ≤ p ≤ ∞. We know that Tt f (x) is C 1 in time and that the two derivatives in space depend on the coefficients regularity.2 We consider the SDE coefficients Cb2,α, so the stochastic flow has two spatial derivatives (theorem 7). In this chapter we consider the Cauchy problem for general linear parabolic equations, however, if we consider stopping times we can do the same for initial-boundary problems (if we suppose a smooth domain, or at least with the interior cone property, see[Du]). In the first chapter we saw that if f ∈ Cb2 ∩ L1 (Rd ) and we consider the heat equation ut (t, x) = −H0 u(t, x) = 1 ∆u(t, x), 2 u(0, x) = f (x) we have, in the notations used in the text, ~ t (x))] = E0 [f (x + W ~ t (0))] = e−H0 t f (x) u(t, x) = Tt f (x) = Ex [f (X 0 0 and these expectations are with respect to the Wiener measure. So, we are integrating in functions. We see that the kernel is the integral with respect 1 We consider that the semigroup takes function in f ∈ Cb2 . This result is not optimal, see theorem 11. 2 We can write the derivatives of f with the Bismut-Elworthy-Li formula (theorem 11). 41 42 CHAPTER 3. PARABOLIC EQUATIONS to the brownian bridge measure in the functions,3 i.e. Z dWxy . p(t − s, x, y) = Cxy [s,t] We need another integration to obtain the measure induced by the brownian motion Z Z dWx . p(t, x, y)dy = Cx [0,T ] R Remark 7 Using the theorem 11 and the theorem 7 we can conclude that if f ∈ Cb and our diffusion X0t (x) has C ∞ (Rd ) coefficients, then Tt f (x) ∈ C ∞ (Rd ). For example, this occurs with the heat equation. Remark 8 In the third section of this chapter we will write p(t − s, x, y) = W (s, x, t, y). This is because we want to conserve in the possible the Feynman notation. Remark 9 If we do not have an unique diffusion (we have many) for a parabolic problem this is not a problem, because the measures induced are equivalent. And so the expectation is well-defined (see [F]). Let4 d d X ∂ ∂2 1 X bi (x) ai,j (x) + − c(x) à = A − c(x) = 2 ∂xi ∂xj ∂xi i=1 i,j=1 be an elliptic operator with Cb2,α coefficients. Moreover we suppose c(x) ≥ 0.5 We suppose that A the generator of an Itô diffusion, this is (ai,j (x)) = σ(x)σ t (x) for a Cb2,α matrix σ. The stochastic flow defined by the SDE ~ ~ ~ ~, dX(t) = ~b(X(t))dt + σ(X(t))d W X(0) = x is C 2 in the spatial variable (see theorem 7). Given a parabolic equation ∂u(t, x) = Ãu(t, x), ∂t u(0, x) = f (x). (3.1) We have the following results: 3 This is true for a general diffusion. The sign of c is to have a contraction semigroup on L∞ . We remarck that the coefficients of the elliptic operator are time-independent. 5 If we take a negative c, we can obtain non-trivial statonary solutions. These solutions are related with the eigenvalues of the elliptic problem. 4 43 3.1. A GENERAL PARABOLIC EQUATION Theorem 21. Given the equation (3.1) fulfilling the previous hypothesis and with c = 0. Let f ∈ Cb2 (Rd ) and let u(t, x) be its classical solution. Then we have ~ 0t (x))] = eAt f (x) u(t, x) = Tt f (x) = Ex [f (X where the measure is the induced by the diffusion measure. And conversely, if we define v(t, x) = Tt f (x) then v verifies the equation (3.1) in the classical way. Proof. Fix t. We apply the Itô formula (we can because of the smoothness ~ s (x)), of u) to u(t − s, X 0 Z s Z s ∂ r s ~ 0r (x))·σdW ~. ~ ~ ∇u(t−r, X +A u(t−r, X0 (x))dr+ u(t−s, X0 (x)−u(t, x) = ∂r 0 0 We observe that ∂ ∂ =− . ∂r ∂t No we take s = t and we take expectations, obtaining ~ 0t (x))] − u(t, x) = 0 Ex [f (X and so u(t, x) = Tt f (x). Now let v(t, x) be a function defined as in the statement. In the chapter 1 we studied that v(t, x) is C 1 in time. We deduce that v ∈ C 2 spatially because of the smoothness of the SDE coefficients (see [Ku] or the theorem 7) and the Bismut-Elworthy-Li formula (theorem 11). Obviously hte initial condition holds. Finally we have to prove that the equation holds. We studied that the generator of the diffusion is our elliptic operator A, and so the semigroup solves the parabolic equation. If we consider c(x) ≥ 0, then we have the Feynman-Kac formula (now in the parabolic case). Given the parabolic problem ∂u = Ãu, ∂t u(0, x) = f (x) (3.2) then Theorem 22 (Feynman-Kac (parabolic case)). Given the equation (3.2) fulfilling the previous hypothesis and with c ≥ 0. Let f ∈ Cb2 (Rd ) and let u(t, x) be the classical solution of this problem. Then ~ t (x))e− u(t, x) = T̃t f (x) = Ex [f (X 0 Rt 0 ~ s (x))ds c(X 0 where the measure is the induced by the diffusion measure. ]. 44 CHAPTER 3. PARABOLIC EQUATIONS Proof. Fix t. We consider the processes Z r r ~ s (x))ds, c(X Z0 (x) = − 0 0 r Y0r (x) = eZ0 (x) with differentials ~ 0r (x))dr, dZ0r (x) = −c(X ~ 0r (x))Y0r (x)dr. dY0r (x) = −c(X ~ r (x))Y r (x) is Then the differential of the product u(t − r, X 0 0 ~ 0r (x))Y0r (x)) = d(u(t − r, X ~ 0r (x)))Y0r (x) + udY0r (x). d(u(t − r, X By the Itô formula applied to u we have ~ r (x)) ∂u(t − r, X 0 ~ 0r (x)) dr + ∇u(t − r, X ~ 0r (x)) · σdW ~. − + Au(t − r, X ∂t If we introduce this into the previous equation we obtain ~ r (x)) ∂u(t − r, X r 0 r r ~ ~ + Au(t − r, X0 (x)) dr + d(u(t − r, X0 (x))Y0 (x)) = − ∂t ~ 0r (x)) · σdW ~ Y0r (x) − u(t − r, X ~ 0r (x))c(X ~ 0r (x))Y0r (x). ∇u(t − r, X We integrate until r = t and we take expectations. The result is Z t R ~ r (x)) ∂u(t − r, X ~ s (x)ds 0 r − 0r c(X ~ 0 T̃t f (x)−u(t, x) = E − +Ãu(t−r, X0 (x)) e dr = 0. ∂t 0 Remark 10 We point out that the PDE who appears with the Itô formula with the SDE as in the first chapter is backward in time, i.e. ∂u + Au ∂r To obtain the ’correct’ direction for the time we consider u(t − r, x). However we this is not the unique way. In [Ku] we can see that a backward SDE with final datum x gives us the equation ∂u − Au ∂t ~ t (x), this is our ’initial’ ~ t (x)) particles in X So the idea is we have f (X 0 0 point. These particles move according to the SDE considered.6 Thus u(t, x) = Tt f (x) 6 We recall that in our statement of the SDE (chapter 1) this was our final point, but the intuition comes if we consider this point the initial point. 45 3.1. A GENERAL PARABOLIC EQUATION is the expected amount of particles in the ’final’ point x at time t. We can understand the Itô diffusions as ’characteristic curves’. And ~ t (x) and so, given x, t we calculate u(t, x) looking f in the original point X 0 taking expectations. The same as in the deterministic characteristic curves case. Remark 11 The original problem for (3.2) with ai,j = δi,j , ~b = 0 is quantum mechanics, so commonly we can see 1 ∂u(t, x) = ∆u(t, x) − V (x)u(t, x), ∂t 2 u(0, x) = f (x) We see that ||T̃t f (x)||∞ ≤ ||f (x)||∞ e−t||c(x)||∞ ≤ ||f (x)||∞ so T̃t verifies the same properties that Tt semigroups. Moreover, from mass conservation Z Z Z Z t p(t, x, y)f (y)dydx = Ex [f (X0 (x))]dx ≤ Rd Rd Rd f (y)dy Rd and from being a contraction semigroup on L∞ we conclude 1 1− 1 ||Tt f (x)||p ≤ ||f ||1p ||f ||∞ p , ∀p ∈ [1, ∞], ∀t ≥ 0. If we have a result about the decay in the L∞ norm the probability density then we show the decay rate in Lp for all p. In [Fr] (Theorem 4.5) we have the following bound C |p(t, x, y)| ≤ d/2 . t Another proof of the result is to consider an adapted to the advection coordinates. If we consider the deterministic characteristic curves Z t ~b(Y ~ ~ (s))ds Y (t) = 0 and we define ~ (t)) v(t, x) = u(t, x − Y thus d X i,j=1 ai,j (x)uxi ,xj = d X ai,j (x)vxi ,xj i,j=1 but ~ ′ (t) = ut − ux~b(x). vt = ut − ux Y 46 CHAPTER 3. PARABOLIC EQUATIONS We see that v solves the ’heat’ equation (with a coefficients ai,j (x)) and we have ||u||∞ = ||v||∞ , so we show the decay (see [I] for the fundamental solution for this problem) and we conclude 1 1− 1p ||Tt f (x)||p ≤ ||f ||1p ||f ||∞ ≤ c 1 (td/2 )1− p , ∀p ∈ (1, ∞], ∀t ≥ 0. Remark 12 We can generalize these methods for bounded domains Dirichlet boundary conditions. We do considering a stopping time τx (ω) as in the previous chapter and the t < τx , t ≥ τx cases, so we obtain a boundary term. We do not do that because is quite similar to the previous chapter method. 3.2 The Fisher equation We consider one dimensional Fisher equation7 1 ∂ 2 u(t, x) ∂u(t, x) = + u(t, x)2 − u(t, x), ∂t 2 ∂x2 u(0, x) = f (x) (3.3) 1 0.9 0.8 0.7 u 0.6 0.5 0.4 0.3 0.2 0.1 0 −20 −15 −10 −5 0 x 5 10 15 20 Figure 3.1: Traveling wave solution of (3.3). To obtain our stochastic representation we consider a slightly different process. The process is a branching brownian motion. Be a particle with a brownian path and a exponential time8 , T . At this exponential time the particle branchs in two identical particles, so each particle follows a brownian path with branching. These particles are independent ones from the others. Ww write x1 (t), ...xn (t) for the positions of the particles with P (n = k) = e−t (1 − e−t )k−1 . 7 8 Another name for this equation is Kolmogorov-Petrovskii-Piskunov equation. This is P (T > t) = e−t . 47 3.2. THE FISHER EQUATION We want to show that, under certains assumptions in f , the solution of (3.3) is u(t, x) = Ex [f (x + x1 (t)), ..., f (x + xn (t))]. The equation is used in population dynamics, if the population considered presents movement (diffusion) and grow (birth).9 This is what the process means, birth and diffusion. We need an assumption in f to guarantee the existence of the expectation. Exactly we have Theorem 23. Given f ∈ Cb2 (Rd ),1 ≥ f ≥ 0, we have that u(t, x) = Ex [f (x + x1 (t)), ..., f (x + xn (t))] verifies the equation (3.3). Proof. We consider the T ≤ t or T > t cases. Thus we have Z ∞ e−s ds = e−t P (T > t) = t and then, if T > t, u(t, x) = Ex [f (x + W0t (0))] = e−H0 t f (x). Suppose that we are in the other case, T ∈ (s, s + ds), with s + ds < t. The probability of this situation is e−s ds. The two new particles start their movement at x + x1 (T ), and gives us two independent examples of the same process but translated in space (x + x1 (T )) and time (t − s). As we have independence the expectation of the product is the product of expectations and we have u2 (t − s, x + x1 (t)). Taking the expectations in all position x + x1 (t) = y we obtain the term Z ∞ P (x + x1 (s) ∈ dy)u2 (t − s, y) = e−sH0 u2 (t − s, x). −∞ And if we put all the previous calculations together we conclude Z ∞ Z t −H0 t P (x+x1 (s) ∈ dy)u2 (t−s, y) u(t, x) = P (T > t)e f (x)+ P (T ∈ dy) −∞ 0 or equivalently −t −H0 t u(t, x) = e e f (x) + Z 0 t e−s e−sH0 u2 (t − s, x)ds. We have u(0, x) = f (x). 9 Actually the equation in the population dynamics has a reaction u − u2 , but from (3.3) we recover the other with the change u 7→ 1 − u. 48 CHAPTER 3. PARABOLIC EQUATIONS We change variables s′ = t − s obtaining −t −H0 t u(t, x) = e e f (x) − Z t ′ ′ es −t e(s −t)H0 u2 (s′ , x)ds′ . 0 Now taking derivatives in t we have ∂u(t, x) = −e−t e−H0 t f (x) + e−t (−H0 )e−H0 t f (x) + u2 (t, x)− ∂t Z t 0 ∂ s′ −t (s′ −t)H0 2 ′ e e u (s , x) ds′ ∂t We observe that the last term is the needed term to complete −u and uxx . So u(t, x) verifies the equation (3.3) with the initial datum f . The smoothness is not a problem because of the previous expression and the f and the process considered regularity. Previously we have representation formulas for linear equations. This is the first non-linear equation. See [McK] for a more detailed study. The idea of a particle with movement and branch can be used in other semilinear parabolic (or elliptic) equations with a polinomic nonlinearity. 3.3 Feynman and quantum mechanics In this section we study the Feynman path integral and its relation with the objects defined in the previous chapters. We consider the onedimensional case, but for a general case it is the same. We want to preserve the notation. The notation is different of the notation in the previous chapters. We want to preserve the Feynman calculations and ideas, so we do not care about the rigour.10 We think that this section has an historical interest besides the academic interest, so we have to conserve the original notations and calculations. We have seen how to obtain a PDE solution thanks a functional integration. With this idea we study quantum mechanics. However, this beautiful idea11 is more extent between phisicist than between mathematicians. From a mathematician point of view it is not a completly succesful method because of some problems with the measures in functions (see below). Feynman, in [Fe], says: 10 Feynman, in [FH], says ’The physicist cannot understand the mathematician’s care in solving an idealized physical problem. The physicist knows the real problem is much more complicated. It has already been simplified by intuition, which discards the unimportant and often approximates the remainder. 11 Beautiful at least in the humble author’s opinion. 3.3. FEYNMAN AND QUANTUM MECHANICS 49 The formulation given here suffers from a serious drawback. The mathematical concepts needed are new. (...) One needs, in adittion, an appropiate measure for the space of the argument functions x(t) of the functionals. We follow closely [Fe] (this is a review of the Feynman’s thesis [Fe2]). We start with a ’Chapman-Kolmogorov’ equation (see chapter 1). Let A, B, C be three measurements of the state of a certain physical system such that the system is completly know. Let Pab be the probability of, given A = a, having B = b. In a similar way we define Pbc . Then, if we assume independence, we have Pabc = Pab Pbc and we expect the relation Pac = X Pabc . b This is the gratest difference between classical mechanics and quantum mechanics. In the classical formulation the previous equation is true, while in the quantum formulation is not. The reason is that the intermediate quantum state b is not well defined forever. We have to measure (and so the system has an interference) to becomes true the previous equation. What we have in the quantum case is that there exists complex numbers, φij , such that Pab = |φab |2 , Pbc = |φbc |2 , and we know the following relation12 φac = X Pac = |φac |2 φab φbc . b The physical meaning of this equation is that the probability of a particle without relativistic effects and spinless goes from a to c can be calculate as the square of some complex quantities, each one associated to a possible path. We know by intuition the main result. We know the least action principle, who says that the classical path is a minimum of the action functional. Z tb A = L(X ′ (t), X(t), t)dt ta where L is the lagrangian of the system. For example, for a particle with mass m moving under the influence of a potential V (x) the lagrangian is 1 mX ′ (t)2 − V (X(t)). 2 12 These equations are the Chapman-Kolmogorov equation seen previously. 50 CHAPTER 3. PARABOLIC EQUATIONS Remark 13 Rigorously, it is not needed that our path would be a minimizer, it is enough for it to be a critical point of the considered functional. In this context we use the equations to solve a variational problem, in other times the situation is the converse, we solve a variational problem to solve a PDE problem (for example the Dirichlet problem for the Poisson equation). If P (ta , xa , tb , xb ) is the probability of our particle13 moves from the point xa at time ta = 0 to the point xb at time tb = T we have P (b, a) = |K(ta , xa , tb , xb )|2 for a funtion K with K(ta , xa , tb , xb ) = X φ(X(t)) all paths from (ta , xa ) to (tb , xb ) Remark 14 It is necessary remark that we do not specify the paths we consider, but that the two extremal points are fixed suggests the space (1.16). The physicist are not used to specify the space they consider. The idea is that all path contribute, but in a different manner. Finally φ(X(t)) = Cei/~A (X(t)) where C is a constant to normalize. Phisically this means that our particle, for example a photon, travels for ’all’ possible paths between two points, but in different phases. This is the de Broglie wave-particle duality. Before we continue we are going to think how recover the classical mechanics doing the Planck constant 14 goes to zero. We obtain in a natural way the scale in that the quantum mechanics is a good model. These limits are know as ’semiclassical limits’, because they are not classical (~ 6= 0) but the system behave as it would be classical. If we do a small perturbation in the classical scale the contribution of the action is small in the classical scale, but it is not in the Planck constant scale, where the changes are big. Then our angle15 oscillates in such a way that the total contribution is zero. If we consider a path, X1 , who is not a critical point of the functional, there exist another path, X2 , close to the former and such that the contribution of X2 is the opposite of the contribution of X1 . So we only have to take account of the path in a neighbourhood of X, where X is a critical point of the action. And in the classical limit (~ → 0) the unique important path is the critical point of the functional. 13 We consider the pinned Wiener measure, so the two boundary points are fixed. This is not a problem (see chapter 1), is essentially the same as the Wiener measure used in the previous sections. 14 The Planck constant is asocciated to the quantization. 15 We have a complex exponential. 3.3. FEYNMAN AND QUANTUM MECHANICS 51 To define the path integral we consider a sequence of times16 , ti = ta +εi, i = 0, 1, ...N and the position of the particle at these times, Xi = X(ti ). Then Z K(ta , xa , tb , xb ) ≈ C φ(X1 , ...XN −1 )dX1 dX2 ...dXN −1 . We need to take the limit and a constant to normalizen, and this is a problem. However, in the case of a particle moving under a potential V the constant is (see [FH]) −N C=A = 2πi~ε m −N/2 . So we have (in this case the limit exists (see [FH])) Z dX1 dXN −1 1 ... ei/~A (X1 ,...XN−1) K(ta , xa , tb , xb ) = lim ε→0 A A A (3.4) where A (X1 , ...XN −1 ) is the integral over the path who takes values Xi at times ti and linear between them.17 This definition of a path can be a problem although we do not take the limit, because in the points with X ′ (t) discontinuous (the positions Xi ) the second derivative is not finite, this is that the aceleration is not finite. Feynman knows that this can be a problem, but he also says that he can ’solve’ it with the substitution of X ′′ (t) with the finite diferences ε12 (Xi+1 − 2Xi + Xi−1 ). Feynman is not worried about these problems and says Nevertheless, the concept of the sum over all paths, (...), is independent of a special definition and valid in spite of the failure of such definitions. And so he writes the path integral, understood as the limit when N → ∞ in the previous equation,18 Z K(ta , xa , tb , xb ) = ei/~A (X(t)) DX(t). (3.5) In [Fe] we can see how, with formal calculations, Feynman ’show’ that K defined as above verifies the Schrödinger equation i~ 16 ∂ϕ(t, x) ~2 ∂ 2 ϕ(t, x) =− + V (x)ϕ(x, t) = Hϕ(t, x). ∂t 2m ∂x2 This is the way we define the cylinders in the chapter 1. Taking limits we obtain a path nowhere differentiable, exactly the same as in the brownian case. 18 All path integrals are understood as a limit process in N . 17 52 CHAPTER 3. PARABOLIC EQUATIONS It is a well-known fact that, if f is a given initial value, the solution can be written as, Z K(0, x, t + s, y)f (y)dy ϕ(t + s, x) = ZR Z K(0, x, s, z)K(s, z, t + s, y)f (y)dzdy = ZR R K(0, x, s, z)ϕ(s, z)dz. = R This equation gives us a N times iteration. If we consider ti = ta + iε we obtain the formula (3.4). Moreover we can write K(ta , xa , tb , xb ) = e−(i(tb −ta )/~)H (xa , xb ). Previously we wrote p(t − s, x, y) = Z Cxy [s,t] dWxy (3.6) for the heat kernel, where the initial point x and the final point y are fixed. This remind the previous calculation (3.5).19 We are going to show the relation with the integral with respect to the Wiener measure. We consider the equation 1 ∂ 2 ρ(t, x) ∂ρ(t, x) = − V (x)ρ(t, x). ∂t 2 ∂x2 With V = 0 (the heat equation case) we have, if f is a given initial value20 Z W (0, x, t + s, y)f (y)dy ρ(t + s, x) = ZR Z W (0, x, s, z)W (s, z, t + s, y)f (y)dzdy = ZR R W (0, x, s, z)ρ(s, z)dz. = R If we iterate N times, with times ti = ta + iε we obtain the formula Z N PN Y 2 ρ(tb , xb ) = C(ε) e(−1/2ε) l=0 (Xl+1 −Xl ) ρ(ta , xa ) dxl . l=1 If we compare the two formulas we expect that, as the calculation is valid for all N , in the limit N → ∞ we have Z R tb ′2 (3.7) W (ta , xa , tb , xb ) = N1 e−1/2 ta X (t)dt DX(t) 19 20 In this context the kernel is called ’propagator ’. We change the notation for the kernel to conserve the Feynman notation. 3.3. FEYNMAN AND QUANTUM MECHANICS 53 where N1 is a constant to normalize. Recall that W (ta , xa , tb , xb ) = Z x Cxab ([ta ,tb ],R) dWxxab . In the free particle (with mass m) case we have Z R tb ′2 K(ta , xa , tb , xb ) = N2 ei/~ ta 1/2·X (t)dt DX(t). (3.8) We observe the similarities between the kernels (3.7) and (3.8). However there are big differences between them. The integral in (3.7) is the Wiener integral, and so completly rigourous. The integral in (3.8) is not rigurous. A problem is the measure considered, who is finitely additive (the Feynman measure or a Wiener measure with complex diffusion constant are finitely additives (see [Kl] and references therein). Other problem is that we write X ′ (t), but the paths considered are in the space (1.16), and so they are related to the brownian bridge, thus they are nowhere difRas X are brownian paths, we understand (see [Kl]) Rferentiables. RHowever, dW ′ X (s)dt = dt dt = dW . Thus this terms are Itô stochastic integrals. If we consider a non-zero potential the Feynman-Kac formula (see section 1 of this chapter) gives us Z R tb R tb xb −V (X(t))dt t a e ta −V (X(t))dt dWxxab W (ta , xa , tb , xb ) = Exa [e ]= x Cxab ([ta ,tb ],R and in the Feynman’s notation this is (formally at least) Z R tb ′2 W (ta , xa , tb , xb ) = N3 e ta −1/2·X (t)−V (X(t))dt DX(t). (3.9) (3.10) Again we have similarities between them, but only in the Wiener case there are rigorous integrals. We have seen that the path integral gives us a kernel. Actually if 1 H = − ∆ + V (x) 2 then W (ta , xa , tb , xb ) = e−(tb −ta )H (xa , xb ). We are going to see why in the Wiener case the path integral is valid. Taking limits in N we have that the integral in the path space (the limit of the product of measures in the space is the measure in the path space) is infinite. Indeed Z Z Y N dX(ti ) = ∞. DX(t) = lim ε→0 i 54 CHAPTER 3. PARABOLIC EQUATIONS So the exponential in (3.7) has to vanish. In other way the integral will not be well defined. This happens when the path is nowhere differentiable, as in the brownian bridge (or in the brownian motion) case. This is the process who induce the measure considered (see section 1 of this chapter). There are attempts to rigorize this integrals. For example, Itô consider a regularization term and pass to the limit to vanish it. Indeed he writes Z R tb ′′2 R tb 1 1 ′2 ′2 lim N (ν) ei/~ ta [ 2 mX (t)−V (X(t))]dt e− 2ν ta [X (t)+X (t)]dt DX(t). ν→∞ The idea is that the paths are smoother than before because the second derivative (and not the first derivative) is infinite. See [Kl] for other ways to improve the path integral. Feynman writes his thesis (1942) considering space-time paths. Some years after (1951) he considers phase space paths. To do that we recall the relation between the hamiltonian and the lagrangian of a system: L(X ′ (t), X(t)) = p(t)q ′ (t) − H(p(t), q(t)). Fixed ta and tb , we consider the phase space paths q(t) who, at these times, are in the fixed positions qa and qb . We consider the phase space paths p(t) with ta ≤ t ≤ tb but now the initial and final positions are free. p, q are brownian paths. Thus, if we follow the previous ideas we expect21 Z R tb ′ K(ta , qa , tb , qb ) = M e(i/~) ta p(s)q (s)−H(p(s),q(s))ds Dp(t)Dq(t). (3.11) We recall that as p, q are brownian paths, they are R ′nowhere differentiable. However as we R said before, we can understand the q (s)dt terms as stochastic integrals dW . The great advantage of the pase space path integrals is that we can apply it for relativistic particles. So we have a way to define the quantum fields. For example, if we consider a free relativistic particle, with hamiltonian in units such that the speed of light is 1 p H(p(t), q(t)) = p2 (t) + m2 we have the kernel R(ta , qa , tb , qb ) = M Z e(i/~) R tb ta p(s)q ′ (s)− √ p2 (s)+m2 dt DpDq. Remark 15 This is a non-local operator. Recall that px = −i~ 21 ∂ ∂x As previously, these are formal calculations (see [Kl] for the rigorization). 3.3. FEYNMAN AND QUANTUM MECHANICS 55 thus ∂2 . ∂x2 Before we conclude we have to make Rsome comments. Above we said that we can understand certain terms ( q ′ (s)dt) as stochastic integrals, however we do not say in which sense, Itô or Stratonovich. We consider now the Stratonovich sense, because the relativity impose certain change of coordinates. This is the reason of why the Schrödinger equation is not a relativistic equation. We need an equation with same order in all variables, and the Schrödinger equation does not hold this condition. So if we can apply the usual chain rule we have an advantage. We can rigorize the previous calculations if we consider the regularization (see [Kl]) Z √2 R tb R ′2 1 ′ ′2 2 lim Mν e(i/~) ta p(s)q (s)− p (s)+m dt e− 2ν p (s)+q (s)dt DpDq p2x = −~ ν→0 Remark 16 |ϕ(t, x)|2 is the probability of find our particle in the point x at time t, but |ϕ(p, q)| is the probability of being in a certain state (p, q). So the interpretation is harder. Chapter 4 Representation formulas in fluid dynamics We consider the homogeneus, isotherm, isotropic and incompressible NavierStokes equations. ∂~u u · ∇)~u + ∇p − ν∆~u = 0 ∂t + (~ (4.1) [N S] ∇ · ~u = 0 The spatial domain is Td , so we have periodic boundary conditions. The initial data is f~(x) ∈ C k+1,α , with k ≥ 1. If we do not have viscosity, i.e. if ν = 0, we obtain the Euler equations ∂~u u · ∇)~u + ∇p = 0 ∂t + (~ (4.2) [Euler] ∇ · ~u = 0 Figure 4.1: Navier-Stokes solution at time 10. We will obtain a probabilistic representation for the Navier-Stokes equations. Using this representation we show the local existence (in time) of classical solution for the Navier-Stokes equations. However, we start with the Burgers equation. 56 4.1. THE 1-DIMENSIONAL BURGERS EQUATION 57 Figure 4.2: Stokes problem solution. 4.1 The 1-dimensional Burgers equation We start with the Cauchy problem for the inviscid Burgers equation, but the representation is for the viscid one.1 So, given an initial value f ∈ Cb2 , we consider the equations vt + vvx = 0 (4.3) ν vxx (4.4) 2 The idea is to use the Hopf-Cole transformation and the representation for the heat equation. vt + vvx = Lemma 4 (Hopf-Cole). Let u(t, x) be a classical solution of the heat equation with viscosity ν/2, then v = −ν(log u)x (4.5) is a classical solution of the viscid Burgers equation (4.4). Proof. Calculating the derivatives of (4.5) we obtain vt = −ν uxt u − ut ux , u2 2 uxx ux uxx u − u2x uxx v 2 vx = −ν = −ν − + , = −ν u2 u u u ν uxxx u − ux uxx 2vvx + , u2 ν and using that u(t, x) is a solution of the heat equation, vxx = −2 uxt u − ux ut ν vxx = −ν + vvx = vt + vvx . 2 u2 Thus v(t, x) solves the viscid Burgers equation. 1 Previously we said that (example 1) if we do not have diffusion the measure in the continuous functions is singular (in the sense that it is supported in an unique function). 58 CHAPTER 4. FLUID DYNAMICS 1.5 1 0.5 0 −0.5 −1 −1.5 0 1 2 3 4 5 6 7 Figure 4.3: Inviscid Burgers equation at different times. Proposition 5. Let v(t, x) be the solution of the viscid Burgers equation (4.4) with initial data f ∈ Cb2 . Then the following representation formula holds Z √ν(W t (x)) 0 −1 v(t, x) = −ν log Ex exp( f (s)ds) . (4.6) ν −∞ x Proof. We have that −1 u(t, x) = exp( ν Z x v(t, s)ds) −∞ solves the heat equation ν uxx 2 Z −1 x u0 (x) = exp( f (s)ds). ν −∞ ut = with initial data We know (chapter 3) that u(t, x) has the following representation Z √ν(W t (x)) 0 √ −1 t u(t, x) = Ex [u0 ( ν(W0 (x)))] = Ex exp( f (s)ds) . ν −∞ We use the formula (4.5) to conclude v(t, x) = −ν log Ex −1 exp( ν Z √ ν(W0t (x)) −∞ f (s)ds) . x (4.7) 4.2. THE D−DIMENSIONAL BURGERS EQUATIONS 59 1 0.8 0.6 0.4 0.2 0 −0.2 −0.4 −0.6 −0.8 −1 −4 −3 −2 −1 0 1 2 3 4 Figure 4.4: Burgers equation with different dissipation rates. 4.2 The d−dimensional Burgers equations We consider now the d−dimensional Burgers equations (now ~v is a vector) ~vt + (~v · ∇)~v = 0 (4.8) ~vt + (~v · ∇)~v = ν∆~v . (4.9) and the viscous case With initial data f~ ∈ Cb2 . First, we suppose that we have a potential solution, i.e. ~v (t, x) = ∇H(t, x). This hypothesis, that phisically implies a irrotational flow, gives us the HopfCole transformation. We have to suppose that f~(x) = ∇H0 (x). The Hopf-Cole transformation is ∇H(t, x) = ~v (t, x) = −ν ∇u(t, x) = −ν∇ log(u(t, x)) u(t, x) And we have that if u(t, x) is a solution of ut = ν ∆u 2 with initial data u0 (x) = exp H0 (x) −ν (4.10) 60 CHAPTER 4. FLUID DYNAMICS then ~v (t, x) is a solution of ~vt = ν ν 1 ∆~v − (~v · ∇)~v = ∆~v − ∇|~v |2 2 2 2 f (x) = ∇H0 (x). Remark 17 The last equality shows us the two possible generalizations of the inertial term, (~v · ∇) and 21 ∇||~v ||22 , of the Burgers equation are the same in the gradient function case. In other cases is not true. Now we use the representation of the heat equation (see chapter 3) √ √ 1 t t u(t, x) = Ex u0 ( ν(W0 (x)))] = Ex exp(− H0 ( ν(W0 (x))))] . ν We conclude using the formula (4.10) √ 1 ~v (t, x) = −ν∇ log E exp(− H0 ( ν(Bt + x)))] . ν (4.11) Now we consider the general case (a non-potential flow). So the initial value is not necessarily a gradient, f~(x) 6= ∇H0 (x). But we suppose that our initial data is smoother. We consider, as the spatial domain, the set Td . This domain is bounded but we do not need stopping times becuase of the periodic boundary conditions. The idea is to start with a inviscid equation and consider particles trajectories with white noise. Finally we consider the noise √ ~. 2νdW With generator √ ( 2ν)2 ∆ = ν∆. 2 Then, taking expectations, we obtain the solution to the viscid equation. Remark 18 We can understand the Itô diffusions as ’random characteristic curves’. We consider the inviscid Burgers equation (4.8). We do not have pressure nor dissipative terms and so the velocity can be transported by the flow. ~ a) is the fluid mapping,2~v (t, X(t, a)) is constant in time, and so If X(t, we have ~v (t, X(t, a)) = f (a). Summarizing, ’we go back to labels and we see there the initial velocity’. This is the method of characteristics. Thus the system ~ ′ (t, a) = ~v (t, X(t, ~ a)), ~v (t, X(t, a)) = f (a) X ~ with initial data X(0, a) = a is equivalent to the Burgers equation (4.8) before the formation of shocks. ~ a) is, if we fix t = s, an homeomorphism between the occupied volumes The flow X(t, at time 0 and s, and, if we fix a, is the trajectory of this particle (see Figure 4.5). 2 4.2. THE D−DIMENSIONAL BURGERS EQUATIONS 61 Figure 4.5: Flow. Considering a as points in the initial volume U0 , while x are points in the time t volume, Ut , we write ~ x) = X(t, ~ a)−1 = a, de modo que ~v (t, x) = f (A(t, ~ x)). A(t, Both volumes U0 and Ut are the torus Td , but we need to discriminate the variables. The transformations between the volumes are X(t, a) = x and A(t, x) = a. ~ a)−1 = A(t, ~ x) = a as the spatial inverse. Remark 19 We write X(t, The spatial inverse exists because of Z t ~ = ∇ · ~v det ∇X ~ ⇒ det ∇X ~ = exp( ∇ · ~v ds) > 0. ∂t det ∇X 0 The idea is to disturb the ODE ~ ′ = ~v X with the previous white noise. We obtain the SDE √ ~. ~ = ~v dt + 2νdW dX We consider ν = 1/2 without loss of generality. The main theorem is Theorem 24 (Burgers equation). Let f~ ∈ C k+1,α , k ≥ 1 be a divergence-free ~ a) ~ t (0) a d-dimensional Wiener process. Let ~v (t, x), X(t, vector field, and W 0 62 CHAPTER 4. FLUID DYNAMICS be solution of the stochastic system ~ ~ dX = ~v dt + dW ~ x) = (X(t, ~ a))−1 A(t, ~ x))] ~v = Ex [f~(A(t, (4.12) ~ with initial data X(0, a) = a and periodic boundary conditions for ~v (t, x) ~ and X(t, a) − I. Then ~v (t, x) is a classical solution of (4.9) (with ν = 1/2), with f~ as initial value. Proof. We define ~ 0t (0)) ~v ω = ~v (t, x + W ~ ω as the solution of and Y ~ ω )′ (t, a) = ~v ω (t, Y~ ω ), Y ~ ω (0, a) = a. (Y Let ~ ω (t, x) B ~ ω. be the spatial inverse of Y We define ~ 0t (0)) = f~(B ~ ω (t, x − W ~ 0t (0))) = f~(A(t, ~ x)). w(t, ~ x−W where the las equality is a consequence of ~ B ~ ω (t, x − W ~ 0t (0))) = Y ~ ω (t, B ~ ω (t, x − W ~ 0t (0))) + W ~ 0t (0) = x; X(t, thus ~ x) = B ~ ω (t, x − W ~ t (0)). A(t, 0 We apply the generalized Itô formula (theorem A.8) to ~ t (0))ω = f~(A(t, ~ x)) w(t, ~ x−W 0 obtaining ~ t (0)) − f~(x) = w ~ ω (t, x − W 0 Z t t − Z 0 0 ~ s (0)) w ~ ω (ds, x − W 0 ~ s (0))dW ~ ∇w ~ ω (s, x − W 0 Z 1 t ~ s (0))ds + ∆w ~ ω (s, x − W 0 2 0 Z t j ω s j t ~ ~ ∂j w ~ (ds, x − W0 (0)), x − W0 (0) . + 0 Taking expectations in the previous equation the left hand side resulting is 4.2. THE D−DIMENSIONAL BURGERS EQUATIONS 63 ~v (t, x) − Ex [f~(x)]. The right hand side is Z t Z t 1 ω s ω s ~ ~ Ex w ~ (ds, x − W0 (0)) + Ex ∆w ~ (s, x − W0 (0))ds 2 0 0 because the stochastic integration term vanishes after take the expectation, and the quadratic variation term is zero because of w(t, ~ x) ∈ C 1 in time (the differential is given by the transport equation) and w(t, ~ x) ∈ C k+1,α (as ~ smooth as f ) in space, and so the function and the derivatives to the k−th order are of bounded variation. We have that Z t Z 1 t 1 ω s ~ ∆w ~ (s, x − W0 (0))ds = ∆~v (s, x)ds. Ex 2 0 2 0 We need the convective term. But ~ ω (t, x)) w ~ ω (t, x) = f (B solves w ~ tω (t, x) + (~v ω (t, x) · ∇)w ~ ω (t, x) = 0. with the initial data f~. The term we obtain is Z t Z t ω s ω s ~ ~ Ex w ~ (ds, x − W0 (0)) = Ex w ~ t (s, x − W0 (0)) 0 0 Z t ~ 0t (0)) · ∇) Ex [−(~v ω (s, x − W = 0 ω ~ t (0))] × w ~ (s, x − W 0 Z t = − (~v (s, x) · ∇)~v (t, x)ds. 0 We put all these calculations together and take the time derivative. We suppose we have a solution to the stochastic system, and then we have the representation formula. Conversely, if ~v (t, x) is a classical solution of the equation (4.9) the stochastic system has a solution (see [CI],[Iy] and [Iy2]). The idea is, if ~v is a classical solution, the equation ~ = ~v dt + dW ~ dX ~ Moreover, has solution, and so there exists A. ~ x))] ~v (t, x) = Ex [f~(A(t, 64 CHAPTER 4. FLUID DYNAMICS because of a result (about stochastic partial differential equations) contained in [CI] and from the uniqueness of classical solutions for the viscid Burgers equations. If we show that our stochastic system has a solution, then we havea local in time solution to the equation (4.9). It is a local solution because we use a fixed point method to show the existence of solution to the system (see [Iy]). Remark 20 The boundary conditions are the natural ones because of ~ ~ X(0, a+L~ej ) = a+L~ej = X(0, a)+L~ej . And this is the periodicity condition ~ a) − I. for X(t, Remark 21 We use the notation in [CI] to remark the similarities between this section and the following one. 4.3 The incompressible Navier-Stokes equations In this section we show a representation formula for the classical solution of the Navier-Stokes equations (4.1) as a stochastic system and a functional integral. If we have a classical solution of (4.1) then the solution verifies the probabilistic representation. And conversely, if we have a solution of the stochastic system then the ~u(t, x) is a classical solution of Navier-Stokes equations. We use this fact to show the existence (local in time) of classical solution for the Navier-Stokes. We need some preliminary results.3 The fields ~v ∈ (L2 )d ∩ (C ∞ )d can be written in a orthogonal way as a divergence-free field and a gradient field. So we define Definition 8. We write P for the Leray-Hodge projector, i.e. this operator, given a field, returns the divergence-free component. P : (L2 )d ∩ (C ∞ )d 7→ S where S denotes the set of the divergence-free fields (the solenoidal fields). Proposition 6 (Eulerian-Lagrangian formulation). Let k ≥ 0 and f~(x) ∈ ~ C k+1,α such that ∇· f(x) = 0. Then ~u(t, x) satisfies the incompressible Euler equations (4.2) with an initial datum f~(x) if and only if the pair of functions ~ a) satisfy the stochastic system ~u(t, x), X(t, ~ ′ (t, a) = ~u(t, x) X ~ x) = X ~ −1 (t, a) A(t, ~ t (t, x)f~(A(t, ~ x))] ~u(t, x) = P[(∇A) (4.13) (4.14) (4.15) ~ with initial data X(0, a) = a. 3 We write P the Laray-Hodge projector in the divergence-free fields. Recall that P denotes the probability. 4.3. THE INCOMPRESSIBLE NAVIER-STOKES EQUATIONS 65 See [C] for the proof. Lemma 5. Let ~u(t, x) be a velocity field. The commutator [∂t + (~u · ∇), ∇] is [∂t + (~u · ∇), ∇] = −(∇~u(t, x))t ∇ Proof. [∂t + (~u · ∇), ∇]f~(t, x) = (∂t + (~u · ∇))∇f~ − ∇(∂t + (~u · ∇))f~(t, x) = (~u · ∇)∇f~(t, x) − (~u · ∇)∇f~(t, x) − (∇~u(t, x))t ∇f~(t, x) ~ a) and Lemma 6. Given a Lipschitz, divergence-free field ~u(t, x), and X(t, ~ x) functions defined as A(t, ~ ′ (t, a) = ~u(t, x) X ~ x) = X ~ −1 (t, a) A(t, ~ X(0, a) = a We define ~v (t, x) as the solution of the evolution equation (∂t + (~u · ∇))~v (t, x) = ~z(t, x) for certain field ~z, with initial datum ~v0 . Then if we define w(t, ~ x) as ~ t (t, x)~v (t, x)] w(t, ~ x) = P[(∇A) we have that w(t, ~ x) is the solution of ~ t )~z (t, x) (∂t + (~u(t, x) · ∇))w(t, ~ x) + (∇~u(t, x))t w(t, ~ x) + ∇p(t, x) = ((∇A) ∇ · w(t, ~ x) = 0 w(0, ~ x) = P~v0 (x) See [CI] for the proof. In the proof we use the particular case ~v (t, x) = ~ ~ f (A(t, x)) and ~z = 0. Theorem 25 (Navier-Stokes equations). Let f~ ∈ C k+1,α,k ≥ 1 be a given ~ t (0) a d-dimensional Wiener process. Let the pair solenoidal field, and W 0 ~ a) a solution to the stochastic system ~u(t, x), X(t, ~ = ~udt + dW ~ dX −1 ~ = X ~ A ~ t f~(A)]] ~ ~u = Ex [P[(∇A) (4.16) (4.17) (4.18) ~ ~ − I are periodic in the with initial datum X(a, 0) = a and such that ~u and X boundary. Then ~u satisfies the incompressible Navier-Stokes equation with f~ as initial data. 66 CHAPTER 4. FLUID DYNAMICS Proof. We define uω as ~ t (0)). ~uω (t, x) = ~u(t, x + W 0 ~ ω be the solution of Let Y ~ ′ (t, a) = ~uω (t, Y ~ ω (t, a)), Y ~ ω (0, a) = a. Y ~ ω (t, x) be the spatial inverse of Y ~ ω . We observe that Let B ~ B ~ ω (t, x − W ~ 0t (0))) = Y ~ ω (t, B ~ ω (t, x − W ~ 0t (0))) + W ~ 0t (0) = x X(t, so ~ x) = B ~ ω (t, x − W ~ 0t (0)). A(t, If we write θx h(y) = h(y − x) for the traslation then A = θW ~ t (0) B. 0 We define wω as ~ ω )t (f~(B ~ ω (t, x)))]. wω (t, x) = P[(∇B ~ x)) Applying the lemma 6 in the particular case ~z = 0 and ~v (0, x) = f~(A(t, we have (∂t + (~uω (t, x) · ∇))w ~ ω (t, x) + (∇~uω (t, x))t w ~ ω (t, x) + ∇q ω (t, x) = 0 ω ∇·w ~ (t, x) = 0 w(0, ~ x) = Pf~(x). (4.19) Using the definition of ~u, ~ t f~(A(t, ~ x))]] ~u = Ex [P[(∇A) ~ t f~(θ t = Ex [P[(∇θ t B) W0 (0) ~ W0 (0) B(t, x))]] ~ t f~(B(t, ~ x)))]] = Ex [P[θW0t (0) ((∇B) ~ t f~(B(t, ~ x))])] = Ex [θW0t (0) (P[(∇B) = Ex [θW0t (0) wω ]. The hypothesis f~ ∈ C k+1,α and the existence of solution to the stochastic system (4.18) theorem (see the next section for the proof) gives us the regularity needed to apply the generalized Itô formula (theorem A.8). Recall that w ~ ω is the F in the theorem A.8. We need C 2 in space and C 1 in time. By the lemma 6 we have the C 1 in time condition, and the existence to the system (4.18) theorem gives us the regularity in space. In addition, ~ t (0) is a martingale doing the g(t) in the theorem. x−W 0 4.3. THE INCOMPRESSIBLE NAVIER-STOKES EQUATIONS 67 ~ t (0)), Applying the generalized Itô formula (theorem A.8) to w ~ ω (t, x − W 0 Z t ~ s (0)) ~ t (0)) − f~(x) = w ~ ω (ds, x − W w ~ ω (t, x − W 0 0 0 Z t ~ s (0))dW ~ ∇w ~ ω (s, x − W − 0 0 Z 1 t ~ s (0))ds + ∆w ~ ω (s, x − W 0 2 0 Z t j ω s j t ~ (0)), x − W ~ (0) . ∂j w ~ (ds, x − W + 0 0 0 w ~ω Because of the regularity of we have that the quadratic variation term vanishes. Moreover after taking the expectation Ex the stochastic integration term also disappear. Thus we obtain Z t Z t ω s ~ (0)) + 1 ∆~u(s, x)ds. ~u(t, x) − f~(x) = Ex w ~ (ds, x − W 0 2 0 0 The term with w ~ ω is Z t Z t ω s ω s ~ Ex w ~ (ds, x − W0 (0)) = Ex w ~ t (s, x − W0 (0))ds 0 0 Z t ~ s (0)) = −Ex (~u(s, x) · ∇)w ~ ω (s, x − W 0 0 ~ s (0)) + (∇~u(s, x))t w ~ ω (s, x − W 0 ~ 0s (0))ds + ∇q ω (s, x − W Z t = − [(u · ∇)u + ∇pds] 0 where 1 2 ω |u| + Ex [θW ~ t (0) q ]. 0 2 Finally, the incompressibility holds because of the incompressibility of w ~ω ~ ω. and ~u = Ex θW ~ t (0) w p= 0 Remark 22 In the previous proof we used the existence theorem to (4.18) (theorem 26 in the following section). In [Iy] we can see the result for a substance transported by the fluid. This is Proposition 7. (Transport) Let ~u ∈ C 1 be a velocity field of a fluid and Θ(t, x) a classical solution of Θt (t, x) + (~u(t, x) · ∇)Θ(t, x) − ν∆Θ(t, x) = 0, Θ(0, x) = f (x). 68 CHAPTER 4. FLUID DYNAMICS Then ~ x))] Θ(t, x) = Ex [f (A(t, where A is as in the theorem 25. There is a result for the vorticity. ~ (t, x) be the vorticity of a fluid, then Proposition 8 (Vorticity). Let V ~ (t, x) = Ex [(∇X ~V ~0 )(A(t, ~ x))]. V If d = 2, then ~ (t, x) = Ex [V ~0 (A(t, ~ x))]. V Remark 23 We obtain these formulas taking the expectation in the Euler case formulas. Remark 24 We can use the proposition 8 to obtain a second version of the theorem 25 using the Bioy-Savart law. Remark 25 We can do the samePfor a operator with anisotropic diffu2 sion, i.e. a more general operator as di,j=1 ai,j (x) ∂x∂j ∂xi . 4.4 Proof of local existence for Navier-Stokes In this sectio we show the local existence for the stochastic system and so, the local existence of classical solution of Navier-Stokes equations (4.1). Theorem 26 (Existence for the stochastic system). Let f~ ∈ C k+1,α, k ≥ 1, a divergence-free field. Then there is a time T = T (L, ||f~||C k+1,α , k, α) but independent of the viscosity ν, and a pair of functions ~u, λ ∈ C([0, T ], C k+1,α ) ~ = I + λ satisfies the system (4.18). In addition, there exist such that ~u, X Λ such that ||~u(t)||C k+1,α ≤ Λ. Remark 26 The norm in C k,α is defined as X X Lα |D m ~u(x, t) − D m~u(y, t)| . ||~u||k,α = L|m| sup |D m ~u| + Lk sup |x − y|α x∈Ω x,y∈Ω |m|≤k |m|=k For the norm in C([0, T ], C k,α ) we take the supremum ||~u||C([0,T ],C k,α ) = sup ||~u||k,α . 0≤t≤T We need some definitios and some bounds. See [Iy] for the proof of these bounds. ~ : C k,α × C k+1,α 7→ C k,α as Definition 9. We define the Weber operator W ~ (~v , ~l) = P[(I + (∇~l)t )~v ]. W 4.4. PROOF OF LOCAL EXISTENCE FOR NAVIER-STOKES 69 Proposition 9. If k ≥ 1, and ~l1 , ~l2 , ~v1 , ~v2 ∈ C k,α are functions such that ||∇~li ||k−1,α ≤ C then ~ (~v1 , ~l1 ) − W ~ (~v2 , ~l2 )||k,α ≤ c(||~v2 ||k,α ||∇~l1 − ∇~l2 ||k−1,α + ||~v1 − ~v2 ||k,α ). ||W ~ (v, l) ∈ C k,α and we have Lemma 7. Let k ≥ 1 and ~v , ~l ∈ C k,α. Then W ~ (v, l||k,α ≤ c(1 + ||∇l||k−1,α )||v||k,α . ||W ~ a) be a solution of the system Lemma 8. Let ~u ∈ (C[0, T ], C k+1,α ) and X(t, ~ − I, l = A ~ − I and Λ = supt (||~u||k+1,α ). Then (4.18), and consider λ = X there exists c = c(k, α, Λ) such that the following bounds hold ||∇λ||k,α ≤ cΛt exp(cΛt/L), L cΛt ||∇~l||k,α ≤ exp(cΛt/L). L Lemma 9. Let ~u1 , ~u2 ∈ C([0, T ], C k+1,α ) such that supt ||~ui ||k+1,α ≤ Λ and ~ 1, X ~ 2, A ~1, A ~ 2 be functions defined as in (4.18). Then there exists a time let X T = T (k, α, Λ) and a constant c = c(k, α, Λ) such that the following bounds hold Z t ~ ~ ||~u1 − ~u2 ||k,α ||X1 − X2 ||k,α ≤ c exp(cΛt/L) 0 ~1 − A ~ 2 ||k,α ≤ c exp(cΛt/L) ||A Z t 0 ||~u1 − ~u2 ||k,α for all t ∈ [0, T ] Proof (of the theorem). Consider a time T (we will take T small, see below) and a number Λ (we will take Λ big). We consider the spaces k+1,α ~ U = ~u ∈ C([0, T ], C ), ∇ · ~u = 0, ~u(0, x) = f (x), ||~u||k+1,α ≤ Λ and 1 k+1,α ~ ~ ), ||∇l||k,α ≤ L = l ∈ C([0, T ], C ∀t ∈ [0, T ]l(0, x) = 0 . 2 ~ u , λu = If ~u in U the system (4.18) has a solution. Thus we can define X ~ ~ ~ Xu − I and lu = Au − I. We consider the operator W : U → 7 U ~ (f~(A ~ u ), ~lu ). W (~u) = Ex W 70 CHAPTER 4. FLUID DYNAMICS We want to show that W is Lipschitz with respect to the norm ||~u||U = sup ||~u||k,α t and then, if we take T small enough W will be a contraction and by Banach’s fixed point theorem it will have a fixed point. The previous results gives us that, if we take Λ proportional to k+2 3 , ||f~||k+1,α 2 ||W (~u1 ) − W (~u2 )||k,α cΛ exp(cΛt/L) ≤ L Z t 0 ||~u1 − ~u2 ||k,α and so if T = T (k, α, L, Λ) is small, the operator W is a contraction. Applying the Banach’s fixed point theorem and taking into account that U is closed we conclude that the sequence given by ~un+1 = W (~un ) converges to a function ~u in the norm C k,α. ~u is a fixed point of the operator W and so a solution of (4.18). ~un is a strongly convergent in the k norm, and converges weakly in the k + 1 norm, so the limits are the same. So ~u is a C k+1,α function and we have the bound ||~u(t)|| ≤ Λ because of ~u ∈ U. As we have a solution of (4.18), applying the theorem 25 a local in time classical solution of (4.1). Chapter 5 Differential games and equations We have studied in the previous chapters how the paths of certain Markov processes are related (they are characteristic curves) with elliptic and parabolic equations. We showed (chapter 3, section 2) that for some semilinear equations we can find a representation formula if we take more difficult Markov process. In this chapter we study the relationship between certain differential games (we call them tug of war and they are (most of times) Markov process) and certain equations, the 1-laplacian, the p−laplacian and the ∞−laplacian. We prove that the games’ positions are the characteristic curves for this operators. In this chapter we follow closely [PSSW],[Ob],[Ev3],[BEJ], [ACJ], [KS1] and [KS2] 5.1 The operators Definition 10. The operator d X uxi uxj ux ,x ∆∞ u = |∇u|2 i j i,j=1 is the ∞-laplacian. The operator ∆p u = ∇ · (∇u|∇u|p−2 ) is the p-laplacian. The operator ∆1 u = ∇ · is the 1-laplacian. 71 ∇u |∇u| 72 CHAPTER 5. DIFFERENTIAL GAMES AND EQUATIONS We remark that ∆1 is an orthogonal to the gradient diffusion. Indeed, if we consider the d = 2 case we have that the non-divergence form of the operator is given by the matrix A= 1 |∇u| u2x1 |∇u|2 ux 1 ux 2 |∇u|2 1− ux 1 ux 2 |∇u|2 u2x2 1 − |∇u| 2 . At each point x we consider the basis given by ∇u and ∇u⊥ . If we calculate A∇u we see that it is 0, and so there is no diffusion in this direction. Conversely the ∞−laplacian is a diffusion only in the gradient direction. The matrix now is A= u2x1 |∇u|2 ux 1 ux 2 |∇u|2 ux 1 ux 2 |∇u|2 u2x2 |∇u|2 . We have that A∇u⊥ = 0, and so there is diffusion only in the gradient direction. The previous calculus are only informal because if ∇u = 0 they do not make sense. However they gives us some intuition. Remark 27 The operators are used in computer vision, to conserve the contours (∆1 ) or to difuminate (∆∞ ). 0.75 0.7 0.65 0.6 0.55 0.5 0.5 0.5 0 0 −0.5 −0.5 Figure 5.1: An ∞−harmonic function. 73 5.2. THE GAMES The geometrical interpretation of u satisfying ∆1 u = 0 is that the level curves of u has mean curvatures 0. The variational interpretation is that u is the minimum of Z |∇u| J(u) = U with given boundary data. For the p−laplacian the variational interpretation is that u is the minimum of Z |∇u|p J(u) = U with given boundary data. The variational interpretation of ∆∞ u = 0 in U with Lipschitz boundary values is that u satisfy the condition LipV (u) = Lip∂V (u) ∀V ⊂ U where U, V are domains and LipV (u) indicates the Lipschitz constant in the domain of the function u, V . This kind of functions are called absolutely minimizing Lipschitz extension. During this chapter we consider U a boundad domain, a continuous function g defined in ∂U and we study the following equations Lu = 0 in U, u = g in ∂U where L is a differential operators before. 5.2 The games All games considered are two players and zero-sum games, so a player pays to the another player. We suppose that the player 2 pays player 1 the appropiated quantity (if negative he earns). In all games the players move a token in the considered domain, and when the token hits the boundary the game ends and the player 2 pays. Each game is determined by the positions of the token at each turn xk . Both players have strategies, but it seems reasonable that the ’good’ strategies will be markovian. The p-laplacian game and the 1−laplacian game are studied briefly. We are interested in the ∞−laplacian. We suppose the domain U is as regular as we need to take the needed limits. 5.2.1 ’Tug of war’ We are going to study this game in full detail. Let U ⊂ Rd , x0 ∈ U be as before and g be a Lipschitz function supported on the boundary of U . Each player choose a vector ~aik ∈ B0 (ε) (k is the turn and i is the player). Then a fair coin is thrown to decide the player who can move the token. If the 74 CHAPTER 5. DIFFERENTIAL GAMES AND EQUATIONS player 1 wins the coin then the new position will be xk+1 = xk + ~a1k . We define the history hk = (x0 ,~a0 , x1 ,~a1 ...) where ~ak is the vector for the player who wins the k-th turn. Let Hk be the space of all possible until the k−th turn and let H∞ be the space fo all possible histories. We observe that the space Hk is a product space. We can understand the payoff function as a function g : H∞ 7→ Rd . Usually the vector ~aik depends on the position xk , i.e, the process will be markovian, however we consider a more general dependency. We consider that the vector depends on the full previous history. We define a strategy Ski (i = 1, 2 is for the players) as a function Ski : Hk 7→ B0 (ε). The function gives the following movement for the player i. We define Si = {Ski }k . Then the initial point x0 and the strategies for both players define a probability in H∞ (use the Kolmogorov’s extension theorem to prove it). If we write xτ ∈ ∂U for the point where the game ends and the given strategies S1 , S2 we define the expected payoffs for both players as Vx0 ,i (S1 , S2 ) = ExS01 ,S2 [F (xτ )] if the games ends almost surely. If the game does not end almost surely Vx0 ,1 (S1 , S2 ) = −∞ and Vx0 ,i (S1 , S2 ) = ∞.1 We define the game’s value (discrete) for player 1 as uε1 = sup inf Vx0 ,1 (S1 , S2 ). S1 S2 Roughly speaking, inf S2 Vx0 ,1 (S1 , S2 ) is the minimum quantity that player 1 wins if we suppose that player 2 plays optimally. So, the supremum is that player 1 maximizes the money. For the player 2 the definition is uε2 = inf sup Vx0 ,2 (S1 , S2 ). S2 S1 Roughly speaking supS1 Vx0 ,2 (S1 , S2 ) is the maximum quantity that player 1 obligates to player 2 to pay. And the infimum is that player 2 minimizes this quantity if player 1 plays optimally. Thus, player 1 maximizes his worst case and player 2 minimizes his worst case. We have uε1 (x) ≤ uε2 . If these quantityes are the same then the game (discrete) has a value. We write uε for the value. We expect that taking limit ε → 0 our value (for the discrete game) converges, in a certain sense, to u solution of ∆∞ u = 0, 1 As in the chapter 2, τ is a stopping time. u|∂U = g. (5.1) 75 5.2. THE GAMES This equation has a dynamic programming principle useful for the numerical schemes (see appendix C and [Ob]). This result is, in a certain sense, the mean value property for the infinite harmonic functions. Lemma 10. We consider the tug of war game without running payoff (f = 0). Then the (discrete) game’s value function, u(x) = uε1 (x), satisfies 1 u(x) = ( sup u(y) + inf u(y)) 2 y∈Bx (ε) y∈Bx (ε) (5.2) for all x ∈ U . If the game does not end then u(x) = −∞. The same for v(x) = uε2 (x), if we consider that if the game does not end then v(x) = ∞. Proof. To prove it we have to take into account the possible coin results. Remark 28 We want to remark the similarities with the chapter 2. However in the second chapter the process are pure diffusions, without strategies. In the chapter 2 we had functional integral and now we integrate in the histories. Remark 29 We expect that the good strategies will be markovian but we do not restrict to the markovian ones. 5.2.2 Approximations by SDE to ∆∞ In [BEJ] we can read that the stochastic differential equation ~ ~ ~ = η(t)dt ~ , X0 = x0 dX + ζ(t)d W is related with the ∆∞ . Let U ⊂ Rd , ε > 0, g and x0 ∈ U be as before. The ~ The total payoff function (final player 1 choose ~ η , and player 2 choose ζ. and running payoff) for player 1 is g(X0τx (x)) − Z 0 τx 2 ~ + ε |~η (s)|2 ds ~η (s) · ζ(s) 4 where τx is the ’hit the boundary’ stopping time. This stochastic game has a (discrete) value that converges to u solution of (5.1). In [PSSW] we can see another SDE giving the infinity-laplacian. Let u ∈ C 2 be a infinity-harmonic function. We define ~ t (x)) = |∇u(Y ~ t (x))|−1 ∇u(Y ~ t (x)) ~r(Y 0 0 0 and ~ t (x))|−2 D 2 u(Y ~ t (x))∇u(Y ~ t (x)) − P ~ t (x)) = |∇u(Y ~s(Y 0 0 0 0 ~ t (x)) (and so orthogonal to the gradient). where P is its projection in ∇u(Y 0 76 CHAPTER 5. DIFFERENTIAL GAMES AND EQUATIONS We define the SDE ~ 0t (x) = ~r(X ~ 0t (x))dW + ~s(X ~ 0 (x))dt, dX X0 = x. ~ t (x)) we obtain Applying the Itô formula to u(X 0 Z t Z t 1 s t ~ s (x))dW. ~ ~ ∇ut · ~r(X ∆∞ u(X0 (x))ds + u(X0 (x)) − u(x) = 0 2 0 0 Taking expectations we conclude ~ t (x))] = u(x). Ex [u(X 0 5.2.3 Existence of game’s value for the ’Tug of war’ In this section we proof the existence of a game’s value for the tug of war game (discrete) without running cost. Theorem 27. Let U ⊂ Rd be a domain and 1 >> ε > 0 a fixed number. We consider the tug of war game with f = 0 and g bounded below (or above) in ∂U a Lipschitz function. Then uε1 = uε2 and the game has a value. Proof. If g is bounded above and not below we consider −g and we change the players. So we can restrict us to the g bounded below case. We have to see that uε2 ≤ uε1 . Recall that player 1 is able to finish the game almost surely, because player 1 can obtain a number as big as player 1 need of heads (or tails). Thus uε1 ≥ inf x∈∂U g(x). Let x0 , x1 , ... be the game’s positions at different turns. We write uε = uε1 . We consider the oscillation δ(x) = sup |uε (y) − uε (x)|. y∈Bx (ε) We define the set X0 = {x ∈ U : δ(x) ≥ δ(x0 )} ∪ ∂U and the index jn = maxj≤n xj ∈ X0 . This index gives us the las turn in the set X0 . Let vn = xjn be the last position in X0 . X0 is the set of points with oscillation bigger than the initial point.. Thanks to the dynamic programming principle we have 2uε (xn ) = sup y∈Bxn (ε) uε (y)+ inf y∈Bxn (ε) uε (y) ⇔ inf y∈Bxn (ε) {uε (xn )−uε (y)} = δ(xn ), and so if the players choose the strategy of maximize (player 1) or minimize (player 2) the function uε , the δ function will not decrease because it has, at least, the same oscillation of the previous position. Thus, with the previous strategies, the game is always in X0 . 77 5.2. THE GAMES We consider the following strategy for player 2: if vn 6= xn , i.e. we are not in X0 , player 2 moves to y such that y is the point of minimum distance between xn and X0 . When xn = vn player 2 choose the new position in such a way uε is minimum. For player 1 we consider all the possible strategies and let the game begin. We remark that player 2 does not play in an optimal way, because X0 contains the boundary, where the game ends and player 2 has to pay. For player 2 it is better that the game does not end, because in this case his payoff is uε2 = ∞. We mention that this strategy is Markovian. Let d be the distance measured in ε−steps, then we define dn = d(xn , vn ) the distance where we consider that we have to pass through the previous positions and mn = uε (vn ) + δ(x0 )dn . We have uε (xn ) = uε (vn ) + (uε (xjn +1 ) − uε (vn )) + (uε (xjn +2 ) − uε (xjn +1 )... (5.3) + ...(uε (xn ) − uε (xn−1 ) n X ε ≤ u (vn ) + δ(xk ) (5.4) ≤ mn (5.6) (5.5) k=jn +1 (because they are not in X0 .) mn is a supermartingale. Indeed, we suppose that xn ∈ X0 and player 1 moves. There are two possibilities, xn+1 is or is not in X0 . If xn+1 ∈ X0 then mn+1 = uε (vn+1 ) − uε (xn ) + uε (xn ) ≤ uε (xn ) + δ(xn ) = mn + δ(xn ). If xn+1 ∈ / X0 then mn+1 = uε (vn+1 ) + δ(x0 )dn+1 ≤ uε (xn ) + δ(xn ) ≤ mn + δ(xn ). Suppossing now that xn ∈ X0 and player two moves we have uε (xn+1 ) = uε (xn ) − δ(xn ) = uε (vn ) − δ(xn ) and in the case xn+1 ∈ X0 the previous equation is mn+1 = mn − δ(xn ) ≤ mn − δ(x0 ). If xn+1 ∈ / X0 we have a contradiction, because of δ(xn ) ≥ δ(x0 ) > δ(xn+1 ) 78 CHAPTER 5. DIFFERENTIAL GAMES AND EQUATIONS and this is not possible because of the dynamic programming principle. This principle gives us that the oscillation in xn+1 (if we choose it maximizing or minimizing uε ) is not decreasing. If we suppose that xn ∈ / X0 and player 2 moves we have, using the previously defined strategy and if vn+1 6= xn+1 , the following inequality mn+1 = uε (vn+1 ) + δ(x0 )dn+1 ≤ uε (vn+1 ) + δ(x0 )d(vn+1 , xn ) − δ(x0 )d(xn+1 , xn ) ≤ mn − δ(x0 ) Si vn+1 = xn+1 then mn+1 = uε (xn+1 )±uε (xn ) ≤ uε (xn )+δ(xn ) ≤ mn +δ(x0 ) (using (5.6) and X0 ). We consider the last case: xn ∈ / X0 and player 1 plays. If player 1 enters in X0 then mn+1 = uε (un+1 )±uε (xn ) ≤ uε (xn )+δ(xn ) ≤ mn +δ(x0 ) (using (5.6) and X0 ). If player 1 does not enter in X0 then mn+1 = uε (vn+1 ) + δ(x0 )d(vn+1 , xn+1 ) ≤ uε (vn ) + δ(x0 )d(vn , xn ) + δ(x0 )d(xn , xn+1 ) ≤ mn + δ(x0 ). Putting together all the previous calculations we have that if player 2 moves mn+1 ≤ mn − δ(x0 ) and if player 1 moves mn+1 ≤ mn + δ(x0 ). Thus 1 (5.7) E[mn+1 |m0 , m1 ...mn ] ≤ mn + (δ(x0 ) − δ(x0 )) = mn . 2 Using the martingale convergence theorem, if τx0 is the previous stopping time if the game starts at x0 , we know that there exists the limit limn→∞ mmin(n,τx0 ) . This and that mn+1 ≤ mn − δ(x0 ) implies that the game ends almost surely. Then the expected payoff with this strategy for player 2 is Ex0 [uε (xτx0 )] = Ex0 [ lim uε (xmin(τx0 ,n) )] n→∞ ≤ Ex0 [mmin(τx0 ,n) )] (using (5.6) and Fatou) ≤ m0 = uε (x0 ) (supermartingale) 79 5.2. THE GAMES This is better than uε2 and so uε2 ≤ uε1 . The oscillation can be zero. In this case the strategy for player 2 is advance straightformward to the boundary until arrive to some point, x′0 , with non-zero oscillation (but uε (x0 ) = uε (x′0 ). Then the strategy becomes the previous one. We need a convergence to the continuous game’s value theorem. Theorem 28. Let g be a bounded below function. Then the continuous game’s value, u, exists and the following holds ||u − uε ||∞ → 0 if ε → 0. In addition u is continuous. And this game’s value is a solution of (5.1). Theorem 29. Let U ⊂ Rd be a bounded domain. Let g be a Lipschitz function supported on the boundary. Then u, the continuous game’s value, is the unique viscosity solution of (5.1). 5.2.4 ’Tug of war with noise’ We consider now the ’Tug of war with noise’ game. In this case the operator is the p-laplacian. Let U ⊂ Rd ,x0 ∈ U and g be as before (f is now identically 0). We measure, µ, uniform in the sphere with radious p consider a probability −1 r = (d − 1)q/p (where p + q −1 = 1) in the orthogonal to ~e1 hyperplane. We define µ~v (S) = µ(Ψ−1 (S)) where Ψ(~v ) = ~e1 .2 At each turn k a fair coin is thrown. This coin gives the turn to each player. The player having the turn choose ~vk , with norm least or equal to ε. The new position is xk = xk−1 + ~vk + ~zk , where ~zk is a random vector with respect to µ~vk . If we are at distance to the boundary least or equal to (1+r)ε the player having the turn has to move to the boundary to a point xk with |xk − xk−1 | ≤ (1 + r)ε, and the game ends. We define uε1 (x) and uε2 (x) as the minimum expected player’s payoffs if the game starts at x0 = x. If they are the same then the game has a value. We suppose that the game (discrete) has a value, then the limit at each point u(x) = limε→0 uε1 (x) is the function indicating the minimum expected value for the payoffs if the game (continuous) start at x0 = x. We have the following result: the function u(x) verifies ∆p u = 0, u|∂U = g (5.8) Remark 30 Show that the game (discrete) has a value, take the limit and see what operators we obtain are results. See [KS1], [KS2] and [PS] for the proof. 2 See [PS] for more details about this probability measure. 80 CHAPTER 5. DIFFERENTIAL GAMES AND EQUATIONS Figure 5.2: The posible positions for the ’Tug of war with noise’ game. 5.2.5 Spencer game We start with the ’Spencer game’. In this case the operator is the 1-laplacian. Given U a domain, x0 ∈ U the point where we start the game, g a continuous function defined in the boundary of U , this function will be our final payoff, and f : U 7→ R the running payoff (at each movement the player who has the turn pays). So, if g = 0 and f is a positive constant, c, the payoffs at the turn k are f (xk ) = c and in the case that the token hits the boundary, g(xk ) = 0. In this case player 1 want to maximize the number of steps to hit the boundary. On the other hand player 2 want to arrive to the boundary as soon as possible to pay the minimum quantity. Each turn k player 2 choose a vector, ~vk , with a fixed norm ε. Player 1 choose a direction to this vector, σk ∈ {1, −1}. At each turn the new position is xk = xk−1 + σk~vk . We define uε1 (x) and uε2 (x) the minimum expected values of the player’s payoffs if the game starts at x0 = x.3 When they are the same quantity we say that the game has a value, u(x0 ). We can understand u as the money the players pay to the casino to enter the game. It is the expected payoff, so player 1 want to maximize and player 2 want to minimize this value. To obtain the differential operator we have to take the limit ε → 0. 3 This definition is not rigorous. See the ∞−laplacian case for a rigorous one. 81 5.2. THE GAMES We suppose the game (discrete) has a value. Then the limit at each point u(x) = limε→0 uε1 (x) is the function indicating the minimum payoff each player hope if the game (continuous) starts at x0 = x. We suppose that player 2 choose the gradient of u direction, thinking in minimize u all it would be possible. Then the player 1 take positive sign and in this new position the player 2 pays more. So we expect that a (good) player 2 choose only the orthogonal to the gradient of u direction. We have the following result: the function u(x) verifies 1 − ∆1 u = f, 2 5.2.6 u|∂U = g (5.9) Other games In [KS1] and [KS2] the authors study games for non-linear parabolic or elliptic equations. For example we explain the game leading to the backward heat equation. We consider the backward Cauchy problem ut (t, x) + uxx (t, x) = 0, u(T, x) = f (x). As before, there are two players, a initial point x0 and a fixed ε. We fix 0 < t < T . Player 1 choose a number α. √ Then, knowing the player 1’s α, player 2 choose b = ±1. Plyer 1 pays 2εαb. Then the time, that when 2 we started was √ t, is t + ε , and the token’s position, originally at x0 , is at xk = xk−1 + 2εb. The game continues until the final time T . Then player 1 has a payoff f (x(T )), where x(T ) is the final token’s position. Player 1 wants to final payoff minus the running cost will be maximum. Player 2 wants to minimize this quantity. The player 1’s value function, uε1 (t, x), is defined as the maximum of the final payoff minus the running cost if the game has an initial time t and a initial position x0 = x. This function converges to the solution of the backwar Cauchy problem for the heat equation. This games have a dynamic programming principle (see [KS1]). The proof (see [KS1]) use the dynamic programming principle and the Taylor formula. In these papers, the authors study the relationship between the economy and these games. Chapter 6 Numerical experiments In the introduction we talked about the applications of these representation formulas to the numerical methods. But first we have to simulate the paths of a given SDE. To do this we apply an explicit Euler method. So, if we have the 1D diffusion given by dY = b(Y )dt + σ(Y )dW and a initial value Y0 . Our method do Y (n + 1) = b(Y (n))T /N + σ(Y (N ))(W (t(n + 1)) − W (t(n)). Now we can apply a Monte-Carlo method to the elliptic or parabolic equations: Let U be the square centered at the origin with side 2. Our boundary value will be g(x) = x1 . The idea is to simulate a big number of brownian paths started at the same point x, look for the boundary points hitted and take the mean of g evaluated at these points. This will be our u(x). In the figure 6.1 we can see the result of an experiment with 100 points in the grid, a time step of 1/9 and 100 diffusions per point. In my PC (1.73 GHz) the time has been 32 seconds. For the parabolic equations (we consider the Cauchy problem) we can do the same. Fix t, then for a given point x we consider a big number the diffusions started at x and we see the value of f at the final point (at time t) of the diffusions. Then we take the mean and this will be our numerical approximation. For example if we consider the heat equation with initial value a characteristic of a given set we have the results (with the same number of points in the grid and same time step) shown in the figures 6.2 and 6.3. We expect that the error will decay if we take a greater number of diffusion, a greater number of points in the grid and a smaller time step, however this is not the case always as we will see. 82 83 Solucion 1.5 1 0.5 0 −0.5 −1 −1.5 1 1 0.5 0.5 0 0 −0.5 −0.5 −1 −1 Figure 6.1: The numerical solution. In [IN] we can read how use the Monte-Carlo method to approximate the solution of the Navier-Stokes and Burgers equations. The idea is to replace the flow X by M copies of it, each driven by an independent Wiener process, and replace the expected value by the mean. This method gives us a second system, the numerical one, which we call Monte-Carlo system. This method is valid for the Navier-Stokes equations in 2D (we know the existence to Euler equations in 2D), however is not valid for the onedimensional Burgers equation. The reason is that the Monte-Carlo system (the numerical one) is dissipative only for short time. When the Monte-Carlo system stops dissipating energy, then the nonlinearity forces the Monte-Carlo system to shock. The really amazing result is that if we solve the Monte-Carlo system for short time, then replace the initial data with the solution in this small time, and restart the procedure this way we obtain the real solution, without unreal shocks!. The stochastic system (see chapter 4) is markovian, but the Monte-Carlo system is not. If we reset often enough then the dissipation is strong enough to do not give us shocks. Thus we conclude that the mean is less dissipative that the expectation. There is research in this field now, for example see the Denis Talay’s research [GKMPPT]. See the appendix C for the Matlab code. 84 CHAPTER 6. NUMERICAL EXPERIMENTS Valor inicial 1 0.8 0.6 0.4 0.2 0 12 10 12 8 10 6 8 6 4 4 2 2 0 0 Figure 6.2: Initial value. Evolucion 1 0.8 0.6 0.4 1 2 0.2 3 4 0 5 0 6 2 7 4 8 6 9 8 10 10 12 11 Figure 6.3: Numerical solution at time 4. Chapter 7 Conclusion Throughout the whole text we have signaled how Markov processes relate to differential equations, starting with the result of Kakutani for harmonic functions and concluding with an operator, the ∞−laplacian, for which the Markov process considered is much more complicated. The markovian paths are like the characteristic curves for the secon order equations. Or, citing Kohn and Serfaty (see [KS1]) ’the first and second-order cases are actually quite similar.’ These probabilistic methods offer a new intuition useful in many problems (Feynman-Kac, the Kolmogorov-Petrovski-Piskounov equation, NavierStokes equations...). Moreover it gives us a new way to obtain well-known results (like the mean value property, Harnack inequality, existence of classical local solution to the 3D Navier-Stokes equations). We can also generalize this technique to non-local operator (see [A], [NT]), which are the generators of more general Markov processes, the Lévy processes.1 Other equations susceptible to apply these methods are the wave equation, the beam equation... (see [DMT]). From the point of view of the numerical analysis these methods are useful in problems with a very difficult geometry, or high dimensions. In these cases a Monte-Carlo method does not have any drawback. We can also use it to divide the domain in subdomains to apply another method (like the finite elements method...). These ideas are useful in many applications (silhouette recognition, fluid mechanics, quantum mechanics...) because, for example, they gives us a third formulation of quantum mechanics, the Feynman’s formulation. This formulation has the serious advantage of be easily generalized to quantum fields. These methods are well adapted to Dirichlet boundary conditions. However, the Neumann boundary conditions can also be studied with these meth1 These processes are like ’brownian motions and jumps’, and these jumps are the reason for which they are non-local. 85 86 CHAPTER 7. CONCLUSION ods if we consider Itô diffusions reflected in the boundary (see [F], [R]). Appendix A Some useful results A.1 A construction for the brownian motion As we can see in the Itô formula (see below), to define the SDE we need the brownian motion’s increments. Tipically the increments are related to the concept of derivative. The brownian motion is not smooth, but we want to conserve this idea. We start with the white noise, formally dW dt . We define the brownian motion with a time integral. The white noise is in L2 (0, 1), where we choose this time interval to fix ideas, but we can do in general. So, if ψn ara a orthonormal basis of L2 (0, 1) we can write dW (ω, t) = ∞ X An (ω)ψn (t). n=0 Using the brownian motion properties and the fact that, formally, derivative of W , we expect that dW dt is the An ∼ N (0, 1) and independent random variables. We are going to see this more carefully. We have An (ω) = Z 1 dW (ω, t)ψn (t)dt. 0 If we suppose that they are normals with mean 0 and variance 1 then we have Z 1 ψn ψm dt = 0. 0 = E[An ]E[Am ] = E[An Am ] = 0 A similar conditions holds for the variance, E[A2n ] = 1. 87 88 APPENDIX A. SOME USEFUL RESULTS We define the brownian motion as Z t Z t ∞ X ψn (s)ds. An (ω) dW (ω, s)ds = W (ω, t) = 0 n=0 0 We do not choose a basis yet. The previous formula is correct with all basis. Let hn be the n−th function for the Haar basis, i.e. h0 = 11(0,1) , h1 = 11(0,1/2) − 11(1/2,1) hn = 2k 1((n−2k )/2k ,(n−2k +1/2)/2k − 2k 1((n−2k +1/2)/2k ,(n−2k +1)/2k where k is in a such way that 2k ≤ n < 2k+1 . In this basis all is simpler, because of Z t hn (s)ds = sn (t) 0 where sn is the n−th Schauder function (also a basis). We show in this way the result Theorem 30. Let An ∼ N (0, 1) independents random variables. Then W (ω, t) = ∞ X Ak (ω)sk (t) k=0 converges uniformly in t almost everywhere in ω. Moreover W defined in such a way is a brownian motion. See [Ev] for the proof. It is a calculation with characteristic functions to see that the increments are as we expect. When we have the onedimensional brownian motion in (0,1) we extend the definition to higher dimensions putting together some onedimensional brownia motions. TO extend to longer times we do the same. A.2 The Kolmogorov’s regularity theorem Theorem 31 (Kolmogorov). Let X be a stochastic process with continuous paths almost everywhere such that E[|X(t) − X(s)|β ] ≤ C(t − s)1+α , ∀t, s ≥ 0 then for all 0 < γ < α β and T > 0 there exists K(ω) in such a way |X(t) − X(s)| ≤ K|t − s|γ . A.2. THE KOLMOGOROV’S REGULARITY THEOREM 89 Proof. We take T = 1 without loss of generality. We take γ in the considered interval. We define for n ≥ 1 i + 1 i 1 n An = X − X n ≥ nγ for some integer i < 2 . 2n 2 2 Es decir, los conjuntos de sucesos tales que, en la partición que consideramos, tengan incrementos grandes. La idea ahora es acotarlos y aplicar BorelCantelli. n −1 2X i 1 i+1 ≥ − X P X P (An ) ≤ 2n 2n 2nγ i=0 n −1 β 2X i+1 1 −β i E X ≤ −X n 2n 2 2nγ i=0 n −1 2X 1 1+α 1 −β ≤ C 2n 2nγ i=0 P Thus, if we sum over n we have that the series ∞ n=1 (An ) is a convergent one and using Borel-Cantelli’s lemma we conclude that almost everywhere ω there exists m such that n ≥ m X i + 1 − X i ≤ 1 . n n 2 2 2nγ But, taking a constant K(ω) we can have the same for all n. We have to choose the constant considering the first m terms. X i + 1 − X i ≤ K , ∀n ≥ 0 (A.1) 2n 2n 2nγ We have to see that the previous equation gives us what we want. We fix ω ∈ Ω such that the previous equation holds. Let t1 and t2 be two dyadic rational numbers1 such that 0 < t2 − t1 < 1. Let n be an integer in such a way that 2−n ≤ t2 − t1 ≤ 2−n+1 . We can write because they are dyadic, t1 = 1 1 1 i − p1 − p2 ... − p , n < p1 < ... < pk n 2 2 2 2k 1 1 j + q1 + q2 ... + n 2 2 2 for certains i, j with i t1 ≤ n ≤ 2 t2 = 1 1 , n < q1 < ... < qk 2qk j ≤ t2 . 2n The rational numbers with a power of two as denominator. 90 APPENDIX A. SOME USEFUL RESULTS Then, recalling what condition holds for the diference t2 − t1 we have that 1 j−i ≤ t2 − t1 < n−1 . 2n 2 Concluding that j = i or j = i + 1. We can use the equation (A.1) with the previously fixed ω and we obtain j − i γ i j X − X n ≤ K n ≤ K(t2 − t1 )γ 2n 2 2 γ X i − 1 − 1 ... − 1 − X i − 1 ... − 1 ≤ K 1 . 2pr 2n 2p1 2p2 2pr 2n 2p1 2pr−1 we can bound, we have to sum, to substract and to apply the triangle inequality, k X 1 X t1 − X i ≤ K (A.2) n p 2 2 rγ r=1 We have pr > n so 1 1 1 1 1 = n pr −n ≤ n r . p r 2 2 2 2 2 In the previous calculation we used the condition n < p1 ... < pr . In addition we can sum all terms in the series obtaining K k ∞ X 1 K X 1 C ≤ ≤ . p γ nγ rγ 2r 2 2 2nγ r=1 r=1 In the las inequality we used that the series converges, because from r onwards the exponent is rγ > 1. Using the properties of n we conclude that C ≤ C(t2 − t1 )γ . 2nγ In a similar way we obtain a bound for |X(t2 ) − X(j/2n )|. Now |X(t1 ) − X(t2 )| = |X(t1 ) − X(i/2n ) + X(i/2n ) − X(j/2n ) + X(j/2n ) − X(t2 )| ≤ C1 (ω)|t2 − t1 |γ for all dyadic rational numbers in [0,1]. We know that the process has continuous paths ans so we conclude the same for all t ∈ [0, 1]. 91 A.3. THE ITÔ FORMULA A.3 The Itô formula We consider the operator Au = d d X ∂u 1 X ∂2u bi (x) ai,j (x) + 2 ∂xi ∂xj ∂xi i=1 i,j=1 where ai,j = (σσ t )i,j . We have the following well-known theorem. Theorem 32 (Itô formula). Let ~ = ~b(X(t), ~ ~ dX t)dt + σdW i.e., ~ t)dt + dX i = bi (X, X ~ t)dW j σ i,j (X, j with bi ∈ L1 ([0, T ]) and σ i,j ∈ L2 ([0, T ]),2 1 ≤ i, j ≤ d. Let u : Rd × [0, T ] 7→ R be a given continuous function with two spatial derivatives and one derivative in time. Then Y (t) = u(X 1 , ..., X d , t) have the differential d d X ∂u 1 X ∂2u ∂u i (X(t), t)dt + (X(t), t)dX + (X(t), t)dX i dX j dY = ∂t ∂xi 2 ∂xi ∂xj i=1 i,j=1 (A.3) where the term with dX i have the following ’rules’ dt2 = 0, dtdW j = 0, dW i dW j = δi,j dt. We can write it in the integral formulation ~ ~ u(X(t), t) − u(X(0), 0) = Z t 0 Z t ∂u ~. ∇u · σdW + Au ds + ∂t 0 We are going to prove the onedimensional version. 2 These spaces contain the function with – »Z T |f |p < ∞ E 0 and p = 1, 2. (A.4) 92 APPENDIX A. SOME USEFUL RESULTS Theorem 33 (Itô formula (1D)). Let dX = b(X(t), t)dt + σ(X(t), t)dW with b ∈ L1 ([0, T ]) and σ ∈ L2 ([0, T ]) and let u : R × [0, T ] 7→ R be a given continuous function with ∂u ∂u ∂ 2 u , , ∂t ∂x ∂x2 continuous. Then Y (t) = u(X(t), t) has the differential dY ∂u 1 ∂2u ∂u (X(t), t)dt + (X(t), t)dX + (X(t), t)σ 2 dt 2 ∂t ∂x 2 ∂x ∂u ∂u 1 ∂2u 2 = (X(t), t) + (X(t), t)b(X(t), t) + (X(t), t)σ dt ∂t ∂x 2 ∂x2 ∂u (X(t), t)σ(X(t))dW. + ∂x = Proof. We start with the case u = xm . We claim that 1 d(X m )(t) = mX(t)m−1 dX + m(m − 1)X m−2 (t)σ 2 dt. 2 The m = 0, 1 cases are trivial. We suppose that it becames tru for m − 1 and we are going to show it for the m case. We will use the following lemma (see [Ev] for the proof): Lemma 11 (Product rule). If we have dX1 (t) = b1 (X(t)1 , t)dt + σ1 (X(t)1 , t)dW dX2 (t) = b2 (X(t)2 , t)dt + σ2 (X2 (t), t)dW with bi ∈ L1 ([0, T ]) and σi ∈ L2 (0, T ). Then the derivative of the product is d(X1 X2 )(t) = X2 (t)dX1 (t) + X1 (t)dX2 (t) + σ1 (X(t)1 , t)σ2 (X(t)2 , t)dt. We write d(X m−1 X) using the lemma and we conclude that the Itô formula holds for this kind of functions. Because of the linearity we conclude that the Itô formula holds for all the polynomials in x. We consider the product of a polynomial in x and another in t, i.e. u(t, x) = f (x)g(t). Then d(f (X(t))g(t)) = f (X(t))g′ dt + gdf (X(t)) 1 = f (X(t))g′ dt + g[f ′ (X(t))dX + f ′′ (X(t))σ 2 dt]. 2 93 A.3. THE ITÔ FORMULA This is the expression we expect. We conclude that the formula holds for all u(x, t) = m X fi (x)gi (t). (A.5) i=1 We use a density argument. Let u be a function as we state above, then there exists un as in (A.5) approximating uniformly in compacts in R×[0, T ] to u and to the mentioned derivatives of u. We conclude the proof taking the limits. We have a relationship between the (1.10) and the elliptic operator Au = d d X ∂u 1 X ∂2u bi (x) ai,j (x) + 2 ∂xi ∂xj ∂xi i=1 i,j=1 where ai,j = (σσ t )i,j . Another useful version is Theorem 34 (Itô formula (stopping times)). Let ~ = ~b(X(t), ~ ~ dX t)dt + σdW i.e., ~ t)dt + dX i = bi (X, X ~ t)dW j σ i,j (X, j with bi ∈ L1 ([0, T ]) and σ i,j ∈ L2 ([0, T ]), 1 ≤ i, j ≤ d. Let u : Rd × [0, T ] 7→ R a given continuous function with two spatial derivatives and one time-derivative. Given τ a stopping time. Then Y (t) = u(X 1 , ..., X d , t) has the derivative dY = d d X ∂u 1 X ∂2u ∂u (X(τ ), τ )dt+ (X(τ ), τ )dX i + (X(τ ), τ )dX i dX j ∂t ∂xi 2 ∂xi ∂xj i=1 i,j=1 (A.6) where the terms with dX i has the following ’rules’ dt2 = 0, dtdW j = 0, dW i dW j = δi,j dt. We can also write in the integral formulation Z τ Z τ ∂u ~ ~ ~. u(X(τ ), τ ) − u(X(0), 0) = ∇u · σdW + Au ds + ∂t 0 0 (A.7) 94 APPENDIX A. SOME USEFUL RESULTS The ’rules’ in the main result can be shown using the joint quadratic variation. Definition 11 (Quadratic variation). Let X(t) be a stochastic process defined for a < t < b and let P = {a ≤ t0 ..., tn ≤ b} be a partition of this interval. We define the quadratic variation related to the given partition as < X(t) >P = k−1 X (X(ti+1 ) − X(ti ))2 + (X(t) − X(tk ))2 i=0 where tk is in a such way that tk < t < tk+1 . If when we refine the partition there exists a limit < X(t) > (in probability) and this limit is independent of the considered partitions then we call this limit the quadratic variation of X(t). If we have a bounded variation and continuous process then its quadratic variation vanishes. So, the quadratic variation of a smooth function vanishes. Definition 12 (Joint quadratic variation). Let M (t), X(t) be two stochastic processes defined for a < t < b and let P = {a ≤ t0 ..., tn ≤ b} be a partition of this interval. We define < X(t), M (t) >P = k−1 X (X(ti+1 )−X(ti ))(M (ti+1 )−M (ti ))+(X(t)−X(tk ))(M (t)−M (tk )) i=0 where tk is in a such way that tk < t < tk+1 . If when we refine the partition there exists a limit < X(t), M (t) > (in probability) and this limit is independent of the considered partitions then we call this limit the joint quadratic variation of X(t) and M (t). The quadratic variation (if there exists) is a continuous process of bounded variation. In addition it is bilinear, symmetric, positive defined and an kind of Schwarz inequality holds | < X(t), M (t) > − < X(s), M (s) > | ≤ p p < X(t) > − < X(s) > < M (t) > − < M (s) >. To see where this process are well-defined, other properties and the Itô integral (or the Stratonovich integral) considering these processes see [Ku]. In chapter 4 we need a generalized version of Itô formula. See [Ku2] for the complete proof. But before this statement we need some definitions: Definition 13 (Local martingale). A stochastic process X(t) adapted3 to a given filtration, Ft , is a local martingale if there exists increasing stopping times, τn , such that X(min{t, τn }) is a martingale. 3 i.e., X(t) is Ft measurable for all t. 95 A.4. EXISTENCE AND UNIQUENESS FOR PDE Definition 14 (Semimartingale). A stochastic process X(t) is a semimartingale if it is sum of a bounded variation process and a local martingale. Theorem 35 (Itô formula (generalized)). Let F~ (x, t) be a process C 2 (in x) and a semimartingale C 1 (in t). Let ~g (t) be a continuous semimartingale such that x and ~g (t) takes values into D ⊂ Rd . Then F~ (~g (t), t) is a continuous semimartingale and satisfies F (~g (t), t) − F (~g (0), 0) = Z d Z X t F (~g (s), ds) + 0 i=1 t 0 ∂F (~g (s), s)dgi (s) ∂xi + d 1 X ∂2F (~g (s), s) < dgi (s), dgj (s) > 2 ∂xi ∂xj + d Z t X i,j=1 i=1 0 ∂F i (~g (s), ds), g (t) . ∂xi An idea to prove it is to see that, with a partition given, we have X F (gt , t) − F (g0 , 0) = F (gtk+1 , tk+1 ) − F (gtk , tk ) = and X X F (gtk+1 , tk+1 ) ± F (gtk , tk+1 ) − F (gtk , tk ) F (gtk , tk+1 ) − F (gtk , tk ) ≈ Z t F (gs , ds). 0 For the other terms we do in a similar way. A.4 Existence and uniqueness for PDE Sometimes we use an existence and uniqueness of classical solution result for certain PDE. Theorem 36. Let à be the elliptic operator defined previously (see chapters 2, 3) with c ≥ 0 and Hölder-α and bounded coefficients. We consider a bounded domain U satisfying the inner sphere property for all point on the boundary. Let f be a bounded and Hölder-α function, and let g be a continuous function. Then the problem Ãu = f if x ∈ U, u|∂U = g has an unique classical solution u ∈ C(Ū ) ∩ C 2,α (U ). See [GT] for the proof. Appendix B Itô integral We have to make sense to the expression Z t GdW 0 where G is a stochastic process and dW dt is the standard white noise. Like before we have a fixed probability space and a filtration1 adapted to the given brownian motion. Definition 15. Given [0, T ] a time interval. We define a partition P as a sequence of times satisfying 0 = t0 < t1 < ... < tn = T. We define the size of the partition as |P | = max |tk+1 − tk |. 0≤k≤n−1 To define the Itô integral we do as follows, first we consider step processes and we approximate more general process with the step processes. We consider the space 2 L (0, T ) = {G, G progressively measurable and such that E 1 Sequence of σ−algebras satisfying the following conditions F(t) ⊂ F(s) si t < s σ(W (t)) ⊂ F(t) ∀t ≥ 0 F(t) independent of σ(W (s) − W (t), ∀s ≥ t) where σ(W (s) − W (t), ∀s ≥ t) is the future of the brownian motion. 96 Z T 2 G dt 0 < ∞}. 97 Definition 16. We define a step process as a process G L2 (0, T ) for which P∈ n there exists time intervals (tk , tk+1 ) such that G(t) = k=1 Gk 1(tk ,tk+1 ) with Gk random variables F(tk )−measurables. To these processes we define the Itô stochastic integral as Z T GdW = 0 n−1 X k=0 Gk (W (tk+1 ) − W (tk )). (B.1) We recall that it is a random variable. From the definition we obtain that this operator is linear. Using the linearity we have E Z T GdW 0 =E n−1 X k=0 Gk (W (tk+1 ) − W (tk )) = 0 (B.2) using the brownian motion properties and the independence hypothesis in the definition of filtration. We know that the brownian motion has bounded quadratic variation, and so 2 Z T Z T 2 G dt . (B.3) GdW =E E 0 0 Indeed, E Z T GdW 0 2 = n−1 X k,j=1 E[Gk (W (tk+1 ) − W (tk ))Gj (W (tj+1 ) − W (tj ))] Si suponemos ahora que j 6= k entonces, aplicando la independencia, tenemos que esos términos se anulan, por lo que sólo quedan los términos j = k. Utilizamos ahora que E[W (tk+1 ) − W (tk )] = tk+1 − tk . We have then the result Theorem 37. The following properties hold for the Itô stochastic integral of a step process: 1. Is linear. 2. We have that E Z T GdW 0 E Z 0 T GdW 2 =E =0 Z 0 T 2 G dt . 98 APPENDIX B. ITÔ INTEGRAL Now, given a general stochastic process in L2 (0, T ), we approximate it with a sequence of step processes. This sequence of random variables will be Cauchy in L2 (0, T ) (moreover, their limit will be G). So, in L2 (Ω) we have Z Z 2 T E 0 T Gm − Gn dW =E 0 (Gm − Gn )2 dt → 0 using the second property in the previous theorem. We can define the integral of the limit as the limit of the integrals in the L2 (Ω) sense. See [Du], [Ev] for more details. The stochastic integral of a L2 (0, T ) process satisfy the same properties that for the step processes. For the indefinite integral we have the following result. This result is needed in the first chapter. Theorem 38. Let I(t) = Z t GdW 0 then I(t) is a martingale. Moreover, it has continuous paths almost surely. See [Ev] for a proof of the second part. There is another tricky detail. We evaluated the Riemann sums in the left hand side point of the interval, tk . If we evaluate the Riemann sum in other point our result will change. The use of the left hand side point of the interval is the Itô approach. If we use the medium point of the interval, then we have the Stratonovich integral. Their properties are very different, for example, in the Itó approach we can not apply the usual chain rule to derive. In the Stratonovich sense we can. However, to use the Itô formula gives us the results in this text. Other advantage of the Itô approach is that the indefinite integral is a martingale. There are a useful conversion formulas to change between them (see [Ev]). We are going to study the following example: Z T 0 1 1 W dW = W 2 (t) − (λ − )T 2 2 where λ is in a such way that we have the Riemann sum mX n −1 k=0 W (τk )(W (tnk+1 ) − W (tnk )) with τk = (1 − λ)tk + λtk+1 . Thus λ=0 99 is the Itô integral, and λ = 1/2 is the Stratonovich integral. We observe the diference between the results. We want to recall that all integration is in the L2 (Ω) sense, this is E mX n −1 k=0 if n → ∞. 2 1 1 W (τk )(W (tk+1 ) − W (tk )) − W 2 (t) − (λ − )T →0 2 2 Appendix C Matlab code C.1 Brownian motion paths function [B,t]=browniano(N,M,T) %This function approach M trajectories of the brownian motion %in then [0,T] interval with step T/N t=0:T/N:T; B=zeros(N+1,M); for j=1:M for i=1:N B(i+1,j)=B(i,j)+normrnd(0,sqrt(1/N)); end end C.2 Brownian bridge paths function [B,P,t1]=puentebrowniano(N,M) %This function approach M trajectories of the brownian bridge %in then [0,1] interval with step 1/N t1=0:1/N:1; B=zeros(N+1,M); P=B; for j=1:M for i=1:N B(i+1,j)=B(i,j)+normrnd(0,sqrt(1/N)); %the brownian motion % incrementsare normals with zero mean and standard %deviation (1/N)^(1/2) P(i,j)=B(i,j)-t1(i)*B(end,j);%we know P(t)=B(t)-t(B(T)) end end 100 C.3. EULER METHOD FOR A SDE C.3 C.3.1 101 Euler method for a SDE 1D case function [t,Y]=sde1(a,b,Y0,T,N) %this function approachs the solution %of the SDE %dY=a(Y,t)dt+b(Y,t)dW %using the Euler method, i.e. %Y(n+1)=a(Y(n))T/N+b(Y(N))(W(t(n+1))-W(t(n)) %a, b are the functions. Y0 is the initial value %T is the final time and N is the number of nodes t=0:T/N:T; Y=zeros(1,N+1); Y(1)=Y0; for i=1:N Y(i+1)=Y(i)+feval(a,Y(i),t(i))*T/N+feval(b,Y(i),t(i))... ...*normrnd(0,sqrt(1/N)); end C.3.2 2D case function [t,Y]=sde2(a,b,Y0,T,dt) %this function approachs the solution %of the SDE %dY=a(Y,t)dt+b(Y,t)dW %using the Euler method, i.e. %Y(n+1)=a(Y(n))T/N+b(Y(N))(W(t(n+1))-W(t(n)) %a, b are the functions. Y0 is the initial value %T is the final time and N is the number of nodes t=0:T/dt:T; Y=zeros(2,dt+1);%two dimensional diffusion Y(:,1)=Y0; for i=1:dt Y(:,i+1)=Y(:,i)+feval(a,Y(:,i),t(i)).*T/dt+feval(b,Y(:,i),t(i))... .*[normrnd(0,sqrt(1/dt)),normrnd(0,sqrt(1/dt))]’; end C.4 Monte-Carlo method for the laplacian function u=lapbrow(g,n,N,B) %this function solves the 2d laplacian %with a given boundary value (g) 102 APPENDIX C. MATLAB CODE %in [-1,1]^2. %n^2 is the number of grid points. %for each point (xi,yj) we calculate B %brownian paths, and their g values and we take the mean. %1/N is the time step x=-1:1/(n-1):1; y=x; [X,Y]=meshgrid(x,y); u=zeros(2*n-1,2*n-1);%the rows are y and the columns are x for a=1:2*n-1 u(1,a)=feval(g,[x(a),1]);%y=1 end for b=1:2*n-1 u(b,1)=feval(g,[-1,y(b)]);% x=-1 end for c=1:2*n-1 u(2*n-1,c)=feval(g,[x(c),-1]);% y=-1 end for d=1:2*n-1 u(d,2*n-1)=feval(g,[1,y(d)]);% x=1 end M=zeros(B,2); for i=2:2*n-2 for j=2:2*n-2 tau=[x(i),y(j)];%the brownian starts in a grid point tau1=[0 0]; for k=1:B while prod(abs(tau1)<[1,1])%this holds if %we are in the square tau1=tau+[normrnd(0,sqrt(1/N)),normrnd(0,sqrt(1/N))]; %a new point, the outer. if prod(abs(tau1)<[1,1]) tau=tau1;%this saves the interior point end end M(k,:)=tau1; C.5. SILHOUETTE RECOGNITION 103 tau1=[0 0]; tau=[x(i),y(j)]; end for l=1:B G(l)=feval(g,M(l,:));%we evaluate g in the l-th row end G; u(j,i)=mean(G); clear G; end end mesh(X,Y,u);title(’Solution’) C.5 Silhouette recognition This program uses a SOR method to solve the Poisson equation. The domain (the silhouette) is introduced as a .png image, changing the colors and with an ’if’ condition. function [img,img2,u,t,cnt]=imagessor(tol,itmax,image) %This program uses a SOR %method to solve the Poisson equation %tol is the tolerance %itmax is the maximum number of iterations %image is a .png image tic img=imread(image); figure;imagesc(img); input(’Press any key’) img=double(img); [H,W]=size(img) w= 2 / ( 1 + sin(pi/(H+1)) );%our SOR parameter for i=1:H for j=1:W img2(i,j)=abs(img(i,j)-255); %change the colors between them end end img2; figure; imagesc(img2); input(’Press any key’) clear i,j; 104 APPENDIX C. MATLAB CODE %We start the algorithm u=img2; v=u; err=1; cnt=0; while((err>tol)&(cnt<=itmax)) for i=2:H-1 for j=2:W-1 if (img2(i,j)==0) else v(i,j)=u(i,j)+w*(v(i-1,j) + u(i+1,j) + v(i,j-1)... ... + u(i,j+1) +1 - 4*u(i,j))/4; E(i,j)=v(i,j)-u(i,j); end end end err=norm(E,inf); cnt=cnt+1; u=v; end u=flipud(u); figure;imagesc(u); mesh(u) t=toc; The programs to calculate norms, gradients, Φ and Ψ are the followings function [Gux,Guy,NGu,t]=gradient(u) %this program calculates the gradient and its norm %Gux is the first component, %Guy is the second component %NGu is the gradient norm tic [H,W]=size(u); for i=2:H for j=2:W Gux(i,j)=u(i,j)-u(i-1,j); Guy(i,j)=u(i,j)-u(i,j-1); NGu(i,j)=(Gux(i,j)^2+Guy(i,j)^2)^0.5; end end t=toc; C.6. MONTE-CARLO METHOD FOR PARABOLIC EQUATIONS 105 function [Phi,t]=phi(u,NGu) %This progam calculates phi=u+NGu^2 %NGu is the gradient of u norm tic [H,W]=size(NGu); for i=1:H for j=1:W Phi(i,j)=u(i,j)+NGu(i,j)^2; end end t=toc; function [Psi,t]=psiimages(u,Gux,Guy,NGu) %This program calculates psi=-div(gradient(u)/norm(gradient(u)) %NGu is the gradient of u norm %Gux is the first component, %Guy is the second component tic [H,W]=size(NGu); for i=2:H for j=2:W Psix(i,j)=((Gux(i,j)-Gux(i-1,j))*NGu(i,j)-Gux(i,j)... ...*(NGu(i,j)-NGu(i-1,j)))/NGu(i,j)^2; Psiy(i,j)=((Guy(i,j)-Guy(i,j-1))*NGu(i,j)-Guy(i,j)... ...*(NGu(i,j)-NGu(i,j-1)))/NGu(i,j)^2; Psi(i,j)=-Psix(i,j)-Psiy(i,j); end end t=toc; C.6 Monte-Carlo method for parabolic equations function [t,u]=parabolic(T,M,N,a,b,dx,u0) %Code to simulate M diffusions, with N time-grid nodes %the diffusions start at x0. T is the final time (integer), %a is the trasport, a function. %b is the diffusion. %u=u(T,x). %dx is the spatial step %u0 is the initial datum function 106 APPENDIX C. MATLAB CODE %%%% Time tic %%%% Domain x=0:dx:10; y=x; Nx=length(x); Ny=length(y); %%%% Initial datum for i=1:Nx for j=1:Ny uo(i,j)=feval(u0,[x(i),y(j)]); end end clear i,j; figure(1) mesh(uo);title(’Initial value’) %%%%Time code for l=1:T %%%% Code to simulate the diffusions for i=1:Nx %x step for j=1:Ny %y step %%%It simulate M diffusions started at [x(i),y(j)] % We canculate u0 at these M point % and save them as a vector. for k=1:M [t,Y]=sde2(a,b,[x(i),y(j)]’,T,N); uu(k)=feval(u0,Y(:,end)); end %%%% Expectation code uoo(i,j)=mean(uu); %u(l,i,j)=u(l,x(i),y(j)) u(l,i,j)=uoo(i,j); clear uu; end end figure mesh(uoo);title(’Evolution’) end C.7. CODE TO APPROXIMATE THE ∞−LAPLACIAN 107 t=toc; C.7 Code to approximate the ∞−laplacian function [u,cnt,t]=infinitylaplacian(B,f,tol,itmax,N) %This program approximates the infinity laplacian %with f as right hand side term %and boundary values B on the boundary of the square %[-1/2,1/2]^2 %tol is the tolerance %itmax is the maximum number of iterations %(N+1)^2 is the number of grid points %B and f are matrices tic; u=B; u1=u; umax=u; umin=u; v=u; %The discretization of -inflap(u)=f is %2u_ij-sup_{i,j vecinos} u -inf u=f_ij %then u_ij=(f_ij+sup u + inf u)/2 err=1; cnt=0; while((err>tol)&(cnt<=itmax)) for j=2:N for i=2:N %we calculate the supremums and the infimums umax(i,j)=max([v(i-1,j) v(i,j-1) u(i+1,j) u(i,j+1)]); umin(i,j)=min([v(i-1,j) v(i,j-1) u(i+1,j) u(i,j+1)]); u1(i,j)=umax(i,j)+umin(i,j);% asi el inflap es 2u-u1 v(i,j)=(f(i,j)+u1(i,j))/2; E(i,j)=v(i,j)-u(i,j); end end err=norm(E,inf); cnt=cnt+1; u=v; end u=flipud(u’); t=toc; Index ∞−laplacian, 71, 73 σ−algebra, 22 p-laplacian, 79 p−laplacian, 71 ’tug of war with noise’, 79 ’tug of war’, 73 1-laplacian, 71, 80 de Broglie, 50 decision tree, 37 differential game, 71 diffusion, 19, 23, 35, 43 Dirichlet boundary conditions, 29, 46 domain, 24 dynamic programming principle, 75, 76, 78, 81 absolutely minimizing Lipschitz extension, 73 elliptic equations, 29 action, 50 elliptic operator, 26, 37 advection, 45 Euler equations, 56 existence and uniqueness (PDE), 95 Banach’s fixed point theorem, 70 existence and uniqueness (SDE), 15 Bismut-Elworthy-Li formula, 28 expectation, 22 Borel sets, 20 exponential, 46 Borel-Cantelli Lemma, 16 branching brownian motion, 46 Feller semigroup, 27 brownian bridge, 23, 42, 53 Feynman masure, 53 brownian motion, 7, 8, 11, 13, 35, 42, Feynman path integral, 48 87 Feynman’s formulation, 48 Feynman-Kac formula, 33, 43, 53 Cauchy problem, 41 Fisher equation, 46 central limit theorem, 8 flow, 18 chain rule, 55 functional, 50 Chapman-Kolmogorov equation, 49 functional integration, 48 Chapman-Kolmogorov equations, 21 fundamental solution, 28, 46 characteristic curves, 5, 60 characteristic form, 38 generator, 25, 26, 28, 30, 43 Chevichev inequality, 16 Green operator, 39 classical mechanics, 49 Gronwall inequality, 15 classical solution, 32, 38 concavity, 37 Hölder, 11, 18 contraction semigroup, 24 Haar basis, 88 convexity, 37 hamiltonian, 38, 54 critical point, 50 harmonic function, 32 cylinders, 20 heat equation, 8, 26, 58 108 109 INDEX history, 73 Hopf-Cole transformation, 57 images, 37 inviscid Burgers equation, 60 inviscid Burgers equations, 57 irrotational flow, 59 Itô diffusion, 19, 28 Itô equations, 19 Itô formula, 19, 27, 30, 31, 33, 34, 44, 91 Itô integral, 13, 30, 96 operators, 23 ordinary equation, 18 Ornstein-Uhlenbeck equation, 12 parabolic equation, 8, 42 parabolic equations, 41 path, 10 path integral, 22 path space, 20 PDE, 30 phase space paths, 54 pinned Wiener measure, 20, 50 Planck constant, 50 joint quadratic variation, 94 Poisson equation, 35 population dynamics, 47 kernel, 41, 52, 54 Kolmogorov extension theorem, 20, potential, 33, 38, 49 progressively measurable, 14 21 Kolmogorov regularity theorem, 10, quadratic variation, 11, 94 18 quantum fields, 54 Kolmogorov’s regularity theorem, 89 quantum mechanics, 48, 49 Kolmogorov-Petrovskii-Piskunov equation, 46 random characteristic curves, 60 random walks, 35 Lévy processes, 28 Regularity for SDE, 18 lagrangian, 50, 54 regularization, 54, 55 Langevin equation, 12, 19 relativistic effects, 49 laplacian, 30 relativistic particles, 54 least action principle, 49 representation formula, 30 lema de Borel-Cantelli, 89 Riemann integral, 13 Lipschitz, 28, 33, 70 local martingale, 94 Schauder basis, 88 Schrödinger equation, 33, 51 Markov, 12, 19, 23 SDE, 13, 61, 75 martingale, 94 semigroup, 24, 43 matrix, 38 semimartingale, 95 mean value property, 32 sequence, 98 minimizer, 50 silhouette, 35 singularities, 60 Navier-Stokes equations, 56 space-time paths, 54 Neumann boundary conditions, 29 Spencer game, 80 non-local operator, 54 spin, 49 numerical methods, 82 state of a system, 49 step process, 97 ODE, 25, 61 stochastic equation, 18, 26 operador de Weber, 68 110 stochastic flow, 18, 41, 42 stochastic game, 75 stochastic integrals, 55 stochastic processes, 13 stopping time, 29, 38, 93 stopping times, 41 strategy, 73 Stratonovich equations, 19 Stratonovich integral, 13, 19, 98 sucessive approximation, 15 supermartingala, 78 thresholding, 37 total variation, 11 transition density, 24, 40, 41 transition function, 24 tug of war, 71, 76 unbounded domains, 39 viscid Burgers equation, 57, 58, 60, 64 viscid Burgers equations, 57 wave-particle duality, 50 white noise, 12, 87 Wiener measure, 9, 19, 23, 41, 53 INDEX Bibliography [A] P.Amore, ’Alternative representation for non-local operators and path integrals’, arXiv:hep-th/0701032v3. [ACJ] G.Aronsson,M.Crandall, P.Juutinen, ’A tour on the theory of absolutely minimizing functions’, Bull. Amer. Math. Soc., 41 (2004), 439-505. [App] D.Applebaum, Lévy Processes and Stochastic Calculus, Cambridge Studies in Advanced Mathematics, 2004. [BEJ] E.Barron,L.Evans,R.Jensen, ’Infinity laplacian, Aronsson’s equation and their generalizations’, Trans.Amer.Math.Soc. 360 (2008), 77-101 [BF] Y.N. Blagoveshcenskii y M.I. Freidlin, ’Some properties of diffusion processes depending on a parameter’, Dokl. Akad. Nauk. SSSR, 138 (1961), 508-511. [C] P.Constantin, ’An Eulerian-Lagrangian approach for incompressible fluids: local theory’. J.Amer.Math.Soc. 14 (2001), no.2, 263-278. [CR] K.L. Chung y A.M. Rao, ’Feynman-Kac functional and the Schrödinger equation’. Seminar on Stochastic processes Birkhäuser, 1981. [ChGR] F.Charro, J.Garcı́a y J.D.Rossi, ’A mixed problem for the infinity laplacian via tug-of-war games’. Calc.Var.Partial Differential Equations 34 (2009), 307-320. [CI] P.Constantin y G.Iyer, ’A stochastic lagrangian representation of the 3-dimensional incompressible Navier-Stokes equations’, aparecerá en Communications on Pure and Applied Mathematics. [DMT] R. Dalang, C.Mueller, R. Tribe, ’A Feynman-Kac-type formula for the deterministic and stochastic wave equations and other p.d.e.’s’, arXiv:0710.2861v1 [math.PR]. 111 112 BIBLIOGRAPHY [Du] R.Durret, Stochastic calculus: Press, 1996. a practical introduction, CRC [Dy] E.B.Dynkin, Markov processes, vol I., Springer, 1965. [E] A.Einstein, Investigations on the Theory of the Brownian Movement, ed. por R.Förth, Dover, 1956. [Ev] L.C.Evans, Stochastic Differential http://math.berkeley.edu/evans. [Ev2] L.C.Evans, Partial Diferential Equations, AMS, 2008. [Ev3] L.C.Evans, ’The 1-laplacian, the ∞−laplacian and differential games’, Contemp. Math., 446 (2007), 245-254. [Fe] R.P. Feynman, ’Space-time approach to non-relativistic quantum mechanics’, Rev. of Mod. Phys, 20 (1948), 367-387. [Fe2] R.P. Feynman, The principle of least action in quantum mechanics, ed por Laurie M. Brown, World scientific, 2005. [FH] R.P. Feynman,A.R. Hibbs, Quantum mechanics and Path Integrals, McGraw-Hill, 1965. [F] M. Freidlin, Fuctional integration and partial differential equations, Princeton university press, 1985. [Fr] A. Friedman, Stochastic differential equations and applications, vol.1, Academic press, 1975. [GJ] J.Glimm, A.Jaffe Quantum physics, a functional integral point of view second edition, Springer-Verlag, 1987. Equations, [GKMPPT] C.Graham, T.Kurtz, S.Meleard, P.Protter, M.Pulvirenti, D.Talay, Probabilistic Models for Nonlinear Partial Differential Equations, Lecture Notes in Mathematics 1627, Springer-Verlag, 1996. [GGSBB] L.Gorelick, M.Galun, E.Sharon, R.Basri, A.Brandt, ’Shape Representation and Classification Using the Poisson Equation’, IEEE transaction on pattern analysis and machine intelligence, 28 (2006), no.12, 1991-2004. [GT] D.Gilbarg, N.S.Trudinger, Elliptic partial differential equations of second order, Springer-Verlag, 1970. [I] S. Itô, Diffusion equations, American Mathematical Society, 1992. BIBLIOGRAPHY 113 [Iy] G.Iyer ’A stochastic Lagrangian formulation of the incompressible Navier-Stokes and related transport equations’, Ph.D. Thesis, University of Chicago, 2006. [Iy2] G.Iyer, ’A stochastic lagrangian proof of global existence of the Navier-Stokes equations for flows with small Reynolds number’, Ann. Inst. H. Poincaré Anal. Non Linéaire 26 (2009), 181-189. [IN] G.Iyer, A.Novikov, ’The regularizing effects of resetting in a particle system for the Burgers equation’, (Preprint). [K] Kac, M. ’Wiener and integration in function spaces.’, Bull. Amer. Math. Soc. 72 (1966) 53-68. [K2] Kac, M. ’On some connections between probability theory and differential and integral equations.’, Proceedings of the Second Berkeley Symposium on Mathematical Statistics and Probability (1950), 189215. [Kl] J.R. Klauder, ’The Feynman path integral: slice’,arXiv:quant-ph/0303034v1. [Kle] H. Kleinert, Path integrals in quantum mechanics,statistics, polymer physics and financial markets, fourth edition, World Scientific, 2006. [KS1] R.Kohn S.Serfaty, ’A deterministic-control-based approach to motion by curvature’, Comm.Pur.Appl.Math, 59 (2006), 344-407. [KS2] R.Kohn S.Serfaty, ’Second order PDE’s and deterministic games’, (Preprint) [Ku] H.Kunita, Stochastic differential equation and stochastic flows of diffeomorphism, Lecture Notes in Math. vol 1097, Springer, 1984. [Ku2] H.Kunita, Stochastic flows and stochastic differential equations, Cambridge studies in advanced mathematics, 1997. an historical [LSU] O. Ladyzenskaya, V.Solonnikov, N.Uralceva, Linear an quasilinear equations of parabolic type, American Mathematical Society, 1968. [McK] H.P. McKean, ’Application of brownian motion to the equation of Kolmogorov-Petrovskii-Piskunov’, Comm. Pur. Appl. Math., 28 (1975), 323-331. [MP] P.Mörters y Y.Peres, Brownian motion, disponible en versión electrónica en http://people.bath.ac.uk/maspm/book.pdf 114 BIBLIOGRAPHY [NT] Nagasawa, M. and Tanaka, H ’The principle of variation for relativistic quantum particles.’, Séminaire de Probabilités, 35, 1-27 [Ob] A. Oberman, ’A convergent difference scheme for the infinity laplacian: construction of absolutely minimizing Lipschitz extensions.’ Mathematics of computation, 74 (2004), 1217-1230. [O] B.Oksendal, Stochastic differential equations: an introduction with applications, fifth edition, Springer-Verlag, 2000. [PS] Y.Peres, S.Sheffield, ’Tug-of-war with noise: a game-theoretic view of the p-laplacian’, Duke Mathematical Journal, 145, (2008), 91-120. [PSSW] Y.Peres, O.Schramm, S.Sheffield y D.Wilson, ’Tug-of-war and the infinity laplacian’, aparecerá en Jour. Amer. Math. Soc. [R] S. Ramasubramanian, ’Reflecting brownian motion in a Lipschitz domain and a conditional gauge theorem.’ The Indian Journal of Statistics, 63 (2001), 178-193. [S] B.Simon, Functional integration and quantum physics, Academic press, 1979. [Z] J. Zinn-Justin, Path integrals in quantum mechanics, Oxford Graduate Texts, 2005.
© Copyright 2025 Paperzz