Brownian Motion An Introduction to Brownian Motion, Wiener Measure, and Partial Differential Equations Prof. Michael Mascagni Applied and Computational Mathematics Division, Information Technology Laboratory National Institute of Standards and Technology, Gaithersburg, MD 20899-8910 USA AND Department of Computer Science Department of Mathematics Department of Scientific Computing Graduate Program in Molecular Biophysics Florida State University, Tallahassee, FL 32306 USA E-mail: mascagni@fsu.edu or mascagni@math.ethz.ch or mascagni@nist.gov URL: http://www.cs.fsu.edu/∼mascagni Brownian Motion Outline of the Lectures Introduction to Brownian Motion as a Measure Definitions Donsker’s Invariance Principal Properties of Brownian Motion The Feynman-Kac Formula Explicit Representation of Brownian Motion The Karhunen-Loève Expansion Explicit Computation of Wiener Integrals The Schrödinger Equation Proof of the Arcsin Law Advanced Topics Action Asymptotics Brownian Scaling Local Time Donsker-Varadhan Asymptotics Can One Hear the Shape of a Drum? Probabilistic Potential Theory Brownian Motion Introduction to Brownian Motion as a Measure Introduction to Brownian Motion I I I I I def Let Ω = {β ∈ C[0, 1]; β(0) = 0} = C0 [0, 1], be an infinitely dimensional space we consider for placing a probability measure Consider (Ω, B, P), where B is the set of measurable subsets (a σ-algebra) and P is the probability measure on Ω hR i 1 We would like to answer questions like P 0 β 2 (s)ds ≤ α ? We now construct Brownian motion (BM) via some limit ideas Central Limit Theorem (CLT): let X1 , X2 , . . . be independent, identically P distributed( i.i.d.) with E[Xi ] = 0, Var [Xi ] = 1 and define Sn = ni=1 Xi 1. Note if X1∗ , X2∗ , . . . are i.i.d. with E[Xi∗ ] = µ, Var [Xi∗ ] = σ 2 < ∞, then Xi = 2. Then Xi∗ −µ has E[Xi ] = 0, Var [Xi ] = 1 σ Sn √ converges in distribution to N(0, 1) n as n → ∞ Brownian Motion Introduction to Brownian Motion as a Measure Introduction to Brownian Motion I Let X1 , X2 , . . . be as before, then it follows from the CLT that Z α u2 1 Sn lim P √ ≤ α = √ e− 2 du. n→∞ n 2π −∞ I Erdös and Kac proved (we will find the σi (·)’s): q R i h 2 S S Sn 2 α − u2 ≤ α = σ (α) = e 1. limn→∞ P max √1n , √2n , . . . , √ du 1 π 0 n 2 2 S1 +S2 +···+Sn2 2. limn→∞ P ≤ α = σ2 (α) n2 h i S +S +···+Sn ≤ α = σ3 (α) 3. limn→∞ P 1 23/2 n I Let Nn = #{S1 , . . . , Sn |Si > 0}, then if α ≤ 0 0, √ Nn lim P ≤ α = π2 arcsin α, if 0 ≤ α ≤ 1 n→∞ n 1, if α ≥ 1 Brownian Motion Introduction to Brownian Motion as a Measure Definitions Definitions I X1 , X2 , . . . are as above, and ∀n ∈ N and t ∈ [0, 1] define (S √1 , t = 0 (n) χ (t) = Sin i−1 √ , < t ≤ ni , i = 1, 2, . . . , n n n I Let R denote the space of Riemann integrable functions on [0, 1]. I Theorem: F : R → R and with some weak hypotheses, then h i lim P F χ(n) (·) ≤ α = PW [F (β) ≤ α] , n→∞ where PW denotes the probability called “Wiener measure,” and this result is called Donsker’s Invariance Principal Brownian Motion Introduction to Brownian Motion as a Measure Donsker’s Invariance Principal Examples of Donsker’s Invariance Principal R1 β 2 (s) ds, then by the theorem " n # Z 1 X S2 2 i lim P ≤ α = PW β (s)ds ≤ α n→∞ n2 0 1. F [β] = 0 i=1 2. F [β] = β(1), then Z α u2 Sn 1 e− 2 du lim P √ ≤ α = PW [β(1) ≤ α] = √ n→∞ n 2π −∞ ( R 1 1+sgnβ(s) 1, : x >0 3. F [β] = 0 ds, where sgn(x) = Then 2 −1, : x ≤ 0 lim P n→∞ Z 1 1 + sgnβ(s) Nn ≤ α = PW ds ≤ α n 2 0 Brownian Motion Introduction to Brownian Motion as a Measure Donsker’s Invariance Principal Defining Wiener Measure Using Cylinder Sets I For any integer n, any choice of 0 < τ1 < · · · < τn ≤ 1, and any Lebesgue measurable (L-mb) set, E ∈ Rn define the “interval” I = I(n; τ1 ; . . . ; τn ; E) := {β(·) ∈ C0 [0, 1]; (β(τ1 ), . . . , β(τn )) ∈ E} I Let A be the class of intervals containing all the I for all n, τ1 , . . . , τn and all L−mb sets E ∈ Rn , then A is an algebra of sets in C0 [0, 1] I The I’s are the cylinder sets upon which we will define Wiener measure, and then standard measure theoretic ideas to extend to all measurable subsets of the infinite dimensional space, C0 [0, 1] Brownian Motion Introduction to Brownian Motion as a Measure Donsker’s Invariance Principal Defining Wiener Measure Using Cylinder Sets I Given I, we define its measure as µ(I) = p (2π)n τ1 (τ2 Z 1 − τ1 ) · · · (τn − τn−1 ) u2 Z ··· e (u −u )2 (un −u )2 n−1 2 1 −···− − 2τ1 − 2(τ 2(τn −τn−1 ) 1 2 −τ1 ) du1 · · · dun . E I I Let B be the smallest σ−algebra generated by A, this is the class of Wiener measurable (W-mb) sets in C0 [0, 1] This extension of Wiener measure, also creates a probability measure on C0 [0, 1], and expectation w.r.t. Wiener measure will be referred to as a 1. Wiener integral or Wiener integration 2. Brownian motion expectation Brownian Motion Introduction to Brownian Motion as a Measure Donsker’s Invariance Principal Examples I Let A ∈ Rn×n with Aij we have τ1 τ1 A = τ1 τ2 τ1 τ2 = min(τi , τj ), i.e for the case n = 3, τ1 < τ 2 < τ 3 τ1 τ2 τ3 and in general we can write U = (u1 , . . . , un )> and Z Z > −1 1 µ(I) = p · · · e−U A U du1 . . . dun (2π)n det A E I Let β(·) be a BM, and 0 < τ1 < τ2 < 1, then Z b1 2 1 − u e 2τ1 du and P[a1 ≤ β(τ1 ) ≤ b1 ] = 2πτ1 a1 P[a1 ≤ β(τ1 ) ≤ b1 ∩ a2 ≤ β(τ2 ) ≤ b2 ] Z b2 Z b1 2 (u −u )2 1 − u − 2 1 e 2τ1 2(τ2 −τ1 ) du1 du2 = p (2π)2 τ1 (τ2 − τ1 ) a2 a1 Brownian Motion Introduction to Brownian Motion as a Measure Properties of Brownian Motion Useful Properties of Brownian Motion I I S∞ Theorem: Let PI∞= j=1 Ij where Ij ∩ Ik = ∅ ∀i 6= k and I, I1 , I2 , · · · ∈ A, then µ(I) = j=1 µ(Ij ) we will see that the BM, β(t), satisfies: 1. Almost every (AE) path is non-differentiable at every point 2. AE path satisfies a Hölder condition of order α < 12 , i.e. |β(s) − β(t)| ≤ L|s − t|α 3. 4. 5. 6. E[β(t)] = 0 E[β 2 (t)] = t, and so β(t) ∼ N(0, t) β(0) = 0, β(t) − β(s) ∼ N(0, t − s) E[β(t)β(s)] = min(s, t) Brownian Motion Introduction to Brownian Motion as a Measure Properties of Brownian Motion Useful Properties of Brownian Motion I Let E ∈ Rn (L − mb), 0 < τ1 < · · · < τn < 1, I = I(n; τ1 ; . . . ; τn ; E), then Z Z µ(I) = · · · p(τ1 , 0, u1 )p(τ2 − τ1 , u1 , u2 ) · · · E p(τN − τn−1 , un , un−1 ) du1 · · · dun where p(t, x, y ) = I (x−y )2 2t Note that p(t, x, y ) = ψ(t, x, y ), the fundamental solution for the initial value problem for the heat/diffusion equation ψt = I √ 1 e− 2πt 1 ψyy , 2 ψ(0, x, y ) = δ(y − x) µ is finitely additive since integrals are additive set functions Brownian Motion Introduction to Brownian Motion as a Measure Properties of Brownian Motion Useful Properties of Brownian Motion I Theorem 1: Let a > 0, 0 < γ < 1 2 and define Aa,γ = {β ∈ C0 [0, 1]; |β(τ2 ) − β(τ1 )| ≤ a|τ2 − τ1 |γ ∀τ1 , τ2 ∈ [0, 1]} For any interval I ⊂ C0 [0, 1] s.t. I ∩ Aa,γ = ∅ there is a K indepedent of a for which m(I) < Ka 4 − 1−2γ I Remark: Aa,γ is a compact set in C0 [0, 1] and eventually one can prove that AE β ∈ C0 [0, 1] satisfy some Hölder condition I Theorem 2: µ is countably additive on A, i.e. if In ∈ A, n ∈ N disjoint (Ij ∩ Ik = ∅, j 6= k ) then I= ∞ [ n=1 In ∈ A ⇒ µ(I) = ∞ X n=1 µ(In ) Brownian Motion Introduction to Brownian Motion as a Measure Properties of Brownian Motion Useful Properties of Brownian Motion I I Suppose F : C0 [0, 1] → R is a measurable functional, i.e. {β ∈ C0 [0, 1]; F [β] ≤ α} is measurable ∀α We can consider Z E[F ] = EW [F [β(·)]] = F [β(·)]δW , a Wiener integral I Consider Cx [0, t] = {f ∈ C[0, t]; f (0) = x}, then Z (y −x)2 1 √ P [β(0) = x, β(t) ∈ A] = e− 2t dy 2πt A I Furthermore 1 E [β(τ )] = √ 2πτ Z ∞ u2 ue− 2τ du = 0, ∀τ > 0 −∞ 1 E [g (β(τ1 ), . . . , β(τn ))] = p × (2π)n τ1 (τ2 − τ1 ) · · · (τn − τn−1 ) Z Z (un −un−1 )2 u2 (u2 −u1 )2 − 2τ1 − 2(τ −···− 2(τ −τ −τ1 ) n 1 2 n−1 ) du · · · du · · · g(u1 , . . . , un )e n 1 Brownian Motion Introduction to Brownian Motion as a Measure Properties of Brownian Motion Useful Properties of Brownian Motion I I Let us now consider, without proof, a large deviation result for BM: Theorem (The Law of the Iterated Logarithm for BM): Let β(s) ∈ C0 [0, ∞) be ordinary Brownian Motion, then (1) P lim sup √ t→∞ β(t) 2t ln ln t =1 =1 (2) P β(t) lim inf √ = −1 = 1 t→∞ 2t ln ln t Brownian Motion Introduction to Brownian Motion as a Measure Properties of Brownian Motion Dirac Delta Function I Let g be Borel measurable (B-mb), then Z ∞ u2 1 E[g(β(τ ))] = √ g(u)e− 2τ du 2πτ −∞ I Let g(u) = δ(u − x), using the Dirac delta function, then Z ∞ u2 x2 1 1 √ E[δ(β(t) − x)] = δ(u − x)e− 2t du = √ e− 2t 2πt −∞ 2πt thus u(x, t) = E[δ(β(t) − x)] = the heat equation ut = 1 uxx , u(x, 0) = δ(x) 2 x2 √ 1 e − 2t 2πt is the fundamental solution of Brownian Motion The Feynman-Kac Formula The Feynman-Kac Formula I Consider now V (x) ≥ 0 continuous and consider the equation ut = 1 uxx − V (x)u, u(x, 0) = δ(x), 2 then we can write h Rt i u(x, t) = E e− 0 V (β(s)) ds δ(β(t) − x) This is the Feynman-Kac formula I Example: V (x) = x2 1 x2 , ut = uxx − u, u(x, 0) = δ(x), then 2 2h 2 i Rt 2 1 u(x, t) = E e− 2 0 β (s) ds δ(β(t) − x) Brownian Motion The Feynman-Kac Formula The Feynman-Kac Formula I The following is clearly true: P[β(τ ) ≤ x] =P ({β ∈ C0 [0, τ ]; β(τ ) ∈ E = (−∞, x]}) = Z x u2 1 √ e− 2τ du, and similarly 2πτ −∞ With 0 = τ0 ≤ τ1 · · · ≤ τn we have P [β(τ1 ) ≤ x1 , . . . , β(τn ) ≤ xn ] = p Z xn Z ··· −∞ I u2 x1 e (u −u )2 (2π)−n/2 × (τ1 − τ0 )(τ2 − τ1 ) · · · (τn − τn−1 ) (un −u )2 n−1 2 1 −···− − 2τ1 − 2(τ 2(τn −τn−1 ) 1 2 −τ1 ) du1 · · · dun −∞ Hence with Aij = min(τi , τj ) 1 E [g (β(τ1 ), . . . , β(τn ))] = p × (2π)n |A| Z ∞ Z ∞ 1 > −1 ··· g(u1 , · · · , un )e− 2 U A U du1 · · · dun −∞ −∞ Brownian Motion The Feynman-Kac Formula Feynman-Kac Formula: Derivation I Let us consider the Wiener integral below, where expectation is taken over all of C0 [0, t] o n R E e− I I I t 0 V (β(τ )) dτ We will show that this is equal to the solution of the Bloch equation using an elementary proof of Kac We assume that 0 ≤ V (x) < M is bounded from above and non-negative; however, the upper bound will be relaxed We know Z t k ∞ Rt X (−1)k V (β(τ )) dτ /k ! e− 0 V (β(τ )) dτ = k =0 0 I Since V (·) is bounded we also have Z t 0< V (β(τ )) dτ < Mt I This allows us to use Fubini’s theorem as follows (Z k ) ∞ n Rt o X t − 0 V (β(τ )) dτ k E e = (−1) E V (β(τ )) dτ /k ! 0 k =0 0 Brownian Motion The Feynman-Kac Formula Feynman-Kac Formula: Derivation I Now let us consider the moments (Z k ) t µk (t) = E V (β(τ )) dτ 0 I Consider first k = 1 Z t Z tZ Z t Fubini E {V (β(τ ))} dτ = E V (β(τ )) dτ = 0 0 I 0 ∞ V (ξ) √ −∞ ξ2 1 e− 2τ dξ dτ 2πτ The case k = 2 is a bit more complicated (Z 2 ) Z t Z τ2 t Fubini E V (β(τ )) dτ = 2!E V (β(τ1 ))V (β(τ2 )) dτ1 dτ2 = 0 0 Z tZ 0 τ2 E {V (β(τ1 ))V (β(τ2 ))} dτ1 dτ2 = 2! 0 0 ξ2 Z tZ τ2 Z ∞ Z ∞ 2! 0 0 −∞ −∞ − 2τ1 e V (ξ1 )V (ξ2 ) √ − (ξ2 −ξ1 )2 e 2(τ2 −τ1 ) p dξ1 dξ2 dτ1 dτ2 2πτ1 2π(τ2 − τ1 ) 1 Brownian Motion The Feynman-Kac Formula Feynman-Kac Formula: Derivation I For general k we will proceed by defining the function Qn (x, t) as follows 1. Q0 (x, t) = x2 √ 1 e− 2t 2πt 2. Qn+1 (x, t) = Rt R∞ 0 −∞ (x−ξ)2 √ − 1 e 2(τ −t) 2π(τ −t) Rt V (ξ)Qn (ξ, τ ) dξ dτ Qk (x, t) dx I We have that µk (t) = k ! I By the boundedness of V (·) we also have, by induction, that n 0 ≤ Qn (x, t) ≤ (Mt) Q0 (x, t) n! P k Now define Q(x, t) = ∞ k =0 (−1) Qk (x, t) I 0 I This series converges for all x and t 6= 0 and |Q(x, t)| < eMt Q0 (x, t) I One can easily check that the definitions of the Qk (x, t)’s ensures that Q(x, t) satisfies the following integral equation Z tZ ∞ (x−ξ)2 1 1 − p Q(x, t) + √ e 2(t−τ ) V (ξ)Q(ξ, τ ) dξ dτ = Q0 (x, t) 2π 0 −∞ (t − τ ) Brownian Motion The Feynman-Kac Formula Feynman-Kac Formula: Derivation I It also follows that n Rt o Z E e− 0 V (β(τ )) dτ = ∞ Q(x, t) dx −∞ I Recall that his Wiener integral is over all of C0 [0, t], let us restrict this only to a < β(t) < b, thus n Rt o Z b E e− 0 V (β(τ )) dτ ; a < β(t) < b = Q(x, t) dx a I This tell us immediately that Q(x, t) ≥ 0 I Now we will relax the upper bound on V (·) by considering the function ( V (x), if V (x) ≤ M VM (x) = M, if V (x) ≥ M and we denote Q (M) (x, t) as the respective “Q” function Brownian Motion The Feynman-Kac Formula Feynman-Kac Formula: Derivation I By the additivity of Wiener measure we have that n Rt o n Rt o lim E e− 0 VM (β(τ )) dτ ; a < β(t) < b = E e− 0 V (β(τ )) dτ ; a < β(t) < b M→∞ I Furthermore, as M → ∞ the functions Q (M) (x, t) form a decreasing sequence with limM→∞ Q (M) (x, t) = Q(x, t) existing with the resulting limiting function, Q(x, t) satisfying the (Bloch) equation 1 ∂2Q ∂Q = − V (x)Q ∂t 2 ∂x 2 with the initial condition Q(x, t) → δ(x) as t → 0 Brownian Motion The Feynman-Kac Formula Feynman-Kac Formula: Derivation Variation I I I I Recall the integral equation solved by Q(x, t) Z tZ ∞ (x−ξ)2 x2 1 1 1 − p e 2(t−τ ) V (ξ)Q(ξ, τ ) dξ dτ = √ e− 2t Q(x, t)+ √ 2πt 2π 0 −∞ (t − τ ) R∞ −st Let us define Ψ(x) = −∞ Q(x, t)e dt with s > 0, this is the Laplace transform of Q(x, t) Now multiply the integral equation by e−st and integrate out t to get the equation satisfied by the Laplace transform of Q(x, t) Z ∞ √ √ 1 1 √ e− 2s|x−ξ| V (ξ)Ψ(ξ) dξ = √ e− 2s|x| Ψ(x) + 2s −∞ 2s It is easy to verify that Ψ(x) also satisfies the following differential equation 1 00 Ψ − (s + V (x))Ψ = 0, with the following conditions 2 1. Ψ → 0 as |x| → ∞ 2. Ψ0 is continuous except at x = 0 3. Ψ0 (−0) − Ψ0 (−0) = 2 Brownian Motion The Feynman-Kac Formula Explicit Representation of Brownian Motion Explicit Representation of Brownian Motion I Suppose that F [β] = t Z E Rt 0 β 2 (s) ds, then it follows β 2 (s) ds t Z = 0 h Fubini 0 Rt β(s) ds i Z t h i t2 E β 2 (s) ds = s ds = 2 0 I To compute E e I Consider the eigenvalue problem for this integral equation Z t ρ u(s) min(τ, s) ds = u(τ ) 0 , we need to do some classical analysis 0 I Find eigenvalues ρ0 , ρ1 , . . . and corresponding orthonormalized Rt eigenfunctions u0 (τ ), u1 (τ ), . . . with 0 uj (τ )uk (τ ) dτ = δjk , ∀j, k ≥ 0 Brownian Motion The Feynman-Kac Formula Explicit Representation of Brownian Motion Explicit Representation of Brownian Motion I For t > τ we have Z t Z τ su(s) ds + ρ τ u(s) ds = u(τ ) ρ 0 τ d dτ t Z =⇒ ρτ u(τ ) − ρτ u(τ ) + ρ u(s) ds = u 0 (τ ) τ d dτ =⇒ −ρu(τ ) = u 00 (τ ) Thus u 00 (τ ) + ρu(τ ) = 0 and with u(0) = 0, u 0 (t) = 0 we get 1 2 π2 ρk = (k + ) q2 t 2 k = 0, 1, 2, . . . uk (s) = 2 sin (k + 1 ) πs t I 2 t By the spectral theorem the integral equation kernel can be represented as: min(s, τ ) = ∞ X uk (s)uk (τ ) ρk k =0 Brownian Motion The Feynman-Kac Formula Explicit Representation of Brownian Motion Explicit Representation of Brownian Motion I Let α0 (ω), α1 (ω), . . . be i.i.d. N(0, 1), then we claim that the following is an explicit representation of BM ∞ X αk (ω)uk (τ ) = β(τ ) √ ρk (2.1) k =0 I This is a Fourier series with random coefficients and we will prove that this converges for AE path ω with the following properties 1. We use ω to denote an individual sample of i.i.d. N(0, 1) αi (ω)’s 2. E[αi (ω)] = 0, ∀i ≥ 0 3. E[αi (ω)αj (ω)] = δij , ∀i, j ≥ 0 I This is the simplest version of the Karhunen-Loève expansion of stochastic processes Brownian Motion The Feynman-Kac Formula Explicit Representation of Brownian Motion Explicit Representation of Brownian Motion (Proof) I We now use the representation (2.1) to compute some expectations w.r.t. the αi ’s ∼ N(0, 1) # "∞ X αk (ω) i.i.d. N(0,1) & Fubini E = √ uk (τ ) ρk k =0 ∞ ∞ X X E[αk (ω)]uk (τ ) 0 × uk (τ ) = = 0 = E[β(τ )] √ √ ρk ρk k =0 k =0 I We now use the representation (2.1) to compute some expectations "∞ # ∞ X αk (ω) X i.i.d.N(0,1) αl (ω) E = √ uk (τ ) √ ul (τ ) ρk ρl k =0 l=0 ∞ h i X uk2 (τ ) = min(τ, τ ) = τ = E β 2 (τ ) ρk k =0 Brownian Motion The Feynman-Kac Formula Explicit Representation of Brownian Motion Explicit Representation of Brownian Motion (Proof) I Similarly we compute # "∞ ∞ X X αk (ω) i.i.d.N(0,1) αl (ω) E = √ uk (τ ) √ ul (s) ρk ρl k =0 l=0 ∞ X uk (τ )uk (s) = min(τ, s) = E [β(τ )β(s)] ρk k =0 I We have computed the mean, variance, and correlation of the process defined in (2.1), and it is clear that it is ∼ N(0, τ ) and hence Brownian motion, β(τ ) Brownian Motion The Feynman-Kac Formula The Karhunen-Loève Expansion An Introduction to the Karhunen-Loève Expansion I Karhunen-Loève (KL) expansion writes the stochastic processes Y (ω, t) as a stochastic linear combination of a set of orthonormal, deterministic functions in L2 , {ei (t)}∞ i=0 Y (ω, t) = ∞ X Zi (ω)ei (t) i=0 1. Given the covariance function of the random process Y (ω, t) as CYY (s, τ ) the KL expansion is Y (ω, t) = ∞ p X λi ξi (ω)φi (t) i=0 2. Here λi and φi (t) are the eigenvalues and L2 -orthonormal eigenfunctions of the covariance function and ξi (ω)φi (t) are i.i.d. √ random variables whose distribution depends on Y (ω, t), i.e. Zi (ω) = λi ξi (ω), and ei (t) = φi (t) 3. It can be shown that such an expansion converges to the stochastic process in L2 (in distribution) Brownian Motion The Feynman-Kac Formula The Karhunen-Loève Expansion An Introduction to the Karhunen-Loève Expansion 4. By the spectral theorem, we can expand the covariance, thought of as an integral equation kernel, as follows ∞ X CYY (s, τ ) = λi φi (s)φi (τ ) i=0 5. Here λi and φi (t) are the eigenvalues and eigenfunctions of the following integral equation Z ∞ CYY (s, τ )φj (τ ) dτ = λj φj (s) 0 I For ordinary BM, Y (ω, t) = β(t), we have from above 1. CYY (s, τ ) = Cββ (s, τ ) = min(s, τ ) 2 , where ρj = (j + 12 )2 πs2 q 3. φj (t) = uj (t) = 2s sin((j + 21 ) πts ) 2. λj = 1 ρj 4. ξj (ω) = αj (ω) ∼ N(0, 1) P αj (ω)uj (t) 5. Y (ω, t) = ∞ = β(t) √ j=0 ρ j Brownian Motion The Feynman-Kac Formula Explicit Computation of Wiener Integrals Explicit Computation of Wiener Integrals I We are now in position to compute R P i h Rt t ∞ = E e 0 k =0 E e 0 β(s) ds P R ∞ t E e k =0 0 α √k ρk uk (s) ds indep. = ∞ Y α √k E e ρk αk uk (s) √ ρk Rt 0 ds = uk (s) ds = k =0 ∞ Y 1 e 2ρk Rt ( 0 2 uk (s) ds) 1 = e2 R t R t P∞ 0 0 k =0 uk (s)uk (τ ) ρk ds dτ = k =0 1 I Rt Rt 1 e 2 0 0 min(s,τ ) ds dτ = e2 We have used the following results u2 τ2 0 [( 2 Rt +(τ (t−τ ))] dτ t3 =e6 1. E[eαu ] = e 2 , with α ∼ N(0, 1) via moment generating function R R R 2 2. 0t min(s, τ ) ds = 0τ s ds + τt τ ds = τ2 + (τ (t − τ )) Brownian Motion The Feynman-Kac Formula Explicit Computation of Wiener Integrals Explicit Computation of Wiener Integrals I Moreover λ2 E e− 2 Rt indep. = 0 β 2 (s) ds ∞ Y " " E 2 =E e 2 α2 k −λ e 2 ρk # = k =0 = = ∞ Y k =0 ∞ Y k =0 1 √ 2π − λ2 ∞ Y k =0 Z 1 q 1+ 2 ∞ e − α2 α2 k k =0 ρk # 1 √ 2π Z P∞ 2 1+ λ ρ k ∞ e 2 2 − λ2 α ρ k −∞ dα −∞ λ2 ρk 1 = p cosh(λt) = s Q∞ k =0 1 1+ λ2 t 2 (k + 12 )2 +π 2 e− α2 2 dα Brownian Motion The Feynman-Kac Formula The Schrödinger Equation The Schrödinger Equation I Let us review the Schrödinger equation from quantum mechanics 1. The “standard," time-dependent Schrödinger equation " # −~2 ∂ ∆ + V (x, t) Ψ(x, t) = Ĥ(x, t)Ψ i~ Ψ(x, t) = ∂t 2m 2. We can make the equation dimensionless as 1 ∂ −i ψ(x, t) = ∆ − V (r, t) ψ(x, t) = H(x, t)ψ ∂t 2 3. We also are interested in the spectral properties of the time-independent problem 1 ∆ − V (x, t) ψ(x, t) = H(x, t)ψ = λψ 2 Brownian Motion The Feynman-Kac Formula The Schrödinger Equation The Schrödinger and Bloch Equations I We now arrive at the Bloch equation 1. Consider transformation (analytic continuation) of the Schrödinger to imaginary time, τ = it, this gives us the Bloch equation, but is sometimes also called the Schrödinger equation (going back to u(x, t)) 1 ∂u(x, t) = ∆u(x, t) − V (x, t)u(x, t) ∂τ 2 2. The time dependent Bloch equation can be solved via separation of variables as u(x, t) = U(x)T (t), and so we apply this to the Bloch equation ∂u(x, t) = U(x)T 0 (t) = ∂t 1 ∆U(x) − V (x, t)U(x) T (t) 2 Brownian Motion The Feynman-Kac Formula The Schrödinger Equation The Schrödinger and Bloch Equations 3. Placing the time and space dependent on different sides of the equation gives 1 ∆ − V (x, t) U(x) T 0 (t) =λ= 2 , where λ is constant T (t) U(x) 4. Thus we have that T (t) and U(x) satisfy the following equations T 0 (t) − λT (t) = 0, 1 ∆ − V (x, t) U(x) = λU(x) 2 5. Thus the λj ’s and ψj (x, t)’s are eigenvalues and eigenfunctions of the above eigenvalue problem, and the solution by separation variables is u(x, t) = ∞ X j=1 cj e−λj t ψj (x), where, cj = Z ∞ u0 (x)ψj (x) dx −∞ Brownian Motion The Feynman-Kac Formula The Schrödinger Equation The Schrödinger and Bloch Equations I i h 1 Rt 2 Let λ = 1, as t → ∞, E e− 2 0 β (s) ds = √ lim t→∞ I 1 cosh(t) ∼ √ −t 2e 2 and i h 1 Rt 2 1 1 ln E e− 2 0 β (s) ds = − . t 2 Theorem: If V (y ) → ∞ as |y | → ∞, then lim t→∞ i h Rt 1 ln E e− 0 V (β(s)) ds = −λ1 , t where λ1 is the lowest eigenvalue of the Bloch equation 1 00 ψ (y ) − V (y )ψ(y ) = λψ(y ) 2 Brownian Motion The Feynman-Kac Formula The Schrödinger Equation The Schrödinger and Bloch Equations I Feynmann-Kac: Let V be measurable and bounded below, then the solution of the Bloch equation 1 uxx − V (x)u, u(x, 0) = u0 (x) 2 h Rt i is u(x, t) = Ex e− 0 V (β(s)) ds u0 (β(t)) ut = I This equation is the imaginary time analog of the Schrödinger 1 00 ψ (y ) − V (y )ψ(y ) = λψ(y ) 2 Equation 1. Special case: V ≡ 0: Ex [u0 (β(t))] = √ 1 2πt Z ∞ −∞ u0 (y )e− (x−y )2 2t dy = u(x, t) Brownian Motion The Feynman-Kac Formula The Schrödinger Equation Another special case x2 , 2 u0 ≡ 1: h 1 Rt 2 i h 1 Rt i 2 u(x, t) = = Ex e− 2 0 β (s) ds · 1 = E0 e− 2 0 (β(s)+x) ds i h Rt x2t 1 Rt 2 =e− 2 E e−x 0 β(s) ds− 2 0 β (s) ds " # 2 2. For V (x) = =e− x2t 2 =e− x2t 2 ∞ Y 2 − x2 t k =0 ∞ Y =e E e k =0 =e− x2t 2 −x αk k =0 √ρk P∞ " E e α −x √ρk 1 √ 2π 0 Rt 0 k Z Rt ∞ e uk (s) ds− 12 α2 uk (s) ds− 12 ρ k k −x √α ρ Rt k 0 α k k =0 ρk P∞ # 2 uk (s) ds− α2 (1+ ρ1 ) k −∞ x2 1 p e2 cosh(t) R t R t P∞ 0 0 k =0 uk (s)uk (τ ) ρk +1 ds dτ dα Brownian Motion The Feynman-Kac Formula The Schrödinger Equation I Define R(s, τ ; −λ2 ) such that Z t min(s, τ ) = λ2 min(s, ξ)R(ξ, τ ; −λ2 ) dξ 0 Note that R(s, τ ; −1) = − I P∞ k =0 uk (s)uk (τ ) . ρk +1 Consider − ∞ ∞ X uk (s)uk (τ ) X uk (s)uk (τ ) + ρk + λ 2 ρk k =0 k =0 Z tX ∞ ∞ uk (s)uk (ξ) X ul (ξ)ul (τ ) = λ2 dξ ρk ρk + λ 2 0 k =0 l=0 Brownian Motion The Feynman-Kac Formula The Schrödinger Equation I I For 0 ≤ s ≤ t we have ( cosh(λ(t−τ )) sinh(λs) − 2 λ cosh(λt) R(s, τ ; −λ ) = sinh(λτ ) − cosh(λ(t−s)) λ cosh(λt) s≥τ Thus u(x, t) = p I s≤τ Rt Rt x2 x 2 tanh t 1 1 e− 2 (t+ 0 0 R(s,τ ;−1) ds dτ ) = p e− 2 cosh(t) cosh(t) 2 Exercise: compute t) for Vi (x) = x2 , u0 (x) = x. Hint: the solution is h 1 R tu(x, 2 u(x, t) = Ex e− 2 0 β (s) ds β(t) . Calculate h i 1 Rt 2 ũ(x, t, λ) =Ex eλβ(t)− 2 0 β (s) ds , u(x, t) = d ũ(x, t, λ) . dλ λ=0 Brownian Motion The Feynman-Kac Formula Proof of the Arcsin Law Proof of the Arcsin Law I Theorem: Let X1 , X2 , . . . be i.i.d. r.v.’s with P E [Xi ] = 0, Var (Xi ) = 1, and Nn is the number of partial sums Sj = ji=1 Xi out of S1 , . . . , Sn which are ≥ 0: α<0 0 √ Nn lim P < α = Σ(α) = π2 arcsin α 0 ≤ α ≤ 1 n→∞ n 1 α≥1 I Proof: (Using the Feynman-Kac formula and Donsker’s Invariance Principal) Define the random step function (S √1 τ =0 (n) X (τ ) = Sin i−1 √ < τ ≤ ni n n The invariance principle states that for a large class of functionals F and F ∈F h h i i lim P F X (n) (·) ≤ α = PBM [F [β(·)] ≤ α] (2.2) n→∞ Brownian Motion The Feynman-Kac Formula Proof of the Arcsin Law Proof of the Arcsin Law I For example, let t Z F [β] = 0 1 + sgn[β(s)] ds, where sgn(x) = 2 ( 1 −1 I Then (2.2) says that Z 1 1 + sgn[β(s)] Nn ≤ α = PBM ds ≤ α lim P n→∞ n 2 0 I We drop the BM from the probabilities as it is understood of the Brownian motion that is positive x ≥0 x <0 Brownian Motion The Feynman-Kac Formula Proof of the Arcsin Law Proof of the Arcsin Law I Let t Z σ(α, t) = P 0 I 1 + sgn[β(s)] ds ≤ α 2 Then for λ > 0 we can define the Laplace Transform/Moment Generating Function of σ(α, t) Z ∞ R t 1+sgn[β(s)] ds 2 = e−λα dσ(α, t) E e−λ 0 0 I Now define Rt u(x, t; λ) = E e−λ 0 1+sgn[β(s)] 2 ds δ(β(t) − x) Brownian Motion The Feynman-Kac Formula Proof of the Arcsin Law Proof of the Arcsin Law I By Feynman-Kac this is a solution to the following PDE u(x, t; λ)t = 1 u(x, t; λ)xx − λV (x)u(x, t; λ), 2 ( where V (x) = I We also realize that Z ∞ Z u(x, t; λ) dx = −∞ Z ∞ Rt E e−λ 0 u(x, 0; λ) = δ(x) 1 x ≥0 0 x <0 1+sgn[β(s)] 2 ds Fubini δ(β(t) − x) dx = −∞ ∞ E e−λ R t 1+sgn[β(s)] 0 2 ds Rt δ(β(t) − x) dx = E e−λ 0 −∞ ∞ Z 0 e−λα dσ(α, t) 1+sgn[β(s)] 2 ds = Brownian Motion The Feynman-Kac Formula Proof of the Arcsin Law Proof of the Arcsin Law I It is know that u(x, t; λ) also solves the following integral equation −x 2 1 e 2t − u(x, t; λ) = √ 2πt Z t Z ∞ −(x−ξ)2 1 e 2(t−τ ) λ dτ dξV (ξ)u(ξ, τ ; λ) p 2π(t − τ ) 0 −∞ I Now we apply the heat equation operator, ∂ ∂t − 1 ∂2 2 ∂x 2 ∂u 1 ∂2u − = 0 − λV (x)u(x, t; λ) ∂t 2 ∂x 2 I And we the Laplace transform of u(x, t; λ) Z ∞ Ψ(x, s; λ) = e−st u(x, t; λ)dt −∞ to this Brownian Motion The Feynman-Kac Formula Proof of the Arcsin Law Proof of the Arcsin Law I If we take the Laplace transform of the integral equation we get √ 1 Ψ(x, s; λ) = √ e− 2s|x| 2s Z ∞ √ 1 dξV (ξ)Ψ(ξ, s; λ) √ e− 2s|x−ξ| −λ 2s −∞ I This is equivalent to the following ordinary differential equation (ODE) 1 00 Ψ (x) − (s + λV (x))Ψ(x) = 0, Ψ → 0 as |x| → ∞ 2 Ψ(x) and Ψ0 (x) is continuous at x 6= 0, and Ψ0 (0− ) − Ψ0 (0+ ) = 2 Brownian Motion The Feynman-Kac Formula Proof of the Arcsin Law Proof of the Arcsin Law I The solution to the above ODE is ( √ Ψ(x, s; λ) = I x ≥0 x <0 Thus we have that Z ∞ −∞ I √ 2 √ √ e− 2(s+λ)x s+λ+ s √ √ 2 √ √ e− 2sx s+λ+ s 1 Ψ(x, s; λ) dx = p s(s + λ) So we have the following Z ∞ Z ∞ Z ∞ Ψ(x, s; λ) dx = e−st u(x, t; λ) dx ds = −∞ −∞ 0 Z ∞ Z ∞ 1 e−st e−λα dσ(α, t) ds = p s(s + λ) 0 0 Brownian Motion The Feynman-Kac Formula Proof of the Arcsin Law Proof of the Arcsin Law I The last line test us that we know the Laplace transform of Z ∞ F (t) = e−λα dσ(α, t) 0 I The inverse Laplace transform of √ λt F (t) = e− 2 Io ( I 1 s(s+λ) λt )= 2 ∞ Z tells us that e−λα σ 0 (α, t) dα 0 Which is itself the Laplace transform of σ 0 (α, t), so we have √1 0<α<t 0 σ (α, t) = π α(t−α) 0 α>t Brownian Motion The Feynman-Kac Formula Proof of the Arcsin Law Proof of the Arcsin Law I We now integrate the previous result Z α −∞ I σ 0 (ᾱ, t) d ᾱ = σ(α, t) = 0 2 π 1 Setting t = 1 we get the Arcsin Law 0 √ σ(α, 1) = Σ(α) = π2 arcsin t 1 arcsin pα t 0<α 0<α<t α>t 0<α 0 < α < 1 Q. E. D. α>1 Brownian Motion The Feynman-Kac Formula Proof of the Arcsin Law Another Wiener Integral I We wish to compute the probability of P max β(s) ≤ α 0≤s≤t I By Donsker’s Invariance Principal this is equal to S1 S2 Sn lim max √ , √ , · · · , √ ≤ α = P max β(s) ≤ α = H(α, t) n→∞ 0≤s≤t n n n I Consider the step-function potential ( 1 Vα (x) = 0 I x ≥α x <α Since β(·) is a continuous function AE, if max0≤s≤t β(s) ≤ α then Vα (β(s)) = 0 on a set of positive measure Brownian Motion The Feynman-Kac Formula Proof of the Arcsin Law Another Wiener Integral I Consider the following Wiener integral h i Rt lim E e−λ 0 Vα (β(s)) ds = H(α, t) λ→∞ I I This is because the λ limit kills walks that exceed α and only count the walks that satisfy the condition for a fixed λ this is, by Feynman-Kac, the solution to u(x, t; λ)t = I 1 u(x, t; λ)xx − λV (x)u(x, t; λ), 2 ( 1 x ≥α where V (x) = 0 x <α u(x, 0; λ) = 1 The solution of the PDE is very similar to the solution of the PDE from the Arcsin Law, and is left to the reader r Z √α u2 t 2 H(α, t) = e− 2 du π 0 Brownian Motion Advanced Topics Action Asymptotics Action Asympotics: A Heuristic for Wiener Integrals I Von Neumann proved that there is no translationally invariant Haar measure in function space; Wiener measure is not translationally invariant I Consider the following problem where we write our heuristic via a “flat” integral 1 F [β]e− 2 E {F [β]} “ = ” I 0 [β 0 (τ )]2 dτ δβ Here we define the Action as A[β] = − I Rt 1 2 Z t 0 2 β (τ ) dτ 0 This is obviously a heuristic, as BM is nondifferentiable AE Brownian Motion Advanced Topics Action Asymptotics Action Asympotics: A Heuristic for Wiener Integrals I Now consider computing the following with Action Asymptotics h √1 R t i β(s) ds E e 0 I We first compute this using our standard techniques i h √1 R t 1 R t P∞ √ β(s) ds = E e 0 k =0 E e 0 1 P∞ √ E e k =0 α √k ρk Rt 0 uk (s) ds indep. = ∞ Y k =0 I And thus 1 √ 2π αk uk (s) √ ρk ∞ Z e −∞ h √1 R t i t3 β(s) ds lim ln E e 0 = →0 6 √α ρk ds Rt 0 = 2 uk (s) ds − α e 2 dα = Brownian Motion Advanced Topics Action Asymptotics Action Asympotics: A Heuristic for Wiener Integrals I Let’s “derive” the action asymptotics heuristic with a construction due to Kac and Feynman by considering n Rt o F (t) = E e− 0 V (β(τ )) dτ where β(·) ∈ C0 [0, t], and the expectation is taken w.r.t. Wiener measure I Since we assume that V (·) is continuous and non-negative, and Rt β(·) ∈ C0 [0, t]) is continuous, F (t) exists as 0 V (β(τ )) dτ is measurable I Now let us consider a discrete approximation of this Wiener integral by breaking it up into N sized time intervals of size t/N, which gives us F (t) from bounded convergence and the Riemann summability o n t PN tk F (t) = lim E e− N k =1 V (β( N )) N→∞ Brownian Motion Advanced Topics Action Asymptotics Action Asympotics: A Heuristic for Wiener Integrals I If we consider the expectation in the limit we can rewrite it as follows Z ∞ Z ∞ o n t PN PN tk ··· e−h k =1 V (βk ) × lim E e− N k =1 V (β( N )) = lim N→∞ N→∞ −∞ −∞ P(0, β1 ; h)P(β1 , β2 ; h) · · · P(βN−1 , βN ; h) dβ1 dβ2 · · · dβN where we have 1. h = Nt 2. βk = β(kh) 3. P(βk −1 , βk ; h) = √ 1 e− 2πh (βk −βk −1 )2 2h I This limit exists and is equal to the Wiener integral I However, Feynman chose to rewrite the above as (suppressing the limit) with β0 = 0 ( 1 (2πh)N/2 Z ∞ Z ∞ ··· −∞ −h e −∞ PN k =1 V (βk )+ 12 PN k =1 βk −βk −1 h 2 ) dβ1 dβ2 · · · dβN Brownian Motion Advanced Topics Action Asymptotics Action Asympotics: A Heuristic for Wiener Integrals I If we look at the exponent in Feynman’s we notice that ( N ) 2 ) 2 Z t( N X 1 X βk − βk −1 1 dβ h→0 V (βk ) + h→ + V (β(τ )) dτ 2 h 2 dτ 0 k =1 k =1 I This is the Hamiltonian the along the path, β(τ ), and with the classical action along the path is ) 2 Z t( 1 dβ − V (β(τ )) dτ 2 dτ 0 I thus Feynman writes the above integral instead as n Rt o Z − F (t) = E e− 0 V (β(τ )) dτ = e Rt 0 1 2 ( dβ dτ ) 2 +V (β(τ )) dτ d(path) Brownian Motion Advanced Topics Action Asymptotics Action Asympotics: A Heuristic for Wiener Integral I h 1 √ i How does E e F [ β] behave as → 0? I We can approach this with Action Asymptotics h 1 √ i E e F [ β] “ = ” I Now let 1 1 e [F [ω]− 2 Rt 0 [ω 0 (s)]2 ds] δβ Using Laplace asymptotics the above will behave like 1 e I R 2 β] − 12 0t [β 0 (s)] ds δβ e √ β = ω “=” I √ 1 e F[ R supω∈C ∗ [0,t] [F [ω]− 12 0t [ω 0 (s)]2 ds] 0 Where the space C0∗ [0, t] is made up functions, ω(t), with 1. ω(t) continuous in [0, t] 2. ω(0) = 0 3. ω 0 (t) ∈ L2 [0, t] Brownian Motion Advanced Topics Action Asymptotics Action Asympotics: Examples I A conjecture using Action Asymptotics h 1 √ i lim ln E e F [ β] = →0 I Consider F [β] = Rt 0 Z 1 t 0 F [ω] − [ω (s)]2 ds 2 0 ω∈C0∗ [0,t] sup β(s) ds h 1 √ i h √1 R t i β(s) ds E e F [ β] = E e 0 I From the conjecture we have that h lim ln E e →0 1 √ Rt 0 β(s) ds i t Z = sup ω∈C0∗ [0,t] 0 1 ω(s) ds − 2 Z t 0 2 [ω (s)] ds 0 Brownian Motion Advanced Topics Action Asymptotics Action Asympotics: Examples I From the calculus of variations we have that the Euler equation for following maximum principle is Z t Z 1 t 0 2 sup [ω (s)] ds =⇒ ω(s) ds − 2 0 ω∈C0∗ [0,t] 0 1. 1 + ω 00 (s) = 0 2. ω(0) = 0 3. ω 0 (t) = 0 I 2 The solution is ω(s) = − s2 + ts and ω 0 (s) = −s + t so Z t 2 Z s 1 t t3 − + ts ds − [s − t]2 ds = 2 2 0 6 0 Brownian Motion Advanced Topics Brownian Scaling Brownian Scaling I Recall some basic properties of the BM, β(·) and constant, c: 1. 2. 3. 4. 5. 6. I β(τ ) ∼ N(0, τ ) β(cτ ) ∼ N(0, cτ ) √ cβ(τ ) ∼ N(0, cτ ) E[β(τ )β(s)] = min(τ, s) E[β(cτ )β(cs)] = c min(τ, s)√ √ E[β(cτ )β(cs)] = E[ cβ(τ ) cβ(s)] = cE[β(τ )β(s)] = c min(τ, s) Now consider the following h i E esup0≤s≤t β(s) = i h √ E esup0≤τ ≤1 tβ(τ ) = h 1 i √ E e sup0≤τ ≤1 β(τ ) h i E esup0≤τ ≤1 β(tτ ) = h 1 β(τ ) i √ t sup E e 0≤τ ≤1 t = using the substitution t = 1 Brownian Motion Advanced Topics Brownian Scaling Action Asympotics: Examples I So we now have that i h i h 1 √ 1 lim ln E esup0≤s≤t β(s) = lim ln E e sup0≤τ ≤1 β(τ ) t→∞ t →0 By Action Asymptotics we have h 1 i √ lim ln E e sup0≤τ ≤1 β(τ ) = I " # Z 1 1 0 2 [ω (τ )] dτ sup sup ω(τ ) − →0 2 0 ω∈C0∗ [0,1] 0≤τ ≤1 a2 1 = max a − = a>0 2 2 I The supremum comes on straight lines, that minimize arc-length i.e. the second term, so consider ω(τ ) = aτ , and a = 1 is the maximizer Brownian Motion Advanced Topics Brownian Scaling Action Asympotics: Examples I Consider a more complicated problem for Action Asymptotics is h √ i √ 1 E G ( β(·)) e F ( β(·)) h 1 √ i lim “=” →0 E e F ( β(·)) i h √ √ 2 0 1 1 Rt E G ( β(·)) e F ( β(·))− 2 0 [β (s)] ds δβ h E e 1F ( i R √ β(·))− 12 0t [β 0 (s)]2 ds = δβ √ We now change variables with x(·) = β(·) h i 2 1 Rt 0 1 E G (x(·)) e [F (x(·))− 2 0 [x (s)] ds] δx h 1 i 1 Rt 0 2 E e [F (x(·))− 2 0 [x (s)] ds] δx Brownian Motion Advanced Topics Brownian Scaling Action Asympotics: Examples I As → 0 the exponential term goes to something like a “delta" function in function space and we get = G [ω ∗ (·)] where ω ∗ (·) = argsup [F [ω] − A[ω]] ω∈C0∗ [0,t] I We now apply this to some PDE problems: Burger’s Equation ut + uux = uxx , −∞ ≤ x ≤ ∞, t > 0 2 Z ∞ u(x, 0) = u0 (η) dη = o(x 2 ) as |x| → ∞ u0 (x), 0 I We now apply the Hopf-Cole transformation, if we define the solution to (x,t) Burger’s equation u(x, t) = − vvx(x,t) = −∂x [ln v (x, t)] then v (x, t) satisfies 1 Rx vt = vxx , v (x, 0) = e− 0 u0 (η) dη 2 Brownian Motion Advanced Topics Brownian Scaling Action Asympotics: Examples I So by Feynman-Kac we can write the solution as Z ∞ (x−y )2 1 Ry 1 v (x, t; ) = √ e− 0 u0 (η) dη e− 2t dy 2πt −∞ I We now apply the Hopf-Cole transformation (taking the logarithmic derivative) (y −x)2 1 Ry dy (x−y ) − 0 u0 (η) dη+ 2t e t −∞ 2 R ∞ − 1 R0y u0 (η) dη+ (y −x) dy 2t R∞ u(x, t; ) = −∞ I I I e Ry 2 Now let F (y ) = 0 u0 (η) dη + (y −x) , this is the function that Action 2t Asymptotics tells us to minimize (due to the negative sign) Note that lim|y |→∞ Fy(y2 ) = 2t1 by the assumptions, and so there is a minimum, y (x, t) = argmin F (y ) Hopf showed that if at (x, t) there is a single minimizer to F (y ) then lim u(x, t; ) = →0 x − y (x, t) = u0 (y (x, t)) t Brownian Motion Advanced Topics Brownian Scaling Action Asympotics: Examples I Consider the related equation ut + uux = uxx − V 0 (x), 2 Z −∞ ≤ x ≤ ∞, t >0 ∞ u(x, 0) = u0 (x), u0 (η) dη = o(x 2 ) as |x| → ∞ 0 I Again we use the Hopf-Cole transformation to get vt = I 1 vxx − V 0 (x)v , 2 1 v (x, 0) = e− Rx 0 u0 (η) dη And so we can write down the solution to the transformed equation via Feynman-Kac √ √ 1 R β(t) u (η) dη 1 Rt 0 v (x, t; ) = Ex e− 0 V ( β(s)) ds− 0 hR i R √β(t)+x √ − 1 0t V ( β(s)+x) ds 0 u0 (η) dη = E0 e Brownian Motion Advanced Topics Brownian Scaling Action Asympotics: Examples I We now take apply the Hopf-Cole transformation and get i h √ √ 1 E G[ β(·)]e− F [ β(·)] h 1 √ i where we define u(x, t; ) = E e− F [ β(·)] t Z F [β(·)] = 0 t Z G[β(·)] = 0 √ V ( β(s)) ds − √ Z β(t) u0 (η) dη 0 √ √ V 0 ( β(s) + x) ds + u0 ( β(t) + x) Brownian Motion Advanced Topics Brownian Scaling Action Asympotics: Examples I By Action Asymptotics we have that lim u(x, t; ) = G[ω ∗ (·)] where ω ∗ (·) = arginf [F [ω] + A[ω]] →0 I ω∈C0∗ [0,t] If for (x, t) ∃! minimizer, ω ∗ , then the limit exists and is Z t G[ω ∗ (t)] = u(x, t) = V 0 (ω ∗ (s) + x) ds + u0 (ω ∗ (t) + x) 0 I Now consider the related variational problem "Z Z t inf ∗ ω∈C0 [0,t] I ω(t)+x V (ω(s) + x) ds 0 u0 (η) dη + 0 1 2 We refer to the functional to be minimized as H[ω(·)] t Z 0 # [ω 0 (s)]2 ds Brownian Motion Advanced Topics Brownian Scaling Action Asympotics: Examples I To arrive derive an equivalent system via the Calculus of Variations we need to form the Frechet derivative, in the direction of the arbitrary function, Ψ, as follows Z t dH[ω + hΨ] = V 0 (ω(s) + x)Ψ(s) ds + u0 (ω(t) + x)Ψ(t) δH|Ψ = dh 0 h=0 Z t ω 00 (s)Ψ(s) ds + ω 0 (t)Ψ(t) − 0 I Note that the last two terms come from the following computation Z t dJ[ω + hΨ] def 1 J[ω(·)] = [ω 0 (s)]2 ds =⇒ 2 0 dh h=0 Z t Z t 1 = [ω 0 (s) + hΨ0 (s)]2 ds = [ω 0 (s) + hΨ0 (s)]2 ds 2 0 0 Z t Z t = ω 0 (s)Ψ0 (s) ds = ω 0 (s) dΨ0 (s) 0 0 Brownian Motion Advanced Topics Brownian Scaling Action Asympotics: Examples I We now integrate by parts using the natural boundary conditions 1. ω(0) = 0 2. ω 0 (0) = 0 Z t ω 0 (s) dΨ0 (s) = ω 0 (t)Ψ0 (s) − 0 I t Z ω 00 (s)Ψ ds 0 So the solution to this problem is 1. V 0 (ω(s) + x) = ω 00 (s) for 0 ≤ s ≤ t 2. ω(0) = 0 3. ω 0 (t) = −u0 (ω(s) + x) I We can now apply this Hpof’s result with V ≡ 0 1. ω 00 (s) = 0 for 0 ≤ s ≤ t 2. ω(0) = 0 3. ω 0 (t) = −u0 (ω(s) + x) I The solution is then very simply 1. ω(s) = cs for some constant, c 2. ω 0 (s) = c = −u0 (ct + x) y (x,t)−x 3. Let c = = −u0 (y (x, t)) or u0 (t(x, t)) = t I With a unique y (x, t) we get a unique ω ∗ (s) = x−y (x,t) t x−y (x,t) t s Brownian Motion Advanced Topics Brownian Scaling Action Asympotics I I We now consider some tools with the “flat integral" The Cameron-Martin Translation Formula E {F [β + y ]} , with y ∈ C0 [0, t] I We now use the “flat integral" 1 F [β + y ]e− 2 E {F [β + y ]} “ = ” 1 F [ω]e− 2 “=” 1 “ = ”e− 2 Rt 0 Rt 0 [ω (s)]2 ds 0 Rt 0 [β 0 (s)]2 ds δβ, and let ω = β + y (s)−y 0 (s)]2 ds Rt δω 0 0 1 F [ω]e+ 0 [ω (s)y (s)] ds− 2 n o Rt 0 2 1 Rt 0 “ = ”e− 2 0 [y (s)] ds E F [β]e 0 y (s) dβ(s) I 0 [y Rt 0 [ω 0 (s)]2 ds And so our result is that 1 E {F [β + y ]} = e− 2 Rt 0 [y 0 (s)]2 ds n o Rt 0 E F [β]e 0 y (s) dβ(s) , with y ∈ C0 [0, t] δω Brownian Motion Advanced Topics Local Time Local Time I Spectral Theory: I If V (x) ≥ 0 and V (x) → 0 as |x| → ∞ then the eigenvalue problem 1 00 Ψ (x) − V (x)Ψ(x) = −λΨ(x) 2 1. Has discrete spectrum: λ1 , λ2 , · · · 2. With corresponding eigenfunctions: Ψ1 , Ψ2 , · · · I Theorem (1949): lim t→∞ 1 h − 12 R0t V (β(s)) ds i E e = −λ1 t Note: The expectation can start at any x due to ergodicity I Proof We will first prove this using Feynman-Kac h 1 Rt i u(x, t) = Ex e− 2 0 V (β(s)) ds Brownian Motion Advanced Topics Local Time Local Time I Satisfies the following PDE ut = I 1 uxx − V (x)u, 2 u(x, 0) = 1 By separation of variables we have u(x, t) = ∞ X cj e−λj t ψj (x), where, cj = Z u(x, 0)ψj (y ) dy −∞ j=1 I ∞ But since u(x, 0) = 1 we have that cj = two representations must be equal R∞ −∞ ψj (y ) dy , ∀j ≥ 0, and so the Z ∞ h 1 Rt i X u(x, t) = Ex e− 2 0 V (β(s)) ds = e−λj t ψj (x) j=1 I ∞ ψj (y ) dy −∞ And so the largest eigenvalue, λ1 , controls the behavior i 1 h 1 Rt lim E e− 2 0 V (β(s)) ds = −λ1 t→∞ t Brownian Motion Advanced Topics Local Time Local Time I We also have a variational representation of λ1 Z ∞ Z 1 ∞ 0 λ1 = inf V (y )Ψ2 (y ) dy + [Ψ (y )]2 dy 2 −∞ Ψ∈L2 −∞ ||Ψ||=1 I Which has a corresponding Euler equation 1 00 Ψ (x) − V (x)Ψ(x) = −λΨ(x) 2 I I i h 1 Rt We notice that in the Wiener integral representation, E e− 2 0 V (β(s)) ds , since the internal integral is in an negative exponential, the main contribution comes for paths that remain close to where V (·) is smallest, which leads us to dissect this problem as follows Let β(s), 0 ≤ s < ∞; β(0) = x be BM for t > 0 and consider the proportion of time that β(·) spends in a set A ⊂ R Z 1 t `t (β(·), ·) = χ (β(s)) ds t 0 A Brownian Motion Advanced Topics Local Time Local Time I Some properties of Lt (β(·), ·) with t > 0, x fixed, and β(·) a particular, fixed, path 1. Lt (β(·), ·) is a countable additive, non-negative function 2. Lt (β(·), R) = 1 3. Lt (β(·), ·) : Cx [0, t] → M, the space of probability measures on R I As a set function, Lt (β(·), ·) for fixed x ∈ R and t > 0 and for almost all β(·) has a density function which we call the normalized local time Z 1 t `t (β(·), y ) = δ(β(s) − y ) dy and t 0 Z ∞ Lt (β(·), A) = χA (y )`t (β(·), y ) dy I `t (β(·), ·) → 0 as Table → ∞ for compact A and almost every β(·) I Now consider the following representation h Rt i h i R∞ Ex e− 0 V (β(s)) ds = Ex e−t −∞ V (y )`t (β(·),y ) dy −∞ Brownian Motion Advanced Topics Local Time Local Time I For fixed x ∈ R and t > 0 we define a probability measure on M, Qx,t = PL−1 t , as follows I If C ⊂ M then we can write Qx,t (C) = P {β(·) ∈ Cx [0, ∞] : Lt (β(·), ·) ∈ C} I Lt (β(·), ·) is an occupation measure so we can write i i i h Rt h h R∞ R∞ Ex e− 0 V (β(s)) ds = Ex e−t −∞ V (y )`t (β(·),y ) dy = Ex e−t −∞ V (y ) dLt (β(·),y ) h i h i R∞ R∞ Q Q Ex x,t e−t −∞ V (y ) µ(dy ) = Ex x,t e−t −∞ V (y )f (y ) dy I We define F as the space of probability density functions on R, then this an expected value on F I To understand how the expected value on F behaves as t → ∞, we need to understand how Qx,t and therefore also how Lt (β(·), A) behaves as t → ∞ Brownian Motion Advanced Topics Local Time Local Time I Long time behavior of local time measures 1. Lt (β(·), A) → 0 as t → ∞ for A ⊂ R, compact, and AE β(·) 2. `t (β(·), A) → 0 as t → ∞ for A ⊂ R, compact, and AE β(·) by the ergodic theorem for BM, if β(·) were not BM, then this would converge AE to the invariant measure 3. Qx,t (C) → 0 as t → ∞ if C ⊂ M, C 6= M, i.e. C is a reasonable set I Theorem on Speed of Convergence: We first need to put the Levý topology on F 1. If C ∈ F is closed, then lim sup t→∞ 1 ln Qx,t (C) ≤ inf I(f ) f ∈C t 2. If G ∈ F is open, then lim inf t→∞ 3. Where I(f ) = 1 8 1 ln Qx,t (C) ≥ inf I(f ) f ∈G t Z ∞ −∞ n o [f 0 (y )]2 /f (y ) dy Brownian Motion Advanced Topics Donsker-Varadhan Asymptotics Donsker-Varadhan Asympotics I This is a simple case of what is referred to as “Donsker-Varadhan Asymptotics" and are a large deviation result I An example, suppose f (y ) ∼ N(0, σ 2 ), i.e. f (y ) = y f 0 (y ) = − σ3 √ e 2π I(f ) = 1 8 Z ∞ −∞ n 2 − y 2 2σ and f 0 (y )2 = [f 0 (y )]2 /f (y ) o dy = −2 y2 e σ 6 2π 1 1 8 σ4 Z y2 2σ 2 ∞ −∞ y2 σ − √1 e 2σ2 2π , then and finally we have 2 y2 σ2 1 − y √ e 2σ2 dy = = 8σ 4 8σ 2 σ 2π Note: the last integral is the variance, σ 2 , of a N(0, σ 2 ) random variable I We refer to the functional I : F → [0, ∞] as the entropy, and roughly speaking Qx,t (f ) ∼ e−t inff ∈A I(f ) for “nice" A Brownian Motion Advanced Topics Donsker-Varadhan Asymptotics Donsker-Varadhan Asympotics I I I Now let us apply the “Entropy Asymptotics" with the “Flat Integral" i h 1 Rt h i R∞ Q Ex e− 2 0 V (β(s)) ds = Ex x,t e−t −∞ V (y )f (y ) dy for t large R∞ V (y )f (y ) dy R∞ V (y )f (y ) dy +I(f )] “=” e−t “=” e−t [ −∞ −∞ e−tI(f ) δf δf As t → ∞ we use Laplace asymptotics to get Z ∞ Z h 1 Rt i 1 ∞ [f 0 (y )]2 1 V (y )f (y ) dy + dy lim ln Ex e− 2 0 V (β(s)) ds = − inf t→∞ t f ∈y 8 −∞ f (y ) −∞ p R∞ 2 R∞ Let f (y ) = Ψ(y ), then −∞ Ψ (y ) dy = −∞ f (y ) dy = 1 since f (y ) is a p.d.f., and so Ψ(·) ∈ L2 [−∞, ∞] and ||Ψ|| = 1 Brownian Motion Advanced Topics Donsker-Varadhan Asymptotics Donsker-Varadhan Asympotics I We now transform the “Entropy Asymptotics" expression with some substitutions p R R ∞ ∞ 1. Let f (y ) = Ψ(y ), then −∞ Ψ2 (y ) dy = −∞ f (y ) dy = 1 since f (y ) is a 2 p.d.f., and so Ψ(·) ∈ L [−∞, ∞] and ||Ψ|| = 1 2. Also Ψ0 (y ) = √1 2 f (y ) f 0 (y ), and so [Ψ0 (y )]2 = 1 4 f 0 (y )2 f (y ) I These allow us to write Z ∞ Z 1 ∞ [f 0 (y )]2 − inf V (y )f (y ) dy + dy = f ∈y 8 −∞ f (y ) −∞ Z ∞ Z 1 ∞ 0 [Ψ (y )]2 dy = −λ1 − inf V (y )Ψ2 (y ) dy + 2 −∞ Ψ∈L2 −∞ I Theorem:Let Φ : F → R be bounded and continuous then, by the “general structure theorem" h i h i 1 1 Q lim ln Ex x,t e−tΦ(f ) = lim ln Ex e−tΦ(`t (β(·),·)) = − inf [Φ(f ) + I(f )] t→∞ t t→∞ t f ∈F ||Ψ||=1 Brownian Motion Advanced Topics Donsker-Varadhan Asymptotics Donsker-Varadhan Asympotics I This is more subtle than action asymptotics, for example consider h i 1 Q lim ln Ex x,t e+tΦ(f ) = sup [Φ(f ) − I(f )] t→∞ t f ∈F 1. There is always a fight between the two terms in the supremum 2. In statistical mechanics we often consider αΦ(f ) and want to compute supf ∈F [αΦ(f ) − I(f )] = g(α), where α is a convex function of α 3. Them may be a critical value of α, call it α0 , where there is a phase transition, this is due to nonuniqueness in the f that maximized the functional Brownian Motion Advanced Topics Donsker-Varadhan Asymptotics An Example Using Action and Entropy Asymptotics I Now we will use “Entropy Asymptotics" to revisit a topic we have already considered I Recall that ( P ) sup β(s) ≤ α r = 0≤s≤t h E e sup0≤s≤t β(s) i 2 πt α Z 0 ∞ Z e dP{ sup β(s) ≤ α} = = h(t) = eα 0 2 e πt 2 − α2t r dα = with the substitution u = I 2 πt e 0≤s≤t ∞ Z e− (α−t)2 2t t e+ 2 dα = 0 2 2t e π Then we have r 2 2t e π Z ∞ √ t α 0 r α−t √ t 1 lim ln h(t) = t→∞ t r ∞ Z α 0 r ∞ Z u2 e− 2t du, so that we also have e− u2 2 du = 1 2 Z ∞ √ t 2 − α2t2 e dα πt e− u2 2 du Brownian Motion Advanced Topics Donsker-Varadhan Asymptotics An Example Using Action and Entropy Asymptotics I First we turn the t → ∞ limit into an → 0 limit i h i h 1 √ 1 lim ln E esup0≤s≤t β(s) = lim ln E e sup0≤τ ≤1 β(τ ) t→∞ t →0 Recall that by Action Asymptotics we have " # Z h 1 i √ 1 1 0 sup0≤τ ≤1 β(τ ) 2 lim ln E e [ω (τ )] dτ = sup sup ω(τ ) − →0 2 0 ω∈C0∗ [0,1] 0≤τ ≤1 a2 1 = max a − = a>0 2 2 I The supremum comes on straight lines, that minimize arc-length i.e. the second term, so consider ω(τ ) = aτ , and a = 1 is the maximizer I Brownian Motion Advanced Topics Donsker-Varadhan Asymptotics An Example Using Action and Entropy Asymptotics I Now we solve the same problem using Entropy Asymptotics by using a result of Paul Levý that the following have the same probability distributions ( ) P sup β(s) ≤ α = P {t`t (β(·), 0)} 0≤s≤t I Thus we have that h i h i h i h(t) = E esup0≤s≤t β(s) = E et`t (β(·),0) = E etΦ[`t (β(·),0)] , where Φ[f ] = f (0) I So from Entropy Asymptotics we get lim t→∞ I Z i 1 ∞ [f 0 (y )]2 1 1 h ln h(t) = lim E etΦ[`t (β(·),0)] = sup f (0) − dy t→∞ t t 8 −∞ f (y ) f ∈F Recall that f ∈ F is a probability distribution, and so the maximizing family of functions (proven below) is fa (y ) = ae−2a|y | Brownian Motion Advanced Topics Donsker-Varadhan Asymptotics An Example Using Action and Entropy Asymptotics I Recall that f ∈ F is a probability distribution, and so the maximizing family of functions (proven below) is fa (y ) = ae−2a|y | I We can write fa (y ) = ae−2a|y | = ( ae−2ay ae2ay [fa0 (y )]2 = I y ≥0 , so fa0 (y ) = y <0 ( 4a4 e−4ay 4a4 e4ay ( −2a2 e−2ay 2a2 e2ay y ≥0 , and so y <0 y ≥0 = 4a4 e−4a|y | y <0 This gives us Z Z 1 ∞ [f 0 (y )]2 1 ∞ sup f (0) − dy = sup a − 4a3 e−2a|y | dy 8 −∞ f (y ) 8 −∞ a>0 a>0 Z a2 ∞ 1 a2 = sup a − ae−2a|y | dy = sup a − = , which occurs at a = 1 2 −∞ 2 2 a>0 a>0 Brownian Motion Advanced Topics Donsker-Varadhan Asymptotics An Example Using Action and Entropy Asymptotics I Now we find the maximizing family of functions by the same transformation as before p 1. f (y ) = Ψ(y ) or f (y ) = Ψ2 (y ), and so 2 2. f (0) = Ψ (0) 3. I 1 4 f 0 (y )2 f (y ) = [Ψ0 (y )]2 And so we obtain Z Z 1 ∞ [f 0 (y )]2 1 ∞ 0 dy = sup Ψ2 (0) − [Ψ (y )]2 dy sup f (0) − 8 −∞ f (y ) 2 −∞ f ∈F Ψ∈L2 ||Ψ||=1 I Let Ψ(0) = a we get the following constrained Euler-Lagrange equation Ψ00 (y ) − 2λΨ(y ), I Ψ(0) = a This is maximized with a stretched version of Ψ(y ) = e−2|y | Brownian Motion Advanced Topics Can One Hear the Shape of a Drum? Kac’s Drum I I Let Ω ⊂ R2 be an open domain with sufficiently smooth boundary, ∂Ω, so that the following problem has a unique solution 1 ∆u + λu = 0, with u = 0 on ∂Ω 2 Under these circumstances we know that 1. ∃0 < λ1 < λ2 < · · · a discrete spectrum 2. ∃u1 (x, y ) < u1 (x, y ) < · · · corresponding normalized eigenfunctions I Consider C(λ) = X 1 = # of eigenvalues < λ λj <λ I C(λ) is an increasing function in λ, and Hermen Weyl proved that I |Ω|λ as λ → ∞ 2π Additionally, Carlemann proved that X λ u(x, y ) ∼ , ∀ (x, y ) ∈ Ω as λ → ∞ 2π C(λ) ∼ λj <λ Brownian Motion Advanced Topics Can One Hear the Shape of a Drum? Kac’s Drum I I I Now consider starting a BM at (x0 , y0 ) ∈ Ω Let p(x0 , y0 , x, y , t) be the probability density function of a 2D BM starting at (x0 , y0 ) reaching (x, y ) at time t without hitting ∂Ω Einstein-Smoluchowski: Then p(x0 , y0 , x, y , t) is the solution to 1 ∂p = ∆p in Ω ∂t 2 p = 0 on ∂Ω, ∀t > 0 I We note that as t → 0 Z g(x, y )p(x0 , y0 , x, y , t) dx dy → g(x0 , y0 ) I Assume we can find p using separation of variables: p(x0 , y0 , x, y , t) = T (t)U(x, y ), then Ω T ∆U, U = 0 on ∂Ω, ∀t > 0 2 0 ∆U T = = −λ yields T 2 −λt T (t) = e , and U = the eigenfunction corresponding to λ T 0U = Brownian Motion Advanced Topics Can One Hear the Shape of a Drum? Kac’s Drum I So this means that we can write explicitly p(x0 , y0 , x, y , t) = ∞ X e−λj t uj (x0 , y0 )uj (x, y ), and so we know j=1 p(x0 , y0 , x0 , y0 , t) = ∞ X e−λj t uj2 (x0 , y0 ) j=1 I Let p∗ (x0 , y0 , x, y , t) be the probability density function of unrestricted 2D BM starting at (x0 , y0 ) reaching (x, y ) at time t p∗ (x0 , y0 , x, y , t) = I 1 − (x−x2t 0 )2 − (y −y2t 0 )2 e 2πt Thus we conclude that ∞ X j=1 e−λj t uj2 (x0 , y0 ) ∼ p∗ (x0 , y0 , x, y , t) ∼ 1 as t → 0 2πt Brownian Motion Advanced Topics Can One Hear the Shape of a Drum? Kac’s Drum I Karamata Tauberian Theorem: Consider Z ∞ f (t) = e−λt dα(λ), and assume 0 1. The above Laplace-Stiltje’s transform exists 2. α(λ) is non-decreasing on (0, ∞) I If f (t) ∼ At −γ as t → 0 for A and γ constants then α(λ) ∼ I Aλγ as λ → ∞(λ → 0) Γ(γ + 1) We now apply the Karamata Tauberian Theorem to Z ∞ ∞ X X 2 f (t) = e−λt dα(λ) = e−λj t uj2 (x0 , y0 ), where α(λ) = uj (x0 , y0 ) 0 λj <λ j=1 1 2πt as t → 0, and so α(λ) ∼ λ 2π I We know f (t) ∼ I By integrating this over Ω we get Weyl’s theorem as λ → ∞ Brownian Motion Advanced Topics Probabilistic Potential Theory Probabilistic Potential Theory 1. Let Ω ∈ R3 be a bounded closed domain 2. Let r(t) ∈ C be a continuous function starting at the origin 3. Let χΩ (·) be the indicator function of Ω I Consider the following functional on C Z ∞ TΩ (y, r(·)) = χΩ (y + r(τ )) dτ, y ∈ R3 0 I I This functional is the total occupations time of r(·), a 3D BM, in Ω translated by y Now impose Wiener measure on C and consider the following Wiener integral Z ∞ E {TΩ (y, r(·))} = P {y + r(τ ) ∈ Ω} dτ 0 I Note that because we are using Wiener measure we know Z ∞ |r−y|2 1 P {y + r(τ ) ∈ Ω} = e− 2τ dr 3/2 (2πτ ) 0 Brownian Motion Advanced Topics Probabilistic Potential Theory Probabilistic Potential Theory I We now use Fubini’s theorem to exchange the order of integration Z Z ∞ |r−y|2 1 e− 2τ dτ E {TΩ (y, r(·))} = dr 3/2 (2πτ ) Ω Z 0 1 dr = < ∞ in R3 2π Ω |r − y| I We see that in R3 AE BM path starting at y spends a finite amount of time in Ω I Now consider the k th moment of the occupation time Z Z n o [k ] drk k! dr1 dr2 E TΩk (y, r(·)) = · · · ··· (2π)k Ω |r − y| |r − r | |r − rk −1 | 1 2 1 k Ω I k = 1, 2, · · · We focus on the second moment, k = 2 n o Z ∞Z ∞ 2 E TΩ (y, r(·)) = P {y + r(τ1 ) ∈ Ω} P {y + r(τ2 ) ∈ Ω} dτ1 dτ2 0 0 Brownian Motion Advanced Topics Probabilistic Potential Theory Probabilistic Potential Theory I We focus on the second moment, k = 2 n o Z ∞Z ∞ P {y + r(τ1 ) ∈ Ω} P {y + r(τ2 ) ∈ Ω} dτ1 dτ2 E TΩ2 (y, r(·)) = 0 ZZ =2 0≤τ1 <τ2 <∞ 0 dτ1 dτ2 Ω = I 2 Z Z Ω |r −r | |r1 −y|2 1 1 − 2 1 e− 2τ e 2(τ2 −τ1 ) dr1 dr2 3/2 3/2 (2πτ1 ) [2π(τ2 − τ1 )] 2 (2π)2 Z Z Ω Ω dr1 dr2 |r1 − y| |r2 − r1 | The formula for the k th moment suggests that we should consider the following eigenvalue problem Z φ(ρ) 1 dρ = λφ(r), r ∈ Ω 2π Ω |r − ρ| Brownian Motion Advanced Topics Probabilistic Potential Theory Probabilistic Potential Theory I The integral kernel in the eigenvalue problem is Hilbert-Schmidt 1. Since the single integral is convergent, we have Z Z 1 dr dρ < ∞ 2 Ω Ω |r − ρ| 2. We also need to show that the kernel is positive definite: Z Z φ(r)φ(ρ) dr dρ > 0 ∀φ(ρ) 6= 0 in L2 (Ω) Ω Ω |r − ρ| Note that: ∞ Z 0 ∞ |r−y|2 1 e− 2τ dτ = 3/2 (2πτ ) 0 Z −|ζ|2 τ 1 τ 3/2 dτ dζ = eiζ·(r−ρ) e 2 3/2 3/2 (2πτ ) (2π) R3 Z ∞ Z −|ζ|2 τ 1 dτ dζeiζ·(r−ρ) e 2 3 (2π) 0 R3 1 1 = 2π |r − ρ| Z Brownian Motion Advanced Topics Probabilistic Potential Theory Probabilistic Potential Theory I I So Z Z φ(r)φ(ρ) dr dρ = Ω Ω |r − ρ| 2 Z −|ζ|2 τ dζe 2 φ(ρ)eiζ·ρ dρ > 0, ∀φ(ρ) 6= 0 in L2 (Ω) Z ∞ Z 1 dτ (2π)3 0 Ω R3 With the kernel being Hilbert-Schmidt, we know that the integral equation has 1. Discrete spectrum: λ1 , λ2 , · · · 2. With corresponding eigenfunctions that form a complete, orthonormal basis for L2 (Ω) I Lemma: ∞ o X 1 n k λkj −1 E TΩ (y, r(·)) = k! Z j=1 φj (r) dr Ω 1 2π 1. This holds for all y ∈ R3 2. If y ∈ Ω, then we note that 1 2π φj (ρ) Z Ω |ρ − y| dρ = λj φj (y) Z Ω φj (ρ) dρ |ρ − y| Brownian Motion Advanced Topics Probabilistic Potential Theory Probabilistic Potential Theory I Proof: Recall that Z Z o [k ] 1 n k 1 dr1 dr2 drk E TΩ (y, r(·)) = · · · ··· k! (2π)k Ω |rk − rk −1 | Ω |r1 − y| |r2 − r1 | I We recognize this as an iterated integral equation of the form a(y, r1 )a(r1 , r2 ) · · · a(rk −1 , rk ) I We can then rewrite this using Mercer’s theorem representation of the kernel of the integral operator ∞ X 1 = λj φj (ρ)φj (y) |ρ − y| j=1 I Next we apply Mercer’s theorem only to the terms not involving y to get Z Z X ∞ o 1 1 n k 1 E TΩ (y, r(·)) = λkj −1 φj (r1 )φj (rk ) dr1 drk k! 2π Ω |r1 − y| Ω j=1 Brownian Motion Advanced Topics Probabilistic Potential Theory Probabilistic Potential Theory I To review we have that o 1 n k E TΩ (y, r(·)) = k! I (P ∞ R R 1 λkj −1 Ω φj (r) dr 2π Ω R k λ φ (r)φ (y) dr, j j j j=1 Ω j=1 P∞ φj (ρ) |ρ−y| dρ, y ∈ R3 y∈Ω Now let us consider the moment generation function (Laplace transform) with z ∈ C ∞ n o X o zk n k E ezTΩ (y,r(·)) = E TΩ (y, r(·)) k! k =0 I Now we use the above lemma to get Z Z ∞ φj (ρ) z X 1 dρ =1+ φj (r) dr 2π 1 − λj z Ω Ω |ρ − y| j=1 1. This series converges if |z| < λ 1 max 2. The moment generating function is analytic if <{z} < 0 since TΩ ≥ 0 3. The last series is analytic for <{z} < 0, so by analytic continuation this identity holds with <{z} < 0 Brownian Motion Advanced Topics Probabilistic Potential Theory Probabilistic Potential Theory I Let u > 0 and define Z Z ∞ n o φj (ρ) u X 1 h(y, u) = E e−uTΩ (y,r(·)) = 1− φj (r) dr dρ 2π 1 + λj u Ω Ω |ρ − y| j=1 (*) I This series converges on compact sets in C because 1. 1 <1 1 + λj u 2. ∞ Z X φj (r) dr j=1 φj (ρ) Z Ω Ω |ρ − y| 2 dρ ≤ j=1 Z |Ω| Ω I ∞ Z X 2 X ∞ Z φj (r) dr Ω j=1 φj (ρ) Ω |ρ − y| 2 dρ = dρ <∞ |ρ − y| This gives uniform convergence via the Weierstrass M-test and thus this is also analytic Brownian Motion Advanced Topics Probabilistic Potential Theory Probabilistic Potential Theory I If y ∈ Ω then we get h(y, u) = 1 − ∞ X j=1 I And so we can multiply both sides by 1 2π I λj u 1 + λj u Z Ω h(y, u) dy 1 = |y − r| 2π Z Ω Z 1 2π|y−r| φj (r) dr φj (y) Ω and integrate over Ω Z Z ∞ X λj u φj (y) dy dy 1 − φj (ρ) dρ |y − r| 1 + λj u 2π Ω |y − r| Ω j=1 But we know that Z Z ∞ Z X φj (y) dy 1 dy 1 = φj (ρ) dρ 2π Ω |y − r| 2π Ω |y − r| Ω j=1 I Thus we cane write that Z Z Z ∞ X φj (y) dy h(y, u) dy 1 1 1 = φj (ρ) dρ 2π Ω |y − r| 1 + λj u 2π Ω Ω |y − r| j=1 Brownian Motion Advanced Topics Probabilistic Potential Theory Probabilistic Potential Theory I We recognize the left hand side of the previous equation from (*), and so we use this ro rewrite this as Z h(y, u) dy 1 1 = (1 − h(r, u)) , ∀r ∈ R3 2π Ω |y − r| u I Moreover, if we rename variables we get Z h(ρ, u) dρ 1 1 = (1 − h(y, u)) , 2π Ω |y − ρ| u I ∀y ∈ R3 (**) We now make some important observations 1. From (*) we see that if y ∈ / Ω then h(y, u) is harmonic in y, and the series in (*) converges uniformly on compact Ω’s 2. Again from (*) we get 2 1/2 X 2 1/2 ∞ Z ∞ Z φj (ρ) u X h(y, u) > 1 − dρ φj (ρ) dρ 2π Ω Ω |ρ − y| j=1 >1− j=1 u |Ω|1/2 2π Z Ω dρ |ρ − y| 1/2 Brownian Motion Advanced Topics Probabilistic Potential Theory Probabilistic Potential Theory 3. So we now know that 0 ≤ h(y, u) ≤ 1, and so lim h(y, u) = 1 (***) u%∞ 4. And for from Courant-Hilbert II, pp. 245–246 Z h(y, u) dy ∆ = −4πh(y, u) |y − r| Ω I Now apply the Laplacian to both sides of (**) to get 1 −2h(y, u) = − ∆h(y, u) u or we get I 1 ∆h(y, u) − uh(y, u) = 0, y ∈ Ω 2 Now consider U(y) = limu%∞ (1 − h(y, u)) = P {TΩ (y, r(·)) > 0}, this is the capacitory potential (capacitance) and follows easily from the definition of the moment generating function Brownian Motion Advanced Topics Probabilistic Potential Theory Probabilistic Potential Theory I Example: Let Ω be a sphere of radius 1 centered at the origin 1. h(y, u) is clearly spherically symmetric 2. h(y, u) is harmonic outside Ω, so we have h(y, u) = α(u) + β(u), |y| y∈ /Ω α(u) 3. From (***) we see that β(u) = 1 and so h(y, u) = |y| + 1 for y ∈ Ω 4. We also know that for y ∈ Ω we have √ sinh( 2u |y|) h(y, u) = γ(u) |y| 5. If we substitute this into the equation (**) we get that γ(u) = 1 √ √ 2u cosh(a 2u) 6. h(y, u) is continuous ∀y so from the uniform convergence of the series, and so √ α(u) 1 sinh( 2ua) 1 √ +1= √ a 2u cosh( 2ua) a to finally give us √ 2u) 1 − 1 1 − tanh(a √ , y∈ /Ω |y| √ a 2u h(y, u) = √ sinh( 2u|y|) √ , y∈Ω 2u cosh( 2ua)|y| Brownian Motion Advanced Topics Probabilistic Potential Theory Probabilistic Potential Theory I Recall that ( U(y) = lim (1 − h(y, u)) = P TS(0,a) (y, r(·)) > 0 = u%∞ a , |y| y∈ /Ω 1, y∈Ω I This is the capacitory potential of S(0, a) I Now back to the general case, ∀y ∈ R3 we have !Z Z ∞ n o X φj (ρ) dρ 1 1 −uTΩ (y,r(·)) φj (r) dr 1−E e = 1 2π λ + j Ω Ω |ρ − y| u j=1 1. We note that 0 ≤ 1 − h(y, u) ≤ 1 2. The function 1 − h(y, u) is non-decreasing in u: 1 − h(y, u1 ) ≤ 1 − h(y, u2 ) if u1 < u2 3. This is true due to the following 3.1 0 ≤ e−uTΩ (y,r(·)) ≤ 1 and 3.2 lim e u%∞ −uTΩ (y,r(·)) = ( 0, TΩ > 0 1, TΩ = 0 Brownian Motion Advanced Topics Probabilistic Potential Theory Probabilistic Potential Theory I From the previous results and the bounded convergence theorem we have U(y) = lim (1 − h(y, u)) = P {TΩ (y, r(·)) > 0} u%∞ and hence also U(y) = lim ∞ X u%∞ j=1 1 u 1 + λj !Z φj (r) dr Ω 1 2π Z Ω φj (ρ) dρ |ρ − y| and this holds ∀y ∈ R3 Case 1. Let y ∈ Ωo (the interior), clearly the continuity of r(·) immediately implies U (y) = P {TΩ (y, r(·)) > 0} = 1 Remark: with y ∈ summability result Ωo 1 = lim u%∞ we have U (y) = 1 and so we have the following ∞ X j=1 !Z λj λj + 1 u φj (r) dr Ω 1 2π φj (ρ) dρ Z Ω |ρ − y| Brownian Motion Advanced Topics Probabilistic Potential Theory Probabilistic Potential Theory Case 2. Let y ∈ / Ω, we already know that 1 − h(y, u) is harmonic in y, and it is nondecreasing in u, and the previous limit in u exists and equals P {TΩ (y, r(·)) > 0}, thus by Harnack’s theorem, U(y) is harmonic with y∈ / Ω. Assume that Ω ⊂ S(0, a), then P {TΩ (y, r(·)) > 0} ≤ P TS(0,a) (y, r(·)) > 0 From the last problem this means P {TΩ (y, r(·)) > 0} ≤ a , |y| y∈ / S(0, a) and so lim|y|→∞ U(y) = 0 Case 3. Let yo ∈ ∂Ω, and assume that it is regular in the sense of Poincaré: ∃ a sphere S(y∗ , ) lying completely in Ω so that yo ∈ S(y∗ , ) Consider now y∈ /Ω U(y) = P {TΩ (y, r(·)) > 0} ≥ P TS(0,a) (y, r(·)) > 0 = |y − y∗ | As y → yo with y ∈ / Ω we have |y−y → |y −y , and since U(y) ≤ 1 we ∗| o ∗| have finally that lim U(y) = 1 y→yo Brownian Motion Advanced Topics Probabilistic Potential Theory Probabilistic Potential Theory I Thus if Ω is a closed and bounded region, each point on the boundary that is regular in the Poincaré sense has U(y) as the capacitory potential of Ω I Recall that U(y) = lim δ→0 I ∞ X j=1 1 λj + δ Z Ω Z Ω φj (ρ) dρ |ρ − y| We note that this implies that lim |y|(1 − h(|y|, u)) = |y|→∞ I 1 2π φj (r) dr 1 2π Z uh(ρ, u) dρ Ω Again assume that Ω ∈ S(0, a), then h(y, u) = E e−uTΩ ≥ e−uTS(0,a) , a a there for y ∈ / S(0, a) we have h(y, u) ≥ 1 − |y| or 1 − h(y, u) ≤ |y| and so u 2π Z h(ρ, u) dρ ≤ a Ω Brownian Motion Advanced Topics Probabilistic Potential Theory c Michael Mascagni, 2016