Lecture 7. Further topics

In this lecture let X be a locally compact and complete metric space equipped with a Radon measure μ supported on X. Let (ℰ, ℱ = dom(ℰ)) be a Dirichlet form on X. We assume throughout that the heat semigroup Pt is stochastically complete, i.e., Pt1 = 1 for every t ≥ 0.

Contents

Regular Dirichlet forms, Energy measures

We denote by Cc(X) the vector space of all continuous functions with compact support in X and C0(X) its closure with respect to the supremum norm.

A core for (X,μ,ℰ,ℱ) is a subset 𝒞 of Cc(X) ∩ ℱ which is dense in Cc(X) in the supremum norm and dense in in the norm

(||f||^2_{L^2(X,\mu)} + \mathcal{E}(f,f))^{1/2}.

Definition. The Dirichlet form is called regular if it admits a core.

Recall that for any f,g ∈ ℱ, we have

\mathcal{E}(f,g)=\lim_{t\to 0}\frac1t\langle (I-P_t)f,g\rangle=\lim_{t\to 0}\frac1{t}\int_X \int_X(f(x)-f(y)) g(x) p_t(x,dy) d\mu(x),

where pt(x,·) are the heat kernel measures associated to the Dirichlet form (ℰ, ℱ).
From the symmetry property \eqref{heat kernel measure symmetry} of the heat kernel measure one also has

\mathcal{E}(f, g)=\lim_{t\to 0}\frac1{2t}\int_X \int_X(f(x)-f(y))(g(x)-g(y)) p_t(x,dy) d\mu(x).

Lemma. For f,g ∈ ℱ ∩ L(X,μ), fg ∈ ℱ and

\mathcal{E}(fg)^{1/2} \le ||f||_\infty \mathcal{E}(g)^{1/2} +||g||_\infty \mathcal{E}(f)^{1/2}.

Proof. For f, g ∈ ℱ ∩ L(X,μ) such that fg ∈ ℱ,

\mathcal{E}(fg)=\lim_{t\to 0}\frac1{2t}\int_X \int_X(f(x)g(x)-f(y)g(y))^2 p_t(x,dy) d\mu(x).

Write f(x)g(x) – f(y)g(y) = f(x)(g(x) – g(y)) + g(y)(f(x) – f(y)), then by Minkowski’s inequality

\left(\int_X \int_X(f(x)g(x)-f(y)g(y))^2 p_t(x,dy) d\mu(x)\right)^{1/2}

\le ||f||_{\infty}\left(\int_X \int_X (g(x)-g(y))^2 p_t(x,dy) d\mu(x)\right)^{1/2}

+||g||_{\infty}\left(\int_X \int_X(f(x)-f(y))^2 p_t(x,dy) d\mu(x)\right)^{1/2}.

We conclude the expected inequality by multiplying by 1/√2t and taking the limit t → 0 for both sides above.

Theorem (Energy measures) Assume that is regular. For f ∈ ℱ ∩ L(X,μ), there exists a unique Radon measure on X, denoted by dΓ(f), so that for every ϕ ∈ ℱ ∩ Cc(X),

\int_X\phi\, d\Gamma(f)=\frac{1}{2}[2\mathcal{E}(\phi f,f)-\mathcal{E}(\phi, f^2)]

= \lim_{t \to 0} \frac{1}{2t} \int_X \int_X \phi(x) (f(x)-f(y))^2 p_t(x,dy) d\mu(x).

The Radon measure dΓ(f) is called the energy measure of f (and is therefore the weak * limit of (1/(2t)) ∫X (f(x) – f(y))2 pt(x,dy)).

Proof. Let f ∈ ℱ ∩ L(X,μ). For any ϕ ∈ ℱ ∩ Cc(X),

\frac1{2t} \int_X \int_X \phi(x) (f(x)-f(y))^2 p_t(x,dy) d\mu(x)=-\frac1{2t} \langle (I-P_t)f^2 ,\phi\rangle+\frac1{t} \langle (I-P_t)f ,f\phi\rangle.

Letting t → 0, the right-hand side converges to

\frac{1}{2}[2\mathcal{E}(\phi f,f)-\mathcal{E}(\phi, f^2)] .

On the other hand, observing that

\frac1{2t} \int_X \int_X |\phi(x)| (f(x)-f(y))^2 p_t(x,dy) d\mu(x) \le \|\phi\|_{\infty} \frac1{2t} \int_X \int_X (f(x)-f(y))^2 p_t(x,dy) d\mu(x),

we deduce

\left|\frac{1}{2}[2\mathcal{E}(\phi f,f)-\mathcal{E}(\phi, f^2)]\right|\le \|\phi\|_{\infty}\mathcal E(f).

Therefore we conclude the proof by applying the Riesz-Markov representation theorem.

One can actually define dΓ(f,f) for every f ∈ ℱ using the following lemmas.

Lemma. Let f ∈ ℱ. Then fn = min(n, max(-n, f)) ∈ ℱ and ℰ(f – fn) → 0.

Proof. Let f ∈ ℱ. For every x, y ∈ X, we have

|f_n(x) - f_n(y)| \leq |f(x) - f(y)|

and |fn(x)| ≤ |f(x)|. So fn is a normal contraction of f and ℰ(fn) ≤ ℰ(f).

Recall that

\mathcal{E}(f-f_n)=\lim_{t\to 0}\frac1{2t} \int_X \int_X(f(x)-f_n(x)-f(y)+f_n(y))^2 p_t(x,dy) d\mu(x).

Expanding the square inside the integral gives that

\mathcal{E}(f-f_n)=\mathcal E(f)+\mathcal E(f_n)-2\mathcal E(f_n,f)\le 2\mathcal E(f)-2\mathcal E(f_n,f).

Letting n → ∞, we have ℰ(fn,f) → ℰ(f). Therefore ℰ(f – fn) converges to 0 as n → ∞.

Lemma. For f,g ∈ ℱ ∩ L(X,μ) and nonnegative ϕ ∈ ℱ ∩ Cc(X),

\left| \sqrt{\int_X\phi\, d\Gamma(f)} -\sqrt{\int_X\phi\, d\Gamma(g) } \right|^2 \le \int_X\phi\, d\Gamma(f-g) \le ||\phi||_{L^\infty(X,\mu)}\mathcal{E}(f-g)

Proof. The second inequality follows from the proof in Theorem 16.3. For the first inequality, it suffice to show that for any f, g ∈ ℱ ∩ L(X,μ) and any nonnegative ϕ ∈ ℱ ∩ Cc(X),

\sqrt{\int_X\phi\, d\Gamma(f)} \le \sqrt{\int_X\phi\, d\Gamma(g) } +\sqrt{ \int_X\phi\, d\Gamma(f-g) }.

Indeed, this inequality follows from a similar proof as in Lemma 16.2 by noting that f = f – g + g and using Minkowski’s inequality.

Thanks to the previous lemmas, by approximation, one can define dΓ(f) for every f ∈ ℱ by

dΓ(f) = sup{dΓ(fn) : fn = min(n, max(-n, f)), n = 1, 2, …}.

For f,g ∈ ℱ, one can define dΓ(f,g) by polarization

dΓ(f,g) = (1/4)(dΓ(f + g) – dΓ(f – g)).

The following representation theorem for regular Dirichlet forms then holds.

Theorem  (Beurling-Deny) Assume that is regular. For u,v ∈ ℱ, ℰ(u,v) = ∫X dΓ(u,v).

Hunt process associated with a regular Dirichlet form

Definition. A Hunt process with state space X is a family of stochastic process (Xt)t≥0 and probability measures (ℙx)x∈X defined on a measure space (Ω, ℱ), such that (Xt)t≥0 is adapted w.r.t. the right-continuous minimal completed admissible filtration (ℱt)t≥0, X0 = x, x-a.s. and the following hold:

  • (i) x → ℙx(Xt ∈ B) is measurable for all t > 0 and B ∈ ℬ(X),
  • (ii) X is a strong Markov process, i.e. for every stopping time T, XT is T-measurable and for every B ∈ ℬ(X)

    \mathbb{P}_x(X_{T+t} \in B | \mathcal{F}_T) = \mathbb{P}_{X_T}(X_t \in B) \quad \mathbb{P}_x\text{-a.s. on } \{T < \infty\},

  • (iii) X is right-continuous, i.e.

    \lim_{s \downarrow t} X_s = X_t, \quad \forall t \quad \mathbb{P}_x\text{-a.s.}

  • (iv) X is quasi left-continuous, i.e. for all stopping times T and (Tn)n such that Tn ↑ T a.s.

    \lim_{n \to \infty} X_{T_n} = X_T, \quad \mathbb{P}_x\text{-a.s. on } \{T < \infty\}.

Remark 

  1. Note that quasi left-continuity does not necessarily imply left-continuity, because the set

    A = \left\{ \lim_{n \to \infty} X_{s_n} = X_t \right\}

    might depend on the choice of sequence (sn)n, sn ↑ t.

  2. In more generality, one might consider situations where (ℙx(Xt ∈ ·))x,t are sub-probability measures (i.e. all the axioms of probability measures are satisfied but x(Xt ∈ Ω) ≤ 1). In that case we can perform a one-point compactification of X by introducing a cemetery state ∂ ∉ X and redefine x to be a probability measure on X ∪ {∂}.

The following theorem can then be proved, see the book [FOT], Theorem 4.2.1.

Theorem (Fukushima) Assume that is regular, then there exists a Hunt process ((Xt)t≥0, (ℙx)x∈X) such that for μ-a.e. x ∈ X, A ∈ ℬ(X) and t ≥ 0,

\mathbb{P}_x (X_t \in A)=p_t(x,A)

where pt(x, ·) are the heat kernel measures associated to the Dirichlet form (ℰ, 𝒟(ℰ)).

Intrinsic metric

Definition. The Dirichlet form is called strongly local if for any two functions f,g ∈ ℱ with compact supports such that f is constant in a neighborhood of the support of g, we have ℰ(f,g) = 0.

With respect to we can define the following intrinsic metric d on X by

d_{\mathcal{E}}(x,y)=\sup\{u(x)-u(y)\, :\, u\in\mathcal{F}\cap C_0(X)\text{ and } d\Gamma(u,u)\le d\mu\}.

Here the condition dΓ(u,u) ≤ dμ means that Γ(u,u) is absolutely continuous with respect to μ with Radon-Nikodym derivative bounded by 1.

The term “intrinsic metric” is potentially misleading because in general there is no reason why d is a metric on X (it could be infinite for a given pair of points x,y or zero for some distinct pair of points).

Definition. A strongly local regular Dirichlet space is called strictly local if d is a metric on X and the topology induced by d coincides with the topology on X.

Example (Riemannian manifolds) Let (M,g) be a complete n-dimensional Riemannian manifold with Riemannian volume measure μ. We consider the standard Dirichlet form on M, which is obtained by closing the bilinear form

\mathcal{E}(f,g)=\int_\mathbb{M} \langle \nabla f ,\nabla g \rangle d\mu, \quad f,g \in C_0^\infty(\mathbb M).

Then is a strictly local Dirichlet form such that

d_\mathcal{E} (x,y)=d_g(x,y).

Example (Carnot groups) Let G be a Carnot group with sub-Laplacian

L=\sum_{i=1}^d V_i^2

and Dirichlet form

\mathcal{E}(f)=\int_{\mathbb G} \sum_{i=1} (V_if)^2 d\mu.

Then is a strictly local Dirichlet form such that

d_\mathcal{E} (x,y)=d_{CC} (x,y)

where dCC is the so-called Carnot-Carathéodory distance which is defined as follows.

An absolutely continuous curve γ : [0,T] → G is said to be subunit for the operator L if for every smooth function f : G → ℝ we have

\left| \frac{d}{dt} f ( \gamma(t) ) \right| \le \sqrt{ (\Gamma f) (\gamma(t)) }.

We then define the subunit length of γ as s(γ) = T.

Given x,y ∈ G, we indicate with

S(x,y) = {γ : [0,T] → G | γ is subunit for Γ, γ(0) = x, γ(T) = y}.

It is a consequence of the Chow-Rashevskii theorem that

S(x,y) ≠ ∅, for every x, y ∈ G.

One defines then

d_{CC}(x,y) = \inf\{ \ell_s(\gamma) \mid  \gamma \in  S(x,y)\}

Example Consider on the Sierpinski gasket the standard Dirichlet form . Then is regular, but unless f is constant, for f ∈ ℱ, dΓ(f) is singular with respect to the Hausdorff measure μ, see [KM19]. As a consequence is not strictly local.

 Further reading

Far from being exhaustive we mention the following references for further reading:

  1. The book [FOT] is a standard comprehensive reference in the theory of Dirichlet forms and associated Hunt processes, see also [ChenFukushima].
  2. The book [GSC] shows how one can restrict Dirichlet forms to domains (abstract Dirichlet and Neumann boundary conditions).
  3. Parabolic regularity theory for the heat equation can be developed in the setting of abstract strictly local Dirichlet spaces, see [Saloff].
This entry was posted in Dirichlet forms at NYU and tagged , , . Bookmark the permalink.

Leave a comment