This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Brownian windings, Stochastic Green’s formula and inhomogeneous magnetic impurities

Isao Sauzedde University of Warwick [email protected]
Abstract.

We give a general Green formula for the planar Brownian motion, which we apply to study the Aharonov–Bohm effect induced by Poisson distributed magnetic impurities on a Brownian electron in the presence of an inhomogeneous magnetic field.

2020 Mathematics Subject Classification:
Primary 60J65; 60K37 Secondary 60G17

1. Introduction

1.1. Stochastic Green’s formula

For a smooth loop X=(X1,X2):[0,T]2X=(X^{1},X^{2}):[0,T]\to\mathbb{R}^{2} and a point zz outside the range of XX, let 𝐧X(z){\mathbf{n}}_{X}(z)\in\mathbb{Z} be the winding index of XX around zz. For any smooth differential 11-form η=η1dx1+η2dx2\eta=\eta_{1}{\mathop{}\!\mathrm{d}}x^{1}+\eta_{2}{\mathop{}\!\mathrm{d}}x^{2}, the Green formula states that

Xη=2𝐧Xdη,\int_{X}\eta=\int_{\mathbb{R}^{2}}{\mathbf{n}}_{X}{\mathop{}\!\mathrm{d}}\eta,

where dη{\mathop{}\!\mathrm{d}}\eta is the exterior derivative of η\eta. In other words, for two smooth functions η1,η2:2\eta_{1},\eta_{2}:\mathbb{R}^{2}\to\mathbb{R},

0Tη1(Xt)dXt1+0Tη2(Xt)dXt2=2𝐧X(z)(1η2(z)2η1(z))dz.\int_{0}^{T}\eta_{1}(X_{t}){\mathop{}\!\mathrm{d}}X^{1}_{t}+\int_{0}^{T}\eta_{2}(X_{t}){\mathop{}\!\mathrm{d}}X^{2}_{t}=\int_{\mathbb{R}^{2}}{\mathbf{n}}_{X}(z)(\partial_{1}\eta_{2}(z)-\partial_{2}\eta_{1}(z)){\mathop{}\!\mathrm{d}}z.

When the smooth loop is replaced with a Brownian one, such a formula cannot be written down directly. For its left-hand side, we do have a genuine candidate provided by the Stratonovich integrale of η\eta along XX. However, the index function 𝐧X{\mathbf{n}}_{X} fails from being integrable on the vicinity of XX [12], and we need to use some kind of regularization in order to define the right-hand side. In such a framework, the Green formula is thus a convergence result rather than an equality.

In [13], Wendelin Werner proposed two alternative regularizations, for which he was able to prove that the Green formula holds with a convergence in probability.

In [11], I proposed two more regularizations, for which I proved that the Green formula holds with an almost sure limit, but only in the case 1η22η1=1\partial_{1}\eta_{2}-\partial_{2}\eta_{1}=1.

The first goal of this paper is to extend such a formula to any differential 11-form η\eta with regularity 𝒞1+ϵ\mathcal{C}^{1+\epsilon}.

For an integer xx and a positive integer kk, let [x]k[x]_{k} be equal to either x𝟙|x|kx\mathbbm{1}_{|x|\leq k} or max(min(x,k),k)\max(\min(x,k),-k) (the following theorem holds for both choice).

Theorem 1.

Let X:[0,T]2X:[0,T]\to\mathbb{R}^{2} be a Brownian motion, and let 𝐧X{\mathbf{n}}_{X} be the winding function associated with the loop obtained by concatenation of XX with the straight line segment [XT,X0][X_{T},X_{0}] between its endpoints. Then, almost surely, for all ϵ>0\epsilon>0 and all f𝒞bϵ(2)f\in\mathcal{C}_{b}^{\epsilon}(\mathbb{R}^{2}),

2[𝐧X(z)]kf(z)dz\int_{\mathbb{R}^{2}}[{\mathbf{n}}_{X}(z)]_{k}f(z){\mathop{}\!\mathrm{d}}z

converges as kk\to\infty.

Furthermore, if η=η1dx1+η2dx2\eta=\eta_{1}{\mathop{}\!\mathrm{d}}x^{1}+\eta_{2}{\mathop{}\!\mathrm{d}}x^{2} with η1,η2𝒞1+ϵ(2)\eta_{1},\eta_{2}\in\mathcal{C}^{1+\epsilon}(\mathbb{R}^{2}) is such that f=1η22η1f=\partial_{1}\eta_{2}-\partial_{2}\eta_{1}, almost surely,

limk2[𝐧X(z)]kf(z)dz=0TηdX+[XT,X0]η,\lim_{k\to\infty}\int_{\mathbb{R}^{2}}[{\mathbf{n}}_{X}(z)]_{k}f(z){\mathop{}\!\mathrm{d}}z=\int_{0}^{T}\eta\circ{\mathop{}\!\mathrm{d}}X+\int_{[X_{T},X_{0}]}\eta,

where the stochastic integral in the right hand side is to be understood in the sense of Stratonovich.

Corollary 2.

For all xx and yy in 2\mathbb{R}^{2}, the same result holds if the planar Brownian motion is replaced with a planar Brownian loop or a planar Brownian bridge between distinct points.

We will denote this limit as 2𝐧X(z)f(z)dz\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\mathbb{R}^{2}}{\mathbf{n}}_{X}(z)f(z){\mathop{}\!\mathrm{d}}z, since we want to think of it as to the integral of nXn_{X} with respect to the measure f(z)dzf(z){\mathop{}\!\mathrm{d}}z.

1.2. Magnetic impurities

In the theory of weak localization in 2 dimensional crystals, for which we refer to [2], one studies quasiclassical electrons moving inside a metal with magnetic impurities, in the presence of a magnetic fields which induces an Aharonov–Bohm effect on the electrons. In some regime of the parameters, the electron is usually modeled by a planar Brownian trajectory. In particular, for the computation of the weak-localization correction to the Drude conductivity, the electron is modeled by a Brownian loop (see e.g. [7]). The impurities are modeled by a Poisson process of points 𝒫\mathcal{P} with intensity ρdz\rho{\mathop{}\!\mathrm{d}}z in the plane, and the Aharonov–Bohm effect is described by a phase shift exp(iαz𝒫𝐧X(z))\exp(i\alpha\sum_{z\in\mathcal{P}}{\mathbf{n}}_{X}(z)).

In [4], the authors study the limit ρ+\rho\to+\infty with κ=αρ\kappa=\alpha\rho constant, and derive a formula for the phase shift averaged over both 𝒫\mathcal{P} and XX.

For an integrable function fL1(2)f\in L^{1}(\mathbb{R}^{2}), 1ρz𝒫f(z)\frac{1}{\rho}\sum_{z\in\mathcal{P}}f(z) is a Monte–Carlo estimation for 2f(z)dz\int_{\mathbb{R}^{2}}f(z){\mathop{}\!\mathrm{d}}z, and therefore

eiκ2f(z)dz=limρ𝔼𝒫[eiκρz𝒫f(z)].e^{i\kappa\int_{\mathbb{R}^{2}}f(z){\mathop{}\!\mathrm{d}}z}=\lim_{\rho\to\infty}\mathbb{E}^{\mathcal{P}}\big{[}e^{i\frac{\kappa}{\rho}\sum_{z\in\mathcal{P}}f(z)}\big{]}.

However, as it is noticed in [5], for a Brownian loop XX,

𝔼X[eiκ2𝐧X(z)dz]limρ𝔼X,𝒫[eiκρz𝒫𝐧X(z)],\mathbb{E}^{X}\big{[}e^{i\kappa\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-3.40277pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-2.275pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.60416pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.31248pt}}\!\int_{\mathbb{R}^{2}}{\mathbf{n}}_{X}(z){\mathop{}\!\mathrm{d}}z}\big{]}\neq\lim_{\rho\to\infty}\mathbb{E}^{X,\mathcal{P}}\big{[}e^{i\frac{\kappa}{\rho}\sum_{z\in\mathcal{P}}{\mathbf{n}}_{X}(z)}\big{]},

which is due to the lack of integrability of the function 𝐧X{\mathbf{n}}_{X}.

As we proved in [11], the Monte–Carlo method fails in this situation: it is true that XX-almost surely, 1ρz𝒫𝐧X(z)\frac{1}{\rho}\sum_{z\in\mathcal{P}}{\mathbf{n}}_{X}(z) converges in distribution (with respect to 𝒫\mathcal{P}) as ρ\rho\to\infty, but the limit is not deterministic –or should we say, not measurable with respect to XX. It is instead equal to the sum of 2𝐧X(z)dz\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\mathbb{R}^{2}}{\mathbf{n}}_{X}(z){\mathop{}\!\mathrm{d}}z with a centered Cauchy distribution independent from XX. From this result, one can rigorously prove the formula obtained first in [5] for

limρ𝔼X,𝒫[eiκρz𝒫𝐧X(z)].\lim_{\rho\to\infty}\mathbb{E}^{X,\mathcal{P}}[e^{i\frac{\kappa}{\rho}\sum_{z\in\mathcal{P}}{\mathbf{n}}_{X}(z)}].

However, for the scales at play, the magnetic field which induces the Aharonov-Bohm effect cannot be considered as homogeneous in general [8]. Our second goal in this paper is to derive an asymptotic formula for the functional of XX given by

limρ𝔼𝒫[ei1ρz𝒫f(z)𝐧X(z)],\lim_{\rho\to\infty}\mathbb{E}^{\mathcal{P}}[e^{i\frac{1}{\rho}\sum_{z\in\mathcal{P}}f(z){\mathbf{n}}_{X}(z)}],

for a non homogeneous magnetic field ff and a non homogeneous density of impurities.

Theorem 3.

Let f,g𝒞bϵ(2)f,g\in\mathcal{C}^{\epsilon}_{b}(\mathbb{R}^{2}), with g0g\geq 0. For ρ>0\rho>0, let 𝒫\mathcal{P} be Poisson process on 2\mathbb{R}^{2} with intensity ρg(z)dz\rho g(z){\mathop{}\!\mathrm{d}}z, and let XX be either a Brownian motion or a Brownian bridge with duration 11, independent from 𝒫\mathcal{P}. Then, XX-almost surely,

limρ𝔼𝒫[ei1ρz𝒫f(z)𝐧X(z)]=exp(iα𝐧X(z)f(z)g(z)dz|α|201|f(Xt)|g(Xt)dt)\lim_{\rho\to\infty}\mathbb{E}^{\mathcal{P}}[e^{i\frac{1}{\rho}\sum_{z\in\mathcal{P}}f(z){\mathbf{n}}_{X}(z)}]=\exp\Big{(}i\alpha\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X}(z)f(z)g(z){\mathop{}\!\mathrm{d}}z-\frac{|\alpha|}{2}\int_{0}^{1}|f(X_{t})|g(X_{t}){\mathop{}\!\mathrm{d}}t\Big{)}

where 𝔼𝒫\mathbb{E}^{\mathcal{P}} is the expectation over 𝒫\mathcal{P} (conditional on XX).

Although this formula is suited to the problem of magnetic impurities, the following alternative formulation might be more appealing to the reader.

Corollary 4.

Let g𝒞bϵ(2)g\in\mathcal{C}^{\epsilon}_{b}(\mathbb{R}^{2}), with g0g\geq 0. For ρ>0\rho>0, let 𝒫\mathcal{P} be Poisson process on 2\mathbb{R}^{2} with intensity ρg(z)dz\rho g(z){\mathop{}\!\mathrm{d}}z, and XX be either a Brownian motion or a Brownian bridge with duration 11, independent from 𝒫\mathcal{P}. Let also Γ:[0,1]\Gamma:[0,1]\to\mathbb{R} be a standard Cauchy process. Then, for all (f1,,fn)𝒞ϵ(2)(f_{1},\dots,f_{n})\in\mathcal{C}^{\epsilon}(\mathbb{R}^{2}), XX-almost surely, the nn-uple

(1ρz𝒫f1(z)𝐧X(z),,1ρz𝒫fn(z)𝐧X(z))\Big{(}\frac{1}{\rho}\sum_{z\in\mathcal{P}}f_{1}(z){\mathbf{n}}_{X}(z),\dots,\frac{1}{\rho}\sum_{z\in\mathcal{P}}f_{n}(z){\mathbf{n}}_{X}(z)\Big{)}

converges in distribution toward (ξ(f1),,ξ(fn))(\xi(f_{1}),\dots,\xi(f_{n})) where

ξ(f)=𝐧X(z)f(z)g(z)dz+1201f(Xt)g(Xt)dΓt.\xi(f)=\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X}(z)f(z)g(z){\mathop{}\!\mathrm{d}}z+\frac{1}{2}\int_{0}^{1}f(X_{t})g(X_{t}){\mathop{}\!\mathrm{d}}\Gamma_{t}.
Remark 5.

Given f,g𝒞bϵ(2)f,g\in\mathcal{C}^{\epsilon}_{b}(\mathbb{R}^{2}), there always exists a differential 11-form η\eta with regularity 𝒞1+ϵ\mathcal{C}^{1+\epsilon} such that 1η22η1=fg\partial_{1}\eta_{2}-\partial_{2}\eta_{1}=fg, so that 𝐧X(z)f(z)g(z)dz\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X}(z)f(z)g(z){\mathop{}\!\mathrm{d}}z can always be written as a stochastic integral.

Since all the results hold XX-almost surely, the assumptions that the functions are bounded can easily be lifted, but some of the intermediate results come with a quantitative version which depends upon the LL^{\infty} norms.

This paper is built in the continuity of two former papers from the same author, [11] and [9]. It is not necessary to read them to understand the present paper, but we will use some results from those papers, as well as from [10].

2. Notations

2.1. Differential forms and integrals

For α(0,1)\alpha\in(0,1), we define 𝒞α(2)\mathcal{C}^{\alpha}(\mathbb{R}^{2}) as the set of functions f:2f:\mathbb{R}^{2}\to\mathbb{R} such that the semi-norm

|f|𝒞αsupx,y2xyf(x)f(y)|xy||f|_{\mathcal{C}^{\alpha}}\coloneqq\sup_{\begin{subarray}{c}x,y\in\mathbb{R}^{2}\\ x\neq y\end{subarray}}\frac{f(x)-f(y)}{|x-y|}

is finite. We also define 𝒞bα(2)=𝒞α(2)L2(2)\mathcal{C}_{b}^{\alpha}(\mathbb{R}^{2})=\mathcal{C}^{\alpha}(\mathbb{R}^{2})\cap L^{2}(\mathbb{R}^{2}), which we endow with the norm

f𝒞bα=f+|f|𝒞α.\|f\|_{\mathcal{C}^{\alpha}_{b}}=\|f\|_{\infty}+|f|_{\mathcal{C}^{\alpha}}.

For a differential 11-form η=η1dx1+η2dx2\eta=\eta_{1}{\mathop{}\!\mathrm{d}}x^{1}+\eta_{2}{\mathop{}\!\mathrm{d}}x^{2} and α[0,1)\alpha\in[0,1), we write η𝒞1+α(T2)\eta\in\mathcal{C}^{1+\alpha}(T^{*}\mathbb{R}^{2}) if iηj𝒞α(2)\partial_{i}\eta_{j}\in\mathcal{C}^{\alpha}(\mathbb{R}^{2}) for all i,j{1,2}i,j\in\{1,2\}.

Given a curve X:[0,T]2X:[0,T]\to\mathbb{R}^{2}, we write

Xη0Tη1(Xt)dXt1+0Tη2(Xt)dXt2,\int_{X}\eta\coloneqq\int_{0}^{T}\eta_{1}(X_{t}){\mathop{}\!\mathrm{d}}X^{1}_{t}+\int_{0}^{T}\eta_{2}(X_{t}){\mathop{}\!\mathrm{d}}X^{2}_{t},

where these integrals are to be understood either as classical integrals or as Stratonovich integrals, depending on the regularity of XX. No Itô integral will be involved in this paper, and all the stochastic integrals are to be understood in the sense of Stratonovich.

For η𝒞1+α(T2)\eta\in\mathcal{C}^{1+\alpha}(T^{*}\mathbb{R}^{2}), we identify the 22-form dα=(1η22η1)dx1dx2{\mathop{}\!\mathrm{d}}\alpha=(\partial_{1}\eta_{2}-\partial_{2}\eta_{1}){\mathop{}\!\mathrm{d}}x^{1}\wedge{\mathop{}\!\mathrm{d}}x^{2} with the signed measure (1η22η1)dx(\partial_{1}\eta_{2}-\partial_{2}\eta_{1}){\mathop{}\!\mathrm{d}}x, where dx{\mathop{}\!\mathrm{d}}x is the Lebesgue measure on 2\mathbb{R}^{2}.

For a bounded set 𝒟2\mathcal{D}\subset\mathbb{R}^{2} and fLloc1(2)f\in L^{1}_{loc}(\mathbb{R}^{2}), we use the unconventional notation

f(𝒟)=𝒟f(z)dz,f(\mathcal{D})=\int_{\mathcal{D}}f(z){\mathop{}\!\mathrm{d}}z,

and |𝒟||\mathcal{D}| for the Lebesgue measure of 𝒟\mathcal{D}.

2.2. Winding

Given a curve XX on 2\mathbb{R}^{2}, that is a continuous function from [0,T][0,T] to 2\mathbb{R}^{2} for some T>0T>0, we write X¯\bar{X} for the concatenation of XX with a straight line segment from XTX_{T} to X0X_{0}. Although the parameterisation of this line segment does not matter in the following, we will assume it is parameterized by [T,T+1][T,T+1] at constant speed, unless XX is a loop (that is, a curve with XT=X0X_{T}=X_{0}), in which case we set X¯=X\bar{X}=X.

Given a curve XX and a point zz outside the range of X¯\bar{X}, we write 𝐧X(z){\mathbf{n}}_{X}(z) for the winding number of X¯\bar{X} around zz.

For a relative integer kk, we define

𝒜kX={z2Range(X¯):𝐧X(z)=k}.\mathcal{A}_{k}^{X}=\{z\in\mathbb{R}^{2}\setminus\operatorname{Range}(\bar{X}):{\mathbf{n}}_{X}(z)=k\}.

For n>0n>0, we also define

𝒟nX={z2Range(X¯):𝐧X(z)n}=nk<+𝒜kX,\mathcal{D}_{n}^{X}=\{z\in\mathbb{R}^{2}\setminus\operatorname{Range}(\bar{X}):{\mathbf{n}}_{X}(z)\geq n\}=\bigsqcup_{n\leq k<+\infty}\mathcal{A}_{k}^{X},

and

𝒟nX={z2Range(X¯):𝐧X(z)n}=<kn𝒜kX.\mathcal{D}_{-n}^{X}=\{z\in\mathbb{R}^{2}\setminus\operatorname{Range}(\bar{X}):{\mathbf{n}}_{X}(z)\leq-n\}=\bigsqcup_{-\infty<k\leq-n}\mathcal{A}_{k}^{X}.

We also write AkXA_{k}^{X} (resp. DkXD_{k}^{X}) for the Lebesgue measure of 𝒜kX\mathcal{A}_{k}^{X} (resp. 𝒟kX\mathcal{D}_{k}^{X}), and we drop the superscript XX when there is no doubt about the curve we are talking about.

For a real number zz and a positive integer nn, we set

[x]n={nif xn,xif nxn,nif xn.[x]_{n}=\left\{\begin{array}[]{cc}-n&\mbox{if }x\leq-n,\\ x&\mbox{if }-n\leq x\leq n,\\ n&\mbox{if }x\geq n.\\ \end{array}\right.

Once we have shown that the limit

limk2f(z)[nX(z)]kdz\lim_{k\to\infty}\int_{\mathbb{R}^{2}}f(z)[n_{X}(z)]_{k}{\mathop{}\!\mathrm{d}}z

almost surely exists for all f𝒞ϵ(2)f\in\mathcal{C}^{\epsilon}(\mathbb{R}^{2}), we will write 2f(z)nX(z)dz\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\mathbb{R}^{2}}f(z)n_{X}(z){\mathop{}\!\mathrm{d}}z for this limit.

For a locally finite set of points 𝒫\mathcal{P}, we define 𝐧X(𝒫){\mathbf{n}}_{X}(\mathcal{P}) as the sum z𝒫𝐧X(z)\sum_{z\in\mathcal{P}}{\mathbf{n}}_{X}(z). If we are also given a function ff on 2\mathbb{R}^{2}, we define 𝐧X(𝒫,f){\mathbf{n}}_{X}(\mathcal{P},f) as the weighted sum

𝐧X(𝒫,f)=z𝒫𝐧X(z)f(z).{\mathbf{n}}_{X}(\mathcal{P},f)=\sum_{z\in\mathcal{P}}{\mathbf{n}}_{X}(z)f(z).

2.3. Cauchy variables

The Cauchy distribution 𝒞(p,σ)\mathcal{C}(p,\sigma) with position parameter pp and scale parameter σ>0\sigma>0 is the probability distribution on \mathbb{R} which has a density with respect to the Lebesgue measure given at xx by

1πσσ2σ2+(xp)2.\frac{1}{\pi\sigma}\frac{\sigma^{2}}{\sigma^{2}+(x-p)^{2}}.

A Cauchy random variable with position parameter pp and scale parameter σ\sigma is a random variable distributed according to 𝒞(p,σ)\mathcal{C}(p,\sigma). In ordre to unify some results, we will also write 𝒞(p,0)\mathcal{C}(p,0) for a random variable which is actually deterministic and equal to pp.

Following [6, Definition 5.2]111As opposed to [6], we include the trivial case σ=0\sigma=0 in our definition. , we will say that a random variable ZZ on \mathbb{R} lies in the strong domain of attraction of a Cauchy distribution if there exists σ0,δ>0\sigma\geq 0,\delta>0 such that

(Zx)=x+σπx+o(x(1+δ)),(Zx)=x+σπx+o(x(1+δ)).\mathbb{P}(Z\geq x)\underset{x\to+\infty}{=}\frac{\sigma}{\pi x}+o(x^{-(1+\delta)}),\qquad\mathbb{P}(Z\leq-x)\underset{x\to+\infty}{=}\frac{\sigma}{\pi x}+o(x^{-(1+\delta)}).

It then follows from Lemma 5.1 and Theorem 1.2 in [6] that ZZ follows a central limit theorem: if (Zi)i(Z_{i})_{i\in\mathbb{N}} are i.i.d. copies of ZZ, then there exists a unique pp such that

1Ni=1NZiY𝒞(p,σ).\frac{1}{N}\sum_{i=1}^{N}Z_{i}\Longrightarrow Y\sim\mathcal{C}(p,\sigma).

Notice that the same assumptions with δ=0\delta=0 are not sufficient for such a central limit theorem to hold.

The parameters pp and σ\sigma such that Y𝒞(p,σ)Y\sim\mathcal{C}(p,\sigma) are uniquely defined. We call them respectively the position parameter pZp_{Z} of ZZ, and the scale parameter σZ\sigma_{Z} of ZZ.222When ZZ is a Cauchy random variable, it belongs to the strong domain of attraction of a Cauchy distribution. There is thus two definitions of its position parameter, and two definitions of its scale parameter. Of course, the two definitions of its position parameter agree, and the two definitions of its scale parameter agree as well.

3. Former results

We will use the following results from [11], [9] and [10].

Lemma 3.1 (Lemma 5.2 in [11] ).

Assume ZZ belongs to the strong attraction domain of a Cauchy distribution. Then, its position parameter pZp_{Z} is equal to

limn𝔼[[Z]n].\lim_{n\to\infty}\mathbb{E}[[Z]_{n}].

When YY and ZZ lie in the strong attraction domain of Cauchy distributions, or even when they are Cauchy random variables, but they are not independent, Y+ZY+Z does not necessarily belong to the strong attraction domain of a Cauchy distribution. What might be even more surprising is that, even if YY, ZZ, and Y+ZY+Z are Cauchy random variables, pY+Zp_{Y+Z} can differ from pY+pZp_{Y}+p_{Z} (see e.g. [3] for an explicit counter-example). Yet, the following lemma offers conditions weaker then independence under which additivity is restored.

Lemma 3.2 ( Lemma 5.3 in [11] ).

Let n1n\geq 1 and Z1,,ZnZ_{1},\dots,Z_{n} be random variables which each lie in the strong attraction domain of a Cauchy distribution. Assume that there exists δ>0\delta>0 such that, for all i,j{1,,n}i,j\in\{1,\dots,n\}, iji\neq j,

(|Zi|x and |Zj|x)=x+o(x(1+δ)).\mathbb{P}(|Z_{i}|\geq x\mbox{ and }|Z_{j}|\geq x)\underset{x\to+\infty}{=}o(x^{-(1+\delta)}).

Then, Z=i=1nZiZ=\sum_{i=1}^{n}Z_{i} lies in the strong attraction domain of a Cauchy distribution, and pZ=i=1npZip_{Z}=\sum_{i=1}^{n}p_{Z_{i}}.

The following lemma should be compared with the definition of the strong domain of attraction, where the random variable ZZ is given by 𝐧X(P){\mathbf{n}}_{X}(P), with XX fixed and PP a random point distributed according to 1Z𝟙K(z)f(z)dz\frac{1}{Z}\mathbbm{1}_{K}(z)f(z){\mathop{}\!\mathrm{d}}z (when f0f\geq 0), where KK is a convex set containing Range(X)\operatorname{Range}(X).

Lemma 3.3 (Lemma 5 in [9]).

Let X:[0,1]2X:[0,1]\to\mathbb{R}^{2} be a planar Brownian motion. For all β<12\beta<\frac{1}{2}, there exists δ>0\delta>0 such that almost surely, there exists CC such that for all bounded continuous function f𝒞b(2)f\in\mathcal{C}_{b}({\mathbb{R}}^{2}), for all n1n\geq 1,

|2πnf(𝒟n)01f(Xu)du|C(ωf(2X𝒞βnδ)+fnδ),\Big{|}2\pi nf(\mathcal{D}_{n})-\int_{0}^{1}f(X_{u}){\mathop{}\!\mathrm{d}}u\Big{|}\leq C(\omega_{f}(2\|X\|_{\mathcal{C}^{\beta}}n^{-\delta})+\|f\|_{\infty}n^{-\delta}),

where ωf\omega_{f} is the continuity modulus of ff, i.e. ωf(r)=supx,y:|xy|r|f(x)f(y)|\omega_{f}(r)=\sup_{x,y:|x-y|\leq r}|f(x)-f(y)|.

From symmetry of the Brownian motion, Lemma 3.3 also holds with 𝒟n\mathcal{D}_{n} replaced with 𝒟n\mathcal{D}_{-n}.

We will also need some LpL^{p} control.

Lemma 3.4 ( Theorem 6.2 in [11] ).

For all δ<12\delta<\frac{1}{2} and p2p\geq 2, there exists a constant CC such that for all N1N\geq 1,

𝔼[|DN12πN|p]1pCN1δ.\mathbb{E}\big{[}\big{|}D_{N}-\tfrac{1}{2\pi N}\big{|}^{p}\big{]}^{\frac{1}{p}}\leq CN^{-1-\delta}.

Finally, the following lemma will be used to check the condition inside Lemma 3.2.

Lemma 3.5 (Theorem 1 in [10]).

Let X,X:[0,1]2X,X^{\prime}:[0,1]\to\mathbb{R}^{2} be two independent Brownian motions, starting from equal or different points in the plane. Then, n2|𝒟nX𝒟nX|n^{2}|\mathcal{D}^{X}_{n}\cap\mathcal{D}^{X^{\prime}}_{n}| almost surely converges as nn\to\infty.

A few more results will be used, but will be easier to formulate later.

4. Stokes formula

In this section, X:[0,1]2X:[0,1]\to\mathbb{R}^{2} is a standard Brownian motion under \mathbb{P}.

4.1. Existence of a limit

We will first prove the first part of Theorem 1:

Lemma 4.1.

Let ϵ>0\epsilon>0. \mathbb{P}-almost surely, for all f𝒞bϵ(2)f\in\mathcal{C}_{b}^{\epsilon}(\mathbb{R}^{2}), the limits

𝐧X(x)f(x)dxlimN2[𝐧X(z)]Nf(z)dzandlimN2𝐧X(z)𝟙|𝐧X(z)|Nf(z)dz\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X}(x)f(x){\mathop{}\!\mathrm{d}}x\coloneqq\lim_{N\to\infty}\int_{\mathbb{R}^{2}}[{\mathbf{n}}_{X}(z)]_{N}f(z){\mathop{}\!\mathrm{d}}z\qquad\mbox{and}\qquad\lim_{N\to\infty}\int_{\mathbb{R}^{2}}{\mathbf{n}}_{X}(z)\mathbbm{1}_{|{\mathbf{n}}_{X}(z)|\leq N}\ f(z){\mathop{}\!\mathrm{d}}z

exist and are equal. Almost surely, the application f𝐧X(f)f\mapsto{\mathbf{n}}_{X}(f) from 𝒞bϵ(2)\mathcal{C}^{\epsilon}_{b}(\mathbb{R}^{2}) to \mathbb{R} is continuous.

Proof.

We fix β(0,12)\beta\in\big{(}0,\tfrac{1}{2}\big{)}. Let δ>0\delta>0 be such that Lemma 3.3 holds, and let \mathcal{E} be the full probability event on which X𝒞β<\|X\|_{\mathcal{C}^{\beta}}<\infty and Lemma 3.3 holds both for the sequence 𝒟n\mathcal{D}_{n} and the sequence 𝒟n\mathcal{D}_{-n}, with a corresponding random constant CC.

On \mathcal{E}, for all ϵ>0\epsilon>0, with C=4πC,C′′=C(1+|X|𝒞βϵ)C^{\prime}=4\pi C,C^{\prime\prime}=C^{\prime}(1+|X|_{\mathcal{C}^{\beta}}^{\epsilon}), for all f𝒞ϵ(2)f\in\mathcal{C}^{\epsilon}(\mathbb{R}^{2}),

|f(𝒟n)f(𝒟n)|\displaystyle\Big{|}f(\mathcal{D}_{n})-f(\mathcal{D}_{-n})\Big{|} Cn1(ωf(2|X|𝒞βnδ)+fnδ)\displaystyle\leq C^{\prime}n^{-1}(\omega_{f}(2|X|_{\mathcal{C}^{\beta}}n^{-\delta})+\|f\|_{\infty}n^{-\delta})
C′′n1(|f|𝒞ϵnδϵ+fnδ).\displaystyle\leq C^{\prime\prime}n^{-1}(|f|_{\mathcal{C}^{\epsilon}}n^{-\delta\epsilon}+\|f\|_{\infty}n^{-\delta}). (1)

Thus, on \mathcal{E}, the sum

n1(f(𝒟n)f(𝒟n))\sum_{n\geq 1}(f(\mathcal{D}_{n})-f(\mathcal{D}_{-n}))

is absolutely convergent. By applying an Abel summation, we obtain

n=1N(f(𝒟n)f(𝒟n))=2[𝐧X(z)]Nf(z)dz,\sum_{n=1}^{N}(f(\mathcal{D}_{n})-f(\mathcal{D}_{-n}))=\int_{\mathbb{R}^{2}}[{\mathbf{n}}_{X}(z)]_{N}f(z){\mathop{}\!\mathrm{d}}z,

so that the right-hand side is convergent on the event \mathcal{E}.

Besides,

|2[𝐧X(z)]Nf(z)dz2𝐧X(z)𝟙|𝐧X(z)|Nf(z)dz|=N|f(𝒟N+1)f(𝒟N1)|,\Big{|}\int_{{\mathbb{R}}^{2}}[{\mathbf{n}}_{X}(z)]_{N}f(z){\mathop{}\!\mathrm{d}}z-\int_{{\mathbb{R}}^{2}}{\mathbf{n}}_{X}(z)\mathbbm{1}_{|{\mathbf{n}}_{X}(z)|\leq N}f(z){\mathop{}\!\mathrm{d}}z\Big{|}=N|f(\mathcal{D}_{N+1})-f(\mathcal{D}_{-N-1})|,

which, on the almost sure event \mathcal{E}, converges toward 0 as NN goes to infinity (by (1)).

The only thing that remains to be shown is the almost sure continuity of the application f𝐧X(x)f(x)dxf\mapsto\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X}(x)f(x){\mathop{}\!\mathrm{d}}x. Since it is clearly linear, it suffices to show that it is almost surely a bounded operator. By (1),

n=1N|f(𝒟n)f(𝒟n)|\sum_{n=1}^{N}|f(\mathcal{D}_{n})-f(\mathcal{D}_{-n})|

is bounded by C(3)f𝒞bϵC^{(3)}\|f\|_{\mathcal{C}^{\epsilon}_{b}} for a random constant 𝒞(3)\mathcal{C}^{(3)} which depends on ϵ,β\epsilon,\ \beta and δ\delta, but not on ff nor NN. Thus, |𝐧X(x)f(x)dx|C(3)f𝒞bϵ|\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X}(x)f(x){\mathop{}\!\mathrm{d}}x|\leq C^{(3)}\|f\|_{\mathcal{C}^{\epsilon}_{b}}, which concludes the proof. ∎

4.2. Strategy for the Stokes’ formula

In order to conclude the proof of Theorem 1, we now need to identify 𝐧X(x)f(x)dx\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X}(x)f(x){\mathop{}\!\mathrm{d}}x with the Stratonovich integral Xη+[X1,X0]η\int_{X}\eta+\int_{[X_{1},X_{0}]}\eta, when f=1η22η1f=\partial_{1}\eta_{2}-\partial_{2}\eta_{1}.

To this end, we decompose the trajectory XX into several pieces. First, we denote by X(n)X^{(n)} the dyadic piecewise-linear approximation of XX with 2n2^{n} pieces: for λ[0,1],i{0,,2n1}\lambda\in[0,1],\ i\in\{0,\dots,2^{n}-1\}, and t=(i+λ)2nt=(i+\lambda)2^{-n},

Xt(n)=Xi2n+λ(X(i+1)2nXi2n).X^{(n)}_{t}=X_{i2^{-n}}+\lambda(X_{(i+1)2^{-n}}-X_{i2^{-n}}).

For i{0,,2n1}i\in\{0,\dots,2^{n}-1\}, we also set XiX^{i}, the restriction of XX to the interval [i2n,(i+1)2n][i2^{-n},(i+1)2^{-n}]. Finally, set 𝐧Xi(x)f(x)dx\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X_{i}}(x)f(x){\mathop{}\!\mathrm{d}}x the almost sure limit

𝐧Xi(z)f(z)dz=limN2[𝐧Xi(z)]Nf(z)dz.\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X_{i}}(z)f(z){\mathop{}\!\mathrm{d}}z=\lim_{N\to\infty}\int_{{\mathbb{R}}^{2}}[{\mathbf{n}}_{X^{i}}(z)]_{N}f(z){\mathop{}\!\mathrm{d}}z.

By Lemma 4.1, scale invariance, and translation invariance of the Brownian motion, almost surely, 𝐧Xi(x)f(x)dx\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X_{i}}(x)f(x){\mathop{}\!\mathrm{d}}x is well -defined for all n0n\geq 0, for all i{0,,2n1}i\in\{0,\dots,2^{n}-1\}, for all f𝒞ϵ(2)f\in\mathcal{C}^{\epsilon}(\mathbb{R}^{2}).333Since we use the translation invariance, the function ff is replaced with the random function zf(z+Xi2n)z\mapsto f(z+X_{i2^{-n}}). This is not an issue, because Lemma 4.1 holds almost surely for all ff, and not the other way around.

Let us first sketch the strategy of our proof. First, notice that for all z2z\in\mathbb{R}^{2} which does not belong to Range(X)\operatorname{Range}(X) nor to Range(X(n))\operatorname{Range}(X^{(n)}),

𝐧X(z)=i=02n1𝐧Xi(z)+𝐧X(n)(z),{\mathbf{n}}_{X}(z)=\sum_{i=0}^{2^{n}-1}{\mathbf{n}}_{X_{i}}(z)+{\mathbf{n}}_{X^{(n)}}(z),

which essentially comes from the additivity of the winding index, with respect to the concatenation of loops. Thus, it is reasonable to expect that

𝐧X(z)f(z)dz=i=02n1𝐧Xi(z)f(z)dz+2𝐧X(n)(z)f(z)dz.\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X}(z)f(z){\mathop{}\!\mathrm{d}}z=\sum_{i=0}^{2^{n}-1}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X_{i}}(z)f(z){\mathop{}\!\mathrm{d}}z+\int_{\mathbb{R}^{2}}{\mathbf{n}}_{X^{(n)}}(z)f(z){\mathop{}\!\mathrm{d}}z.

By applying the standard Stokes’ formula on the last integral, we get

𝐧X(z)f(z)dz=i=02n1𝐧Xi(z)f(z)dz+X(n)η+[X1,X0]η.\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X}(z)f(z){\mathop{}\!\mathrm{d}}z=\sum_{i=0}^{2^{n}-1}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X_{i}}(z)f(z){\mathop{}\!\mathrm{d}}z+\int_{X^{(n)}}\eta+\int_{[X_{1},X_{0}]}\eta.

As nn goes to infinity, we will see that the contribution from the small pieces (i.e. the sum over ii) vanishes, whilst the integral along X(n)X^{(n)} converges toward the Stratonovich integral Xη\int_{X}\eta, which gives the expected formula.

We will decompose the actual proof into the three following lemma, which we will prove in the three following subsections. Let f𝒞bϵ(2)f\in\mathcal{C}^{\epsilon}_{b}(\mathbb{R}^{2}), and η𝒞1+ϵ(T2)\eta\in\mathcal{C}^{1+\epsilon}(T^{*}\mathbb{R}^{2}) such that f=1η22η1f=\partial_{1}\eta_{2}-\partial_{2}\eta_{1}.

Lemma 4.2.

For all nn, almost surely,

𝐧X(z)f(z)dz=i=02n1𝐧Xi(z)f(z)dz+2𝐧X(n)(z)f(z)dz.\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X}(z)f(z){\mathop{}\!\mathrm{d}}z=\sum_{i=0}^{2^{n}-1}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X_{i}}(z)f(z){\mathop{}\!\mathrm{d}}z+\int_{\mathbb{R}^{2}}{\mathbf{n}}_{X^{(n)}}(z)f(z){\mathop{}\!\mathrm{d}}z. (2)
Lemma 4.3.

As nn goes to infinity,

i=02n1𝐧Xi(z)f(z)dz\sum_{i=0}^{2^{n}-1}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X_{i}}(z)f(z){\mathop{}\!\mathrm{d}}z

converges almost surely toward zero.

Lemma 4.4.

As nn goes to infinity, X(n)η\int_{X^{(n)}}\eta converges almost surely toward ηdX\int\eta\circ{\mathop{}\!\mathrm{d}}X.

Of course, the conclusion that almost surely,

𝐧X(z)f(z)dz=Xη+[X1,X0]η,\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X}(z)f(z){\mathop{}\!\mathrm{d}}z=\int_{X}\eta+\int_{[X_{1},X_{0}]}\eta,

and therefore that Theorem 1 holds, follows directly from these three lemma.

4.3. Additivity

Intuitively, the equality in Lemma 4.2 follows from integration of the almost-everywhere equality

𝐧X(z)=i=02n1𝐧Xi(z)+𝐧X(n)(z),{\mathbf{n}}_{X}(z)=\sum_{i=0}^{2^{n}-1}{\mathbf{n}}_{X^{i}}(z)+{\mathbf{n}}_{X^{(n)}}(z),

applied together with the Stokes formula for X(n)X^{(n)}. However, neither 𝐧X{\mathbf{n}}_{X} nor 𝐧Xi{\mathbf{n}}_{X^{i}} are integrable, we have to deal with the cut-offs that allow to define 𝐧X(z)f(z)dz\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X}(z)f(z){\mathop{}\!\mathrm{d}}z and the 𝐧Xi(z)f(z)dz\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X^{i}}(z)f(z){\mathop{}\!\mathrm{d}}z : in general, for a finite kk,

[𝐧X(z)]ki=02n1[𝐧Xi(z)]k+[𝐧X(n)(z)]k.[{\mathbf{n}}_{X}(z)]_{k}\neq\sum_{i=0}^{2^{n}-1}[{\mathbf{n}}_{X^{i}}(z)]_{k}+[{\mathbf{n}}_{X^{(n)}}(z)]_{k}.
Proof of Lemma 4.2.

From linearity with respect to ff, we can and we do assume f0f\geq 0. In the event that that the restriction of ff to B(0,X)B(0,\|X\|_{\infty}) is identically vanishing, the result is trivial, and we thus assume that

ZB(0,X)f(z)dzZ\coloneqq\int_{B(0,\|X\|_{\infty})}f(z){\mathop{}\!\mathrm{d}}z

is strictly positive.

Let PP be a random point in 2\mathbb{R}^{2} those distribution conditional on XX admits a density with respect to the Lebesgue measure, given by

f(z)𝟙B(0,X)(z)Z.\frac{f(z)\mathbbm{1}_{B(0,\|X\|_{\infty})}(z)}{Z}.

Then, XX-almost surely, PP-almost surely,

𝐧X(P)=i=02n1𝐧Xi(P)+𝐧X(n)(P).{\mathbf{n}}_{X}(P)=\sum_{i=0}^{2^{n}-1}{\mathbf{n}}_{X_{i}}(P)+{\mathbf{n}}_{X^{(n)}}(P).

Notice that, for N0N\geq 0, for X~\tilde{X} equal to either XX, or to one of the XiX^{i}, or to X(n)X^{(n)}, it holds that

(𝐧X~(P)N|X)=1Zf(𝒟NX~),(𝐧X~(P)N|X)=1Zf(𝒟NX~).\mathbb{P}({\mathbf{n}}_{\tilde{X}}(P)\geq N|X)=\frac{1}{Z}f(\mathcal{D}^{\tilde{X}}_{N}),\qquad\mathbb{P}({\mathbf{n}}_{\tilde{X}}(P)\leq-N|X)=\frac{1}{Z}f(\mathcal{D}^{\tilde{X}}_{-N}).

Thus, Lemma 3.3 ensures that XX-almost surely, the random variable 𝐧X~(P){\mathbf{n}}_{\tilde{X}}(P) belong to the strong attraction domain of a Cauchy distribution for either X~=X\tilde{X}=X or X~=Xi\tilde{X}=X^{i}. As for X~=X(n)\tilde{X}=X^{(n)}, |𝐧X~||{\mathbf{n}}_{\tilde{X}}| is bounded by 2n2^{n} and therefore 𝐧X~(P){\mathbf{n}}_{\tilde{X}}(P) also belong to the strong attraction domain of a (degenerate, σ=0\sigma=0) Cauchy distribution.

Let us check that, XX-almost surely, we can apply Lemma 3.2 to the set of variables

(Z0,,Z2n1,Z2n)=(𝐧X0(P),,𝐧X2n1(P),𝐧X(n)(P)).(Z_{0},\dots,Z_{2^{n}-1},Z_{2^{n}})=({\mathbf{n}}_{X^{0}}(P),\dots,{\mathbf{n}}_{X^{2^{n}-1}}(P),{\mathbf{n}}_{X^{(n)}}(P)).

First, for i{0,,2n1}i\in\{0,\dots,2^{n}-1\}, for x2nx\geq 2^{n},

(|𝐧Xi(P)|x and |𝐧X(n)(P)|x)=0=o(x(1+δ)).\mathbb{P}(|{\mathbf{n}}_{X^{i}}(P)|\geq x\mbox{ and }|{\mathbf{n}}_{X^{(n)}}(P)|\geq x)=0=o(x^{-(1+\delta)}).

Besides, for i,j{0,,2n1}i,j\in\{0,\dots,2^{n}-1\}, iji\neq j,

(|𝐧Xi(P)|N and |𝐧Xj(P)|N)\displaystyle\mathbb{P}(|{\mathbf{n}}_{X^{i}}(P)|\geq N\mbox{ and }|{\mathbf{n}}_{X^{j}}(P)|\geq N) =1Zf((𝒟NXi𝒟NXi)(𝒟NXj𝒟NXj))\displaystyle=\frac{1}{Z}f\big{(}\big{(}\mathcal{D}_{N}^{X^{i}}\cup\mathcal{D}_{-N}^{X^{i}}\big{)}\cap\big{(}\mathcal{D}_{N}^{X^{j}}\cup\mathcal{D}_{-N}^{X^{j}}\big{)}\big{)}
CfZ|N|2,\displaystyle\leq\frac{C\|f\|_{\infty}}{Z}|N|^{2},

for a random constant CC. The last equality follows from Lemma 3.5, applied to the independent Brownian motions

X^i:tX(i+1t)2nX(i+1)2n,X^j:tX(j+t)2nX(i+1)2n.\hat{X}_{i}:t\mapsto X_{(i+1-t)2^{-n}}-X_{(i+1)2^{-n}},\qquad\hat{X}_{j}:t\mapsto X_{(j+t)2^{-n}}-X_{(i+1)2^{-n}}.

Notice that the constant C=C(n,i,j)C=C(n,i,j) depends upon ii and jj, but we can replace it with C(n)=maxi,jC(n,i,j)C(n)=\max_{i,j}C(n,i,j) so that it only depends on nn. Furthermore, since there is only countably many couples (i,j)(i,j), the previous inequality holds almost surely for all (i,j)(i,j) simultaneously.

Thus, we can indeed apply Lemma 3.2 to deduce that the, XX-almost surely, the position parameters add up:

p𝐧X(P)=i=02n1p𝐧Xi(P)+p𝐧X(n)(P).p_{{\mathbf{n}}_{X}(P)}=\sum_{i=0}^{2^{n}-1}p_{{\mathbf{n}}_{X^{i}}(P)}+p_{{\mathbf{n}}_{X^{(n)}}(P)}. (3)

Furthermore, since |𝐧X(n)(P)||{\mathbf{n}}_{X^{(n)}}(P)| is bounded, p𝐧X(n)(P)p_{{\mathbf{n}}_{X^{(n)}}(P)} is quickly checked to be equal 𝔼P[𝐧X(n)(P)|X]\mathbb{E}^{P}[{\mathbf{n}}_{X^{(n)}}(P)|X], that is

p𝐧X(n)(P)=1Z2𝐧X(n)(z)f(z)dz.p_{{\mathbf{n}}_{X^{(n)}}(P)}=\frac{1}{Z}\int_{\mathbb{R}^{2}}{\mathbf{n}}_{X^{(n)}}(z)f(z){\mathop{}\!\mathrm{d}}z.

Finally, from Lemma 3.1, we deduce that XX-almost surely,

pnX(P)=limN𝔼P[[𝐧X(P)]N|X]=limN1Z2[𝐧X(z)]Nf(z)dz=1Z𝐧X(z)f(z)dz,\displaystyle p_{n_{X}}(P)=\lim_{N\to\infty}\mathbb{E}^{P}\big{[}[{\mathbf{n}}_{X}(P)]_{N}\big{|}X\big{]}=\lim_{N\to\infty}\frac{1}{Z}\int_{\mathbb{R}^{2}}[{\mathbf{n}}_{X}(z)]_{N}f(z){\mathop{}\!\mathrm{d}}z=\frac{1}{Z}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X}(z)f(z){\mathop{}\!\mathrm{d}}z,

and similarly

pnXi(P)=1Z𝐧Xi(z)f(z)dz.p_{n_{X^{i}}}(P)=\frac{1}{Z}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X^{i}}(z)f(z){\mathop{}\!\mathrm{d}}z.

Thus, Equation 3 turns into

𝐧X(z)f(z)dz=i=02n1𝐧Xi(z)f(z)dz+2𝐧X(n)(z)f(z)dz,\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X}(z)f(z){\mathop{}\!\mathrm{d}}z=\sum_{i=0}^{2^{n}-1}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X^{i}}(z)f(z){\mathop{}\!\mathrm{d}}z+\int_{\mathbb{R}^{2}}{\mathbf{n}}_{X^{(n)}}(z)f(z){\mathop{}\!\mathrm{d}}z,

as announced. ∎

4.4. Contribution from the small loops

We now prove that almost surely,

i=02n1𝐧Xi(z)f(z)dzn0.\sum_{i=0}^{2^{n}-1}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X^{i}}(z)f(z){\mathop{}\!\mathrm{d}}z\underset{n\to\infty}{\longrightarrow}0.

We will first need the following result, which should be compared with Lemma 3.3.

Lemma 4.5.

Let ϵ>0\epsilon>0 and p1p\geq 1. There exists a constant CC and δ>0\delta>0 such that for all f𝒞ϵ(2)f\in\mathcal{C}^{\epsilon}(\mathbb{R}^{2}) and all N1N\geq 1,

𝔼[|f(𝒟NX)12πN01f(Xt)dt|p]1pCN1δf𝒞ϵ.\mathbb{E}\big{[}\big{|}f(\mathcal{D}^{X}_{N})-\frac{1}{2\pi N}\int_{0}^{1}f(X_{t}){\mathop{}\!\mathrm{d}}t\big{|}^{p}\big{]}^{\frac{1}{p}}\leq CN^{-1-\delta}\|f\|_{\mathcal{C}^{\epsilon}}.
Proof.

The proof is largely inspired from [9].

Let T1T\geq 1, which we will later take to be a function of NN. For i{0,,T1}i\in\{0,\dots,T-1\}, let XiX^{i} be the restriction of XX to the interval [iT1,(i+1)T1][iT^{-1},(i+1)T^{-1}]. Let XplX^{pl} be the piecewise linear approximation of XX with TT pieces,

X(i+λ)T1pl=XiT1+λ(X(i+1)T1XiT1),i{0,,T1},λ[0,1].X^{pl}_{(i+\lambda)T^{-1}}=X_{iT^{-1}}+\lambda(X_{(i+1)T^{-1}}-X_{iT^{-1}}),\qquad i\in\{0,\dots,T-1\},\ \lambda\in[0,1].

For i,j{0,,T1}i,j\in\{0,\dots,T-1\}, let

𝒟Ni=𝒟NXi,𝒟Ni,j=(𝒟NXi𝒟NXi)(𝒟NXj𝒟NXj).\mathcal{D}^{i}_{N}=\mathcal{D}^{X^{i}}_{N},\qquad\mathcal{D}^{i,j}_{N}=\big{(}\mathcal{D}^{X^{i}}_{N}\cup\mathcal{D}^{X^{i}}_{-N}\big{)}\cap\big{(}\mathcal{D}^{X^{j}}_{N}\cup\mathcal{D}^{X^{j}}_{-N}\big{)}.

For zz outside Range(X)Range(Xpl)\operatorname{Range}(X)\cup\operatorname{Range}(X^{pl}), we have

𝐧X(z)=i=0T1nXi(z)+𝐧Xpl(z),|𝐧Xpl(z)|T.{\mathbf{n}}_{X}(z)=\sum_{i=0}^{T-1}n_{X^{i}}(z)+{\mathbf{n}}_{X^{pl}}(z),\quad|{\mathbf{n}}_{X^{pl}}(z)|\leq T.

It follows444See Section 3.2 in [11] for more details. that, for all T,M,N1T,M,N\geq 1 such that T(M+1)<NT(M+1)<N,

𝒟NXi=0T1𝒟NTM(T1)ii,j=0ijT1𝒟Mi,jRange(X)Range(Xpl),\mathcal{D}^{X}_{N}\subseteq\bigcup_{i=0}^{T-1}\mathcal{D}^{i}_{N-T-M(T-1)}\cup\bigcup_{\begin{subarray}{c}i,j=0\\ i\neq j\end{subarray}}^{T-1}\mathcal{D}^{i,j}_{M}\cup\operatorname{Range}(X)\cup\operatorname{Range}(X^{pl}),

and therefore

f(𝒟NX)i=0T1f(𝒟NTM(T1)i)+i,j=0ijT1f(𝒟Mi,j).f(\mathcal{D}^{X}_{N})\leq\sum_{i=0}^{T-1}f(\mathcal{D}^{i}_{N-T-M(T-1)})+\sum_{\begin{subarray}{c}i,j=0\\ i\neq j\end{subarray}}^{T-1}f(\mathcal{D}^{i,j}_{M}).

We set t(0,13)t\in(0,\tfrac{1}{3}), m(1+t2,1t)m\in(\tfrac{1+t}{2},1-t), α<12\alpha<\frac{1}{2}, T=NtT=\lfloor N^{t}\rfloor, M=NmM=\lfloor N^{m}\rfloor, and we assume that NN is large enough for the inequality T(M+1)<NT(M+1)<N to hold. We also set N=NTM(T1)N^{\prime}=N-T-M(T-1) to ease notations.

Using the fact that 𝒟Ni\mathcal{D}^{i}_{N^{\prime}} is contained inside the convex hull of XiX^{i}, hence in the ball centered at XiTX_{\frac{i}{T}} with radius X𝒞αTα\|X\|_{\mathcal{C}^{\alpha}}T^{-\alpha}, we deduce that ff is bounded above by f(XiT)+|f|𝒞ϵX𝒞αϵTϵαf(X_{\frac{i}{T}})+|f|_{\mathcal{C}^{\epsilon}}\|X\|^{\epsilon}_{\mathcal{C}^{\alpha}}T^{-\epsilon\alpha} on 𝒟Ni\mathcal{D}^{i}_{N^{\prime}}. Thus,

f(𝒟N)\displaystyle f(\mathcal{D}_{N}) i=0T1f(X(iT))|𝒟Ni|+|f|𝒞ϵX𝒞αϵTϵαi=0T1|𝒟Ni|+fij|𝒟Mi,j|.\displaystyle\leq\sum_{i=0}^{T-1}f(X(\tfrac{i}{T}))|\mathcal{D}^{i}_{N^{\prime}}|+|f|_{\mathcal{C}^{\epsilon}}\|X\|^{\epsilon}_{\mathcal{C}^{\alpha}}T^{-\epsilon\alpha}\sum_{i=0}^{T-1}|\mathcal{D}^{i}_{N^{\prime}}|+\|f\|_{\infty}\sum_{i\neq j}|\mathcal{D}^{i,j}_{M}|.
12πNTi=0T1f(X(iT))+fi=0T1|12πNT|𝒟Ni||+|f|𝒞ϵX𝒞αϵTϵαi=0T1|𝒟Ni|\displaystyle\leq\frac{1}{2\pi NT}\sum_{i=0}^{T-1}f(X(\tfrac{i}{T}))+\|f\|_{\infty}\sum_{i=0}^{T-1}\big{|}\tfrac{1}{2\pi NT}-|\mathcal{D}^{i}_{N^{\prime}}|\big{|}+|f|_{\mathcal{C}^{\epsilon}}\|X\|^{\epsilon}_{\mathcal{C}^{\alpha}}T^{-\epsilon\alpha}\sum_{i=0}^{T-1}|\mathcal{D}^{i}_{N^{\prime}}|
+fij|𝒟Mi,j|\displaystyle\hskip 310.13486pt+\|f\|_{\infty}\sum_{i\neq j}|\mathcal{D}^{i,j}_{M}|
12πN01f(Xt)dt+|f|𝒞ϵX𝒞αϵTϵα2πN+fi=0T1|12πNT|𝒟Ni||\displaystyle\leq\frac{1}{2\pi N}\int_{0}^{1}f(X_{t}){\mathop{}\!\mathrm{d}}t+\frac{|f|_{\mathcal{C}^{\epsilon}}\|X\|_{\mathcal{C}^{\alpha}}^{\epsilon}T^{-\epsilon\alpha}}{2\pi N}+\|f\|_{\infty}\sum_{i=0}^{T-1}\big{|}\tfrac{1}{2\pi NT}-|\mathcal{D}^{i}_{N^{\prime}}|\big{|}
+|f|𝒞ϵX𝒞αϵTϵαi=0T1|𝒟Ni|+fij|𝒟Mi,j|.\displaystyle\hskip 170.71652pt+|f|_{\mathcal{C}^{\epsilon}}\|X\|^{\epsilon}_{\mathcal{C}^{\alpha}}T^{-\epsilon\alpha}\sum_{i=0}^{T-1}|\mathcal{D}^{i}_{N^{\prime}}|+\|f\|_{\infty}\sum_{i\neq j}|\mathcal{D}^{i,j}_{M}|.

Writing (f)+p(f)^{p}_{+} for the positive part of ff, to the power pp, and using the triangle inequality in Lp()L^{p}(\mathbb{P}), we obtain

𝔼[(f(𝒟N)\displaystyle\mathbb{E}\Big{[}\Big{(}f(\mathcal{D}_{N})- 12πN01f(Xt)dt)+p]1p|f|𝒞ϵTϵα2πN𝔼[X𝒞αϵp]1p+f𝔼[|12πN|𝒟N||p]1p\displaystyle\frac{1}{2\pi N}\int_{0}^{1}f(X_{t}){\mathop{}\!\mathrm{d}}t\Big{)}_{+}^{p}\Big{]}^{\frac{1}{p}}\leq\frac{|f|_{\mathcal{C}^{\epsilon}}T^{-\epsilon\alpha}}{2\pi N}\mathbb{E}[\|X\|_{\mathcal{C}^{\alpha}}^{\epsilon p}]^{\frac{1}{p}}+\|f\|_{\infty}\mathbb{E}\Big{[}\Big{|}\frac{1}{2\pi N}-|\mathcal{D}_{N^{\prime}}|\Big{|}^{p}\Big{]}^{\frac{1}{p}}
+|f|𝒞ϵTϵα𝔼[|𝒟N|2p]12p𝔼[X𝒞α2pϵ]12p+f𝔼[(ij|𝒟Mi,j|)p]1p.\displaystyle+|f|_{\mathcal{C}^{\epsilon}}T^{-\epsilon\alpha}\mathbb{E}[|\mathcal{D}_{N^{\prime}}|^{2p}]^{\frac{1}{2p}}\mathbb{E}[\|X\|_{\mathcal{C}^{\alpha}}^{2p\epsilon}]^{\frac{1}{2p}}+\|f\|_{\infty}\mathbb{E}\Big{[}\Big{(}\sum_{i\neq j}|\mathcal{D}^{i,j}_{M}|\Big{)}^{p}\Big{]}^{\frac{1}{p}}.

We now use the asymptotic equivalence NNNN^{\prime}\sim_{N\to\infty}N and 1N1NNNt+m2\frac{1}{N}-\frac{1}{N^{\prime}}\sim_{N\to\infty}N^{t+m-2}, as well as Lemma 3.4, and the following estimations ([11, Lemma 2.4]): for all p1p\geq 1, there exists a constant CC such that for all N1N\geq 1,

𝔼[(ij|𝒟Mi,j|)p]1pClog(N+1)3+2pM2T11p.\mathbb{E}\Big{[}\Big{(}\sum_{i\neq j}|\mathcal{D}^{i,j}_{M}|\Big{)}^{p}\Big{]}^{\frac{1}{p}}\leq C\log(N+1)^{3+\frac{2}{p}}M^{-2}T^{1-\frac{1}{p}}.

We end up with

𝔼[(f(𝒟N)12πN01f(Xt)dt)+p]1p\displaystyle\mathbb{E}\Big{[}\Big{(}f(\mathcal{D}_{N})-\frac{1}{2\pi N}\int_{0}^{1}f(X_{t}){\mathop{}\!\mathrm{d}}t\Big{)}_{+}^{p}\Big{]}^{\frac{1}{p}} C(|f|𝒞ϵN1tϵα+fNm+t2+fN1δ\displaystyle\leq C\big{(}|f|_{\mathcal{C}^{\epsilon}}N^{-1-t\epsilon\alpha}+\|f\|_{\infty}N^{m+t-2}+\|f\|_{\infty}N^{-1-\delta}
+|f|𝒞ϵN1tϵα+flog(N+1)3+2pN2m+ttp),\displaystyle+|f|_{\mathcal{C}^{\epsilon}}N^{-1-t\epsilon\alpha}+\|f\|_{\infty}\log(N+1)^{3+\frac{2}{p}}N^{-2m+t-\frac{t}{p}}\big{)},

for an arbitrary but fixed δ(0,12)\delta\in(0,\frac{1}{2}). The conditions on tt and mm ensures that all the exponents of NN are smaller than 1-1, so that there exists δ\delta^{\prime} and CC such that

𝔼[(f(𝒟N)12πN01f(Xt)dt)+p]1pCf𝒞bαN1δ.\mathbb{E}\Big{[}\Big{(}f(\mathcal{D}_{N})-\frac{1}{2\pi N}\int_{0}^{1}f(X_{t}){\mathop{}\!\mathrm{d}}t\Big{)}_{+}^{p}\Big{]}^{\frac{1}{p}}\leq C\|f\|_{\mathcal{C}_{b}^{\alpha}}N^{-1-\delta^{\prime}}.

The negative part is treated in a similar way, and the lemma follows. ∎

Corollary 4.6.

Let ϵ>0\epsilon>0 and p1p\geq 1. There exists a constant CC such that for all f𝒞bϵ(2)f\in\mathcal{C}_{b}^{\epsilon}(\mathbb{R}^{2}), 𝔼[(𝐧X(z)f(z)dz)p]1pCf𝒞bϵ\mathbb{E}[(\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X}(z)f(z){\mathop{}\!\mathrm{d}}z)^{p}]^{\frac{1}{p}}\leq C\|f\|_{\mathcal{C}_{b}^{\epsilon}}.

Proof.

Let CC and δ\delta be the constants of Lemma 4.5. Then, for all f𝒞bϵf\in\mathcal{C}_{b}^{\epsilon} and nn,

𝔼[|f(𝒟n)f(𝒟n)|p]1p2Cn1δf𝒞bϵ.\mathbb{E}[|f(\mathcal{D}_{n})-f(\mathcal{D}_{-n})|^{p}]^{\frac{1}{p}}\leq 2Cn^{-1-\delta}\|f\|_{\mathcal{C}_{b}^{\epsilon}}.

By triangle inequality in LpL^{p},

𝔼[|n=1(f(𝒟n)f(𝒟N))|p]1p2Cf𝒞bϵn=1N1δCf𝒞bϵ,\mathbb{E}\Big{[}\Big{|}\sum_{n=1}^{\infty}(f(\mathcal{D}_{n})-f(\mathcal{D}_{-N}))\Big{|}^{p}\Big{]}^{\frac{1}{p}}\leq 2C\|f\|_{\mathcal{C}_{b}^{\epsilon}}\sum_{n=1}^{\infty}N^{-1-\delta}\leq C^{\prime}\|f\|_{\mathcal{C}_{b}^{\epsilon}},

as expected. ∎

With this estimation in hand, we can now prove Lemma 4.3.

Proof of Lemma 4.3.

For i{0,,2n1}i\in\{0,\dots,2^{n}-1\}, we define f¯i:2\bar{f}^{i}:{\mathbb{R}}^{2}\to{\mathbb{R}} the constant function whose unique value is equal to f(Xi2n)f(X_{i2^{-n}}), and f~i=ff¯i\tilde{f}^{i}=f-\bar{f}^{i}. Since for all ii, fXi𝐧X(z)f(z)dzf\mapsto\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{X_{i}}{\mathbf{n}}_{X}(z)f(z){\mathop{}\!\mathrm{d}}z is linear, it suffices to show that both

i=12nXi𝐧X(z)f¯i(z)dz=i=12nf(Xi2n)Xi𝐧X(z)dz and i=12nXi𝐧X(z)f~i(z)dz\sum_{i=1}^{2^{n}}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{X^{i}}{\mathbf{n}}_{X}(z)\bar{f}^{i}(z){\mathop{}\!\mathrm{d}}z=\sum_{i=1}^{2^{n}}f(X_{i2^{-n}})\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{X^{i}}{\mathbf{n}}_{X}(z){\mathop{}\!\mathrm{d}}z\qquad\mbox{ and }\qquad\sum_{i=1}^{2^{n}}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{X^{i}}{\mathbf{n}}_{X}(z)\tilde{f}^{i}(z){\mathop{}\!\mathrm{d}}z

almost surely converge toward 0 as nn\to\infty.

From symmetry, for all ii, 𝔼[Xi𝐧X(z)dz|(Xs)si2n]=0\mathbb{E}\Big{[}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{X_{i}}{\mathbf{n}}_{X}(z){\mathop{}\!\mathrm{d}}z|(X_{s})_{s\leq\tfrac{i}{2^{n}}}\Big{]}=0. It follows that, for i<ji<j,

𝔼[f(Xi2n)f(Xj2n)Xi𝐧X(z)dzXj𝐧X(z)dz]=0.\mathbb{E}\Big{[}f(X_{i2^{-n}})f(X_{j2^{-n}})\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{X_{i}}{\mathbf{n}}_{X}(z){\mathop{}\!\mathrm{d}}z\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{X_{j}}{\mathbf{n}}_{X}(z){\mathop{}\!\mathrm{d}}z\Big{]}=0.

Besides, from a simple scaling argument,

𝔼[(Xi𝐧X(z)dz)2]=22n𝔼[(X𝐧X(z)dz)2].\mathbb{E}\Big{[}\Big{(}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{X^{i}}{\mathbf{n}}_{X}(z){\mathop{}\!\mathrm{d}}z\Big{)}^{2}\Big{]}=2^{-2n}\mathbb{E}\Big{[}\Big{(}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{X}{\mathbf{n}}_{X}(z){\mathop{}\!\mathrm{d}}z\Big{)}^{2}\Big{]}.

Notice 𝔼[(X𝐧X(z)dz)2]<\mathbb{E}[(\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{X}{\mathbf{n}}_{X}(z){\mathop{}\!\mathrm{d}}z)^{2}]<\infty, which follows for example from the previous corollary.

We deduce that

𝔼[(i=12nXi𝐧X(z)f¯i(z)dz)2]\displaystyle\mathbb{E}\Big{[}\Big{(}\sum_{i=1}^{2^{n}}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{X^{i}}{\mathbf{n}}_{X}(z)\bar{f}^{i}(z){\mathop{}\!\mathrm{d}}z\Big{)}^{2}\Big{]} =i=12n𝔼[f(Xi2n)2(Xi𝐧X(z)dz)2]\displaystyle=\sum_{i=1}^{2^{n}}\mathbb{E}\Big{[}f(X_{i2^{-n}})^{2}\Big{(}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{X^{i}}{\mathbf{n}}_{X}(z){\mathop{}\!\mathrm{d}}z\Big{)}^{2}\Big{]}
2nf2𝔼[(X𝐧X(z)dz)2].\displaystyle\leq 2^{-n}\|f\|^{2}_{\infty}\mathbb{E}\Big{[}\Big{(}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{X}{\mathbf{n}}_{X}(z){\mathop{}\!\mathrm{d}}z\Big{)}^{2}\Big{]}.

This L2L^{2} convergence rate is sufficient to conclude to the almost sure convergence: for all ϵ>0\epsilon^{\prime}>0,

(nn0:|i=12nXi𝐧X(z)f¯i(z)dz|ϵ)\displaystyle\mathbb{P}\Big{(}\exists n\geq n_{0}:\Big{|}\sum_{i=1}^{2^{n}}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{X^{i}}{\mathbf{n}}_{X}(z)\bar{f}^{i}(z){\mathop{}\!\mathrm{d}}z\Big{|}\geq\epsilon^{\prime}\Big{)} 1ϵ2𝔼[supnn0(i=12nXi𝐧X(z)f¯i(z)dz)2]\displaystyle\leq\frac{1}{\epsilon^{\prime 2}}\mathbb{E}\Big{[}\sup_{n\geq n_{0}}\Big{(}\sum_{i=1}^{2^{n}}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{X^{i}}{\mathbf{n}}_{X}(z)\bar{f}^{i}(z){\mathop{}\!\mathrm{d}}z\Big{)}^{2}\Big{]}
21n0ϵ2f2𝔼[(X𝐧X(z)dz)2]\displaystyle\leq\frac{2^{1-n_{0}}}{\epsilon^{\prime 2}}\|f\|^{2}_{\infty}\mathbb{E}\Big{[}\Big{(}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{X}{\mathbf{n}}_{X}(z){\mathop{}\!\mathrm{d}}z\Big{)}^{2}\Big{]}
n00.\displaystyle\underset{n_{0}\to\infty}{\longrightarrow}0.

In order to deal with the sum involving f~i\tilde{f}^{i}, one must be a bit careful about the way we use the translation invariance and scale invariance of the Brownian motion. We set α<12\alpha<\frac{1}{2} and we define the event

={X𝒞αR},\mathcal{R}=\{\|X\|_{\mathcal{C}^{\alpha}}\leq R\},

for a fixed R1R\geq 1. Let f^i\hat{f}^{i} be the (random) function defined by

f^i(Xi2n+z)={f~i(Xi2n+z)if |z|R2αn,f~i(Xi2n+R2αn|z|z)otherwise.\hat{f}^{i}(X_{i2^{-n}}+z)=\left\{\begin{array}[]{ll}\tilde{f}^{i}(X_{i2^{-n}}+z)&\mbox{if }|z|\leq R2^{-\alpha n},\\ \tilde{f}^{i}(X_{i2^{-n}}+\frac{R2^{-\alpha n}}{|z|}z)&\mbox{otherwise.}\end{array}\right.

In particular, f^i\hat{f}^{i} satisfies the following properties:

  • \scriptstyle\diamond

    f^i=f~i\hat{f}^{i}=\tilde{f}^{i} on B=B(Xi2n,R2αn)B=B(X_{i2^{-n}},R2^{-\alpha n}), so that, in the event \mathcal{R}, f^i(𝒟ni)=f~i(𝒟ni)\hat{f}^{i}(\mathcal{D}^{i}_{n})=\tilde{f}^{i}(\mathcal{D}^{i}_{n}),

  • \scriptstyle\diamond

    |f^i|𝒞ϵ|f|𝒞ϵ|\hat{f}^{i}|_{\mathcal{C}^{\epsilon}}\leq|f|_{\mathcal{C}^{\epsilon}}, and f^iRϵ2ϵαn|f|𝒞ϵ\|\hat{f}^{i}\|_{\infty}\leq R^{\epsilon}2^{-\epsilon\alpha n}|f|_{\mathcal{C}^{\epsilon}},

  • \scriptstyle\diamond

    As a random variable, f^i\hat{f}^{i} is measurable with respect to σ(Xi2n)\sigma(X_{i2^{-n}}).

Set also fˇi(z)=f^i(Xi2n+2n2z)\check{f}^{i}(z)=\hat{f}^{i}(X_{i2^{-n}}+2^{-\frac{n}{2}}z), Xˇi:s[0,1]2n2(X(i+s)2nXi2n)\check{X}^{i}:s\in[0,1]\mapsto 2^{\frac{n}{2}}(X_{(i+s)2^{-n}}-X_{i2^{-n}}), which is a standard planar Brownian motion started from 0, independent from Xi2nX_{i2^{-n}}. Notice that fˇi=f^iRϵ2ϵαn|f|𝒞ϵ\|\check{f}^{i}\|_{\infty}=\|\hat{f}^{i}\|_{\infty}\leq R^{\epsilon}2^{-\epsilon\alpha n}|f|_{\mathcal{C}^{\epsilon}} and |fˇi|𝒞ϵ=2ϵn2|f^i|𝒞ϵ2ϵn2|f|𝒞ϵ|\check{f}^{i}|_{\mathcal{C}^{\epsilon}}=2^{-\frac{\epsilon n}{2}}|\hat{f}^{i}|_{\mathcal{C}^{\epsilon}}\leq 2^{-\frac{\epsilon n}{2}}|f|_{\mathcal{C}^{\epsilon}}, so that

fˇi𝒞bϵ21ϵαn|f|𝒞ϵ.\|\check{f}^{i}\|_{\mathcal{C}_{b}^{\epsilon}}\leq 2^{1-\epsilon\alpha n}|f|_{\mathcal{C}^{\epsilon}}.

On the event \mathcal{R}, we have

𝐧Xi(z)f~i(z)dz=2n𝐧Xˇi(w)fˇi(2n2w)dw.\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X^{i}}(z)\tilde{f}^{i}(z){\mathop{}\!\mathrm{d}}z=2^{-n}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{\check{X}^{i}}(w)\check{f}^{i}(2^{-\frac{n}{2}}w){\mathop{}\!\mathrm{d}}w.

Using Corollary 4.6 with p=1p=1, we deduce

𝔼[𝟙|𝐧Xi(z)f~i(z)dz|]\displaystyle\mathbb{E}\Big{[}\mathbbm{1}_{\mathcal{R}}\Big{|}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X^{i}}(z)\tilde{f}^{i}(z){\mathop{}\!\mathrm{d}}z\Big{|}\Big{]} =2n𝔼[𝔼[|𝐧Xˇi(w)fˇi(2n2w)dw||Xi2n]]\displaystyle=2^{-n}\mathbb{E}\left[\mathbb{E}\left[\big{|}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{\check{X}^{i}}(w)\check{f}^{i}(2^{-\frac{n}{2}}w){\mathop{}\!\mathrm{d}}w\big{|}\middle|X_{i2^{-n}}\right]\right]
2n𝔼[Cfˇi𝒞bϵ]\displaystyle\leq 2^{-n}\mathbb{E}\big{[}C\|\check{f}^{i}\|_{\mathcal{C}^{\epsilon}_{b}}\big{]}
C21nϵαn|f|𝒞ϵ.\displaystyle\leq C2^{1-n-\epsilon\alpha n}|f|_{\mathcal{C}^{\epsilon}}.

Thus,

( and nn0:|i=02n1𝐧Xi(z)f~i(z)dz|ϵ)\displaystyle\mathbb{P}\Big{(}\mathcal{R}\mbox{ and }\exists n\geq n_{0}:\Big{|}\sum_{i=0}^{2^{n}-1}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X^{i}}(z)\tilde{f}^{i}(z){\mathop{}\!\mathrm{d}}z\Big{|}\geq\epsilon^{\prime}\Big{)} 1ϵn=n0i=02n1𝔼[𝟙|𝐧Xi(z)f~i(z)dz|]\displaystyle\leq\frac{1}{\epsilon^{\prime}}\sum_{n=n_{0}}^{\infty}\sum_{i=0}^{2^{n}-1}\mathbb{E}\Big{[}\mathbbm{1}_{\mathcal{R}}\Big{|}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X^{i}}(z)\tilde{f}^{i}(z){\mathop{}\!\mathrm{d}}z\Big{|}\Big{]}
Cϵ,ϵ,α,R2ϵαn0|f|𝒞ϵ\displaystyle\leq C_{\epsilon,\epsilon^{\prime},\alpha,R}2^{-\epsilon\alpha n_{0}}|f|_{\mathcal{C}^{\epsilon}}
n00.\displaystyle\underset{n_{0}\to\infty}{\longrightarrow}0.

Since this holds for all RR, we deduce that i=02n1𝐧Xi(z)f~i(z)dz\sum_{i=0}^{2^{n}-1}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X^{i}}(z)\tilde{f}^{i}(z){\mathop{}\!\mathrm{d}}z almost surely converges toward 0 as nn\to\infty, which concludes the proof. ∎

4.5. Stratonovich integral as a limit of integrals along piecewise-linear paths

It only remains to prove lemma 4.4 which for η𝒞1+ϵ(T2)\eta\in\mathcal{C}^{1+\epsilon}(T^{*}\mathbb{R}^{2}) identifies the limit

limnX(n)η\lim_{n\to\infty}\int_{X^{(n)}}\eta

with the Stratonovich integral of η\eta along XX, which is fairly classical. It is for example a direct consequence of the following lemma.

Lemma 4.7.

For a given dissection Δ=(t0=0,t1,,tn=1)\Delta=(t_{0}=0,t_{1},\dots,t_{n}=1), and X:[0,1]2X:[0,1]\to\mathbb{R}^{2} a Brownian motion, let XΔX_{\Delta} be the piecewise-linear approximation of XX associated with Δ\Delta: for λ[0,1]\lambda\in[0,1] and t=λti+(1λ)ti+1t=\lambda t_{i}+(1-\lambda)t_{i+1},

XΔ(t)=λXti+(1λ)Xti+1.X_{\Delta}(t)=\lambda X_{t_{i}}+(1-\lambda)X_{t_{i+1}}.

For f𝒞1(2)f\in\mathcal{C}^{1}(\mathbb{R}^{2}), let

IΔ1(f)\displaystyle I^{1}_{\Delta}(f) =[ti,ti+1]Δf(Xti+1+Xti2)(X1(ti+1)X1(ti)),\displaystyle=\sum_{[t_{i},t_{i+1}]\in\Delta}f\Big{(}\frac{X_{t_{i+1}}+X_{t_{i}}}{2}\Big{)}(X^{1}(t_{i+1})-X^{1}(t_{i})),
IΔ2(f)\displaystyle I^{2}_{\Delta}(f) =[ti,ti+1]Δf(Xti+1)+f(Xti)2(X1(ti+1)X1(ti)),\displaystyle=\sum_{[t_{i},t_{i+1}]\in\Delta}\frac{f(X_{t_{i+1}})+f(X_{t_{i}})}{2}(X^{1}(t_{i+1})-X^{1}(t_{i})),
IΔ3(f)\displaystyle I^{3}_{\Delta}(f) =01f(XΔ(t))dXΔ(t).\displaystyle=\int_{0}^{1}f(X_{\Delta}(t)){\mathop{}\!\mathrm{d}}X_{\Delta}(t).

Then, almost surely, for all f𝒞1+ϵ(2)f\in\mathcal{C}^{1+\epsilon}(\mathbb{R}^{2}), as |Δ|0|\Delta|\to 0,

IΔ2(f)IΔ1(f)0andIΔ3(f)IΔ1(f)0.I^{2}_{\Delta}(f)-I^{1}_{\Delta}(f)\to 0\quad\mbox{and}\quad I^{3}_{\Delta}(f)-I^{1}_{\Delta}(f)\to 0.
Proof.

Let α(12+ϵ,12)\alpha\in\big{(}\tfrac{1}{2+\epsilon},\tfrac{1}{2}\big{)}. On the almost sure event X𝒞α<\|X\|_{\mathcal{C}^{\alpha}}<\infty, we have

|[ti,ti+1]Δ(f(Xti+1)+f(Xti)2f(Xti+1+Xti2))(Xti+11Xti1)|\displaystyle\Big{|}\sum_{[t_{i},t_{i+1}]\in\Delta}\Big{(}\frac{f(X_{t_{i+1}})+f(X_{t_{i}})}{2}-f\Big{(}\frac{X_{t_{i+1}}+X_{t_{i}}}{2}\Big{)}\Big{)}(X^{1}_{t_{i+1}}-X^{1}_{t_{i}})\Big{|}
[ti,ti+1]Δ12|f(Xti+1)f(Xti+1+Xti2)+f(Xti)f(Xti+1+Xti2)||Xti+11Xti1|\displaystyle\leq\sum_{[t_{i},t_{i+1}]\in\Delta}\frac{1}{2}\Big{|}f(X_{t_{i+1}})-f\Big{(}\frac{X_{t_{i+1}}+X_{t_{i}}}{2}\Big{)}+f(X_{t_{i}})-f\Big{(}\frac{X_{t_{i+1}}+X_{t_{i}}}{2}\Big{)}\Big{|}\Big{|}X^{1}_{t_{i+1}}-X^{1}_{t_{i}}\Big{|}
[ti,ti+1]Δ14|Xti+1Xtif(Xti+1+Xti2)+XtiXti+1f(Xti+1+Xti2)=0||Xti+11Xti1|\displaystyle\leq\sum_{[t_{i},t_{i+1}]\in\Delta}\frac{1}{4}\Big{|}\underbrace{\nabla_{X_{t_{i+1}}-X_{t_{i}}}f\Big{(}\frac{X_{t_{i+1}}+X_{t_{i}}}{2}\Big{)}+\nabla_{X_{t_{i}}-X_{t_{i+1}}}f\Big{(}\frac{X_{t_{i+1}}+X_{t_{i}}}{2}\Big{)}}_{=0}\Big{|}\Big{|}X^{1}_{t_{i+1}}-X^{1}_{t_{i}}\Big{|}
+[ti,ti+1]Δ222+ϵf𝒞ϵ|Xti+1+Xti|2+ϵ\displaystyle\hskip 28.45274pt+\sum_{[t_{i},t_{i+1}]\in\Delta}\frac{2}{2^{2+\epsilon}}\|\nabla f\|_{\mathcal{C}^{\epsilon}}|X_{t_{i+1}}+X_{t_{i}}|^{2+\epsilon}
21ϵf𝒞1+ϵX𝒞α2+ϵ[ti,ti+1]Δ|ti+1ti|α(2+ϵ)|Δ|00.\displaystyle\leq 2^{-1-\epsilon}\|f\|_{\mathcal{C}^{1+\epsilon}}\|X\|_{\mathcal{C}^{\alpha}}^{2+\epsilon}\sum_{[t_{i},t_{i+1}]\in\Delta}|t_{i+1}-t_{i}|^{\alpha(2+\epsilon)}\underset{|\Delta|\to 0}{\longrightarrow}0.

The second convergence is proved in a similar way:

|[ti,ti+1]Δ(titi+1f(XΔ(s))dXΔ(s)f(Xti+1+Xti2)(Xti+11Xti1))|\displaystyle\Big{|}\sum_{[t_{i},t_{i+1}]\in\Delta}\Big{(}\int_{t_{i}}^{t_{i+1}}f(X_{\Delta}(s)){\mathop{}\!\mathrm{d}}X_{\Delta}(s)-f\Big{(}\frac{X_{t_{i+1}}+X_{t_{i}}}{2}\Big{)}(X^{1}_{t_{i+1}}-X^{1}_{t_{i}})\Big{)}\Big{|}
[ti,ti+1]Δ|Xti+11Xti1|\displaystyle\leq\sum_{[t_{i},t_{i+1}]\in\Delta}|X^{1}_{t_{i+1}}-X^{1}_{t_{i}}|
|121(f(λXti+(1λ)Xti+1)+f((1λ)Xti+λXti+1)2f(Xti+1+Xti2))dλ|\displaystyle\hskip 28.45274pt\Big{|}\int_{\frac{1}{2}}^{1}\Big{(}f(\lambda X_{t_{i}}+(1-\lambda)X_{t_{i+1}})+f((1-\lambda)X_{t_{i}}+\lambda X_{t_{i+1}})-2f\Big{(}\frac{X_{t_{i+1}}+X_{t_{i}}}{2}\Big{)}\Big{)}{\mathop{}\!\mathrm{d}}\lambda\Big{|}
21ϵf𝒞1+ϵX𝒞α2+ϵ[ti,ti+1]Δ|ti+1ti|α(2+ϵ)|Δ|00.\displaystyle\leq 2^{-1-\epsilon}\|f\|_{\mathcal{C}^{1+\epsilon}}\|X\|_{\mathcal{C}^{\alpha}}^{2+\epsilon}\sum_{[t_{i},t_{i+1}]\in\Delta}|t_{i+1}-t_{i}|^{\alpha(2+\epsilon)}\underset{|\Delta|\to 0}{\longrightarrow}0.

This concludes the proof of Lemma 4.4, and therefore the proof of Theorem 1 as well. Before we conclude this section, we will shortly prove Corollary 2.

Proof of Corollary 2.

To keep the proof simple, we treat the case when X:[0,1]2X:[0,1]\to\mathbb{R}^{2} is a Brownian loop started from 0. To deal with the case when XX is a Brownian bridge from xx to yxy\neq x, one must also take into account the winding function of the triangle between xx, yy, and X12X_{\frac{1}{2}}, but this is done in a straightforward way.

From linearity, it suffices to prove the result when restricted to functions f0f\geq 0. Furthermore, since the result is trivial in the event f|B(0,X)=0f_{|{B(0,\|X\|_{\infty})}}=0, we assume B(0,X)f(z)dz>0\int_{B(0,\|X\|_{\infty})}f(z){\mathop{}\!\mathrm{d}}z>0.

Let X1X^{1} be the restriction of XX to [0,12][0,\tfrac{1}{2}] , X2X^{2} its restriction to [12,1][\tfrac{1}{2},1], and X^2:t[0,12]X1t\hat{X}_{2}:t\in[0,\tfrac{1}{2}]\mapsto X_{1-t}. Then, the distribution of X1X^{1} (resp. X^2\hat{X}^{2}) admits a density with respect to the density of a standard planar Brownian motion defined on [0,12][0,\tfrac{1}{2}]. Using scale invariance, we can apply Theorem 1 to both X1X^{1} and X^2\hat{X}^{2}. We deduce that, for i{1,2}i\in\{1,2\}, for all ϵ>0\epsilon>0, almost surely, for all f𝒞ϵ(2)f\in\mathcal{C}^{\epsilon}(\mathbb{R}^{2}),

2[𝐧Xi(z)]kf(z)dz\int_{\mathbb{R}^{2}}[{\mathbf{n}}_{X^{i}}(z)]_{k}f(z){\mathop{}\!\mathrm{d}}z

converges as kk\to\infty, and the limits are almost surely equal to respectively X1η+[X12,0]η\int_{X^{1}}\eta+\int_{[X_{\frac{1}{2}},0]}\eta and X2η[X12,0]η\int_{X^{2}}\eta-\int_{[X_{\frac{1}{2}},0]}\eta, where η\eta is such that 1η22η1=f\partial_{1}\eta_{2}-\partial_{2}\eta_{1}=f.

Now we need to show that almost surely, for all f𝒞ϵ(2)f\in\mathcal{C}^{\epsilon}(\mathbb{R}^{2}), 𝐧X1(z)f(z)dz\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X^{1}}(z)f(z){\mathop{}\!\mathrm{d}}z and 𝐧X2(z)f(z)dz\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X^{2}}(z)f(z){\mathop{}\!\mathrm{d}}z add up properly, for which we proceed as in Lemma 4.2, introducing again a random point PP. Going through the same arguments as in the proof of Lemma 4.2, we see that it suffices to show that, XX-almost surely,

|𝒟±NX1𝒟±NX2|=o(N1δ),|\mathcal{D}^{X^{1}}_{\pm N}\cap\mathcal{D}^{X^{2}}_{\pm N}|=o(N^{-1-\delta}), (4)

for the four possible couple of signs in front of NN, and for some δ>0\delta>0.

To prove (4), we further decompose X1X^{1} and X2X^{2} by setting X11X^{11} (resp. X12X^{12}, X21X^{21},X22X^{22}) the restriction of XX to the interval [0,14][0,\tfrac{1}{4}] (resp. [14,12][\tfrac{1}{4},\tfrac{1}{2}], [12,34][\tfrac{1}{2},\tfrac{3}{4}], [34,1][\tfrac{3}{4},1]). Then, 𝒟±NXi𝒟±NXi1𝒟±NXi2\mathcal{D}^{X^{i}}_{\pm N}\subseteq\mathcal{D}^{X^{i1}}_{\pm N^{\prime}}\cup\mathcal{D}^{X^{i2}}_{\pm N^{\prime}} where N=N/2N^{\prime}=\lfloor N/2\rfloor.

We show that almost surely, |𝒟NX11𝒟NX21|=O(N2)|\mathcal{D}^{X^{11}}_{N^{\prime}}\cap\mathcal{D}^{X^{21}}_{N^{\prime}}|=O(N^{-2}), the 15 other intersections are treated either similarly. Conditionally on X12X_{\frac{1}{2}}, X11X^{11} and X21X^{21} are independen. Furthermore, both their distribution, conditional on X12X_{\frac{1}{2}}, have a density with respect to the distribution of a standard Brownian motion with duration 14\frac{1}{4}, started respectively from 0 and X12X_{\frac{1}{2}}. Thus, it suffices to show that for all yy, |𝒟NX11𝒟NX21|=O(N2)|\mathcal{D}^{X^{11}}_{N^{\prime}}\cap\mathcal{D}^{X^{21}}_{N^{\prime}}|=O(N^{-2}) when X11X^{11} and X21X^{21} are independent Brownian motions started respectively from 0 and yy. This follows directly from 3.5, with a scaling of 12\frac{1}{2}. ∎

5. Magnetic impurities

In this section, we fix a function g𝒞bϵ(2)g\in\mathcal{C}^{\epsilon}_{b}(\mathbb{R}^{2}). For all λ>0\lambda>0, we define 𝒫λ\mathcal{P}_{\lambda} a Poisson process on 2\mathbb{R}^{2} with intensity λg(z)dz\lambda g(z){\mathop{}\!\mathrm{d}}z, independent from XX, and Γ:[0,T]\Gamma:[0,T]\to\mathbb{R} a standard Cauchy process, independent from XX. We write 𝔼𝒫\mathbb{E}^{\mathcal{P}} the expectation with respect to 𝒫λ\mathcal{P}_{\lambda}, 𝔼X\mathbb{E}^{X} the one with respect to XX, 𝔼Γ\mathbb{E}^{\Gamma} the expectation with respect to Γ\Gamma and 𝔼=𝔼X𝔼𝒫𝔼Γ\mathbb{E}=\mathbb{E}^{X}\otimes\mathbb{E}^{\mathcal{P}}\otimes\mathbb{E}^{\Gamma} the expectation on the product space (although none of the variables we consider depend on both 𝒫\mathcal{P} and Γ\Gamma, so truly 𝔼=𝔼X𝔼𝒫\mathbb{E}=\mathbb{E}^{X}\otimes\mathbb{E}^{\mathcal{P}} or 𝔼=𝔼X𝔼Γ\mathbb{E}=\mathbb{E}^{X}\otimes\mathbb{E}^{\Gamma}, whichever is relevant).

For a function f𝒞bϵ(2)f\in\mathcal{C}_{b}^{\epsilon}(\mathbb{R}^{2}), we define

ξλ(f)=1λz𝒫λf(z)𝐧X(z),\xi_{\lambda}(f)=\frac{1}{\lambda}\sum_{z\in\mathcal{P}_{\lambda}}f(z){\mathbf{n}}_{X}(z),

as well as

ξ(f)=𝐧X(z)fg(z)dz+1201fg(Xt)dΓt.\xi(f)=\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X}(z)\ f\cdot g(z){\mathop{}\!\mathrm{d}}z+\frac{1}{2}\int_{0}^{1}f\cdot g(X_{t}){\mathop{}\!\mathrm{d}}\Gamma_{t}.

Notice that Γ\Gamma almost surely has a finite pp-variation for all p>1p>1 (see [1, Theorem 4.1]). Since XX-almost surely, (fg)X𝒞ϵ4([0,1])(f\cdot g)\circ X\in\mathcal{C}^{\frac{\epsilon}{4}}([0,1]), the integral 01fg(Xt)dΓt\int_{0}^{1}fg(X_{t}){\mathop{}\!\mathrm{d}}\Gamma_{t} is well-defined as a Young integral.

The main result from this section is the following

Lemma 5.1.

Let f,g𝒞bϵ(2)f,g\in\mathcal{C}^{\epsilon}_{b}({\mathbb{R}}^{2}) be continuous and bounded functions. Assume that gg takes non-negative values. Let

Gβ,f,gk0𝒜k(eikβf(z)1)g(z)dz.G_{\beta,f,g}\coloneqq\sum_{k\neq 0}\int_{\mathcal{A}_{k}}(e^{ik\beta f(z)}-1)g(z){\mathop{}\!\mathrm{d}}z.

Then, XX-almost surely, as β0\beta\to 0,

Gβ,f,g=β0iβ𝐧X(z)fg(z)dz|β|201|f(Xt)|g(Xt)dt+o(β).G_{\beta,f,g}\underset{\beta\to 0}{=}i\beta\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X}(z)fg(z){\mathop{}\!\mathrm{d}}z-\frac{|\beta|}{2}\int_{0}^{1}|f(X_{t})|g(X_{t}){\mathop{}\!\mathrm{d}}t+o(\beta). (5)

Before we dive into the proof of this lemma, we first explain with it implies both Theorem 3 and Corollary 4.

Lemma 5.1 implies Theorem 3 and Corollary 4.

Since the function min(|𝐧Xf|,1)\min(|{\mathbf{n}}_{X}\cdot f|,1) is integrable against the intensity measure λgdz\lambda g{\mathop{}\!\mathrm{d}}z of 𝒫λ\mathcal{P}_{\lambda}, we can use Campbell’s theorem, which gives

𝔼𝒫[eiαξλ(f)]=exp(k0𝒜k(eikαλf(z)1)λg(z)dz)=exp(λGβ,f,g),\mathbb{E}^{\mathcal{P}}[e^{i\alpha\xi_{\lambda}(f)}]=\exp\Big{(}\sum_{k\neq 0}\int_{\mathcal{A}_{k}}(e^{ik\frac{\alpha}{\lambda}f(z)}-1)\lambda g(z){\mathop{}\!\mathrm{d}}z\Big{)}=\exp(\lambda G_{\beta,f,g}),

where β=αλ\beta=\frac{\alpha}{\lambda}.

Besides, conditional on XX, 01f(Xt)g(Xt)dΓ(t)\int_{0}^{1}f(X_{t})g(X_{t}){\mathop{}\!\mathrm{d}}\Gamma(t) is a centered Cauchy random variable with scale parameter 01|f(Xt)|g(Xt)dt\int_{0}^{1}|f(X_{t})|g(X_{t}){\mathop{}\!\mathrm{d}}t, whilst 𝐧X(z)fg(z)dz\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X}(z)fg(z){\mathop{}\!\mathrm{d}}z is deterministic. It follows that

𝔼Γ[eiαξ(f)]=eiα𝐧X(z)fg(z)dz|α|201|f(Xt)|g(Xt)dt,\mathbb{E}^{\Gamma}[e^{i\alpha\xi(f)}]=e^{i\alpha\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-3.40277pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-2.275pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.60416pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.31248pt}}\!\int{\mathbf{n}}_{X}(z)fg(z){\mathop{}\!\mathrm{d}}z-\frac{|\alpha|}{2}\int_{0}^{1}|f(X_{t})|g(X_{t}){\mathop{}\!\mathrm{d}}t},

Thus, Lemma 5.1 implies Theorem 3.

Furthermore, since both ξλ(f)\xi_{\lambda}(f) and ξ(f)\xi(f) are linear in ff, one can use the Cramér-Wold device to deduce Corollary 4 from its special case n=1n=1. By Lévy’s continuity theorem, this specific case is equivalent to the statement that XX-almost surely, for all α\alpha\in\mathbb{R},

𝔼𝒫[eiαξλ(f)]λ𝔼Γ[eiαξ(f)].\mathbb{E}^{\mathcal{P}}[e^{i\alpha\xi_{\lambda}(f)}]\underset{\lambda\to\infty}{\longrightarrow}\mathbb{E}^{\Gamma}[e^{i\alpha\xi(f)}].

From our previous computation, this amount to show that XX almost surely, for all α\alpha\in\mathbb{R},

exp(λGβ,f,g)λexp(iα𝐧X(z)fg(z)dz|α|201|f(Xt)|g(Xt)dt),\exp(\lambda G_{\beta,f,g})\underset{\lambda\to\infty}{\longrightarrow}\exp\Big{(}i\alpha\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X}(z)fg(z){\mathop{}\!\mathrm{d}}z-\frac{|\alpha|}{2}\int_{0}^{1}|f(X_{t})|g(X_{t}){\mathop{}\!\mathrm{d}}t\Big{)},

which follows again from Lemma 5.1. ∎

Proof of Lemma 5.1.

From symmetry, we can assume β>0\beta>0. Performing an Abel summation, we obtain

Gβ,f,g\displaystyle G_{\beta,f,g} =k=1(𝒟keiβkf(1eiβf)gdz+𝒟keiβkf(1eiβf)gdz)\displaystyle=\sum_{k=1}^{\infty}\Big{(}\int_{\mathcal{D}_{k}}e^{i\beta kf}(1-e^{-i\beta f})g{\mathop{}\!\mathrm{d}}z+\int_{\mathcal{D}_{-k}}e^{-i\beta kf}(1-e^{i\beta f})g{\mathop{}\!\mathrm{d}}z\Big{)}
=k=1(ϕk,β(𝒟k)+ϕk,β(𝒟k)),\displaystyle=\sum_{k=1}^{\infty}(\phi_{k,\beta}(\mathcal{D}_{k})+\phi_{-k,\beta}(\mathcal{D}_{-k})),

where

ϕk,β=eiβkf(1esgn(k)iβf)g.\phi_{k,\beta}=e^{i\beta kf}(1-e^{-\operatorname{sgn}(k)i\beta f})g.

The two terms in (5) comes from two different parts in this last sum: the term iβ𝐧X(z)f(z)g(z)dzi\beta\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X}(z)f(z)g(z){\mathop{}\!\mathrm{d}}z comes from the bulk of the sum, that is the part with kk of the order of 11. The second term comes from the tail of the sum, or more precisely from the part of the sum when kk is of the order of β1\beta^{-1}. We will split the sum into several parts. For n,N{}n,N\in\mathbb{N}\cup\{\infty\} with n<Nn<N, we set

Gβ,f,gn,N=k=n+1N(ϕk,β(𝒟k)+ϕk,β(𝒟k)).G_{\beta,f,g}^{n,N}=\sum_{k=n+1}^{N}(\phi_{k,\beta}(\mathcal{D}_{k})+\phi_{-k,\beta}(\mathcal{D}_{-k})).

For N1=N1(β)N_{1}=N_{1}(\beta) and N2=N2(β)N_{2}=N_{2}(\beta) which will be set later on, we decompose Gβ,f,gG_{\beta,f,g} into three parts,

Gβ,f,g=Gβ,f,g0,N1bulk+Gβ,f,gN1,N2tail+Gβ,f,gN2,end.G_{\beta,f,g}=\underbrace{G_{\beta,f,g}^{0,N_{1}}}_{\text{bulk}}+\underbrace{G_{\beta,f,g}^{N_{1},N_{2}}}_{\text{tail}}+\underbrace{G_{\beta,f,g}^{N_{2},\infty}}_{\text{end}}.

As β0\beta\to 0, both N1N_{1} and βN2\beta N_{2} will slowly diverge toward \infty. In particular, N1(β)<<β1<<N2(β)N_{1}(\beta)<<\beta^{-1}<<N_{2}(\beta). The reason why we need to treat the end part in a separate way is that its convergence toward 0 is not absolute, in the sense that the

k=N2+1|ϕk,β(𝒟k)+ϕk,β(𝒟k)|\sum_{k=N_{2}+1}^{\infty}|\phi_{k,\beta}(\mathcal{D}_{k})+\phi_{-k,\beta}(\mathcal{D}_{-k})|

does not converge toward zero as β0\beta\to 0, and one must be a bit careful when dealing with this term. The general term (without the absolute values) slowly oscillates between positive and negative values, and we must take advantage of compensations.

For a given k0k\neq 0, as β0\beta\to 0, uniformly in zz,

ϕk,β(z)=sgn(k)iβf(z)g(z)+O(β2),\phi_{k,\beta}(z)=\operatorname{sgn}(k)i\beta f(z)g(z)+O(\beta^{2}),

and it follows that

ϕk,β(𝒟k)+ϕk,β(𝒟k)=iβ((fg)(𝒟k)(fg)(𝒟k))+O(β2).\phi_{k,\beta}(\mathcal{D}_{k})+\phi_{-k,\beta}(\mathcal{D}_{-k})=i\beta((fg)(\mathcal{D}_{k})-(fg)(\mathcal{D}_{-k}))+O(\beta^{2}).

For k1k\geq 1, let CkC_{k} be such that for all β(0,1)\beta\in(0,1),

|ϕk,β(𝒟k)+ϕk,β(𝒟k)iβ((fg)(𝒟k)(fg)(𝒟k))|Ckβ2,|\phi_{k,\beta}(\mathcal{D}_{k})+\phi_{-k,\beta}(\mathcal{D}_{-k})-i\beta((fg)(\mathcal{D}_{k})-(fg)(\mathcal{D}_{-k}))|\leq C_{k}\beta^{2},

and set N1(β)=min(β13,sup{N:kN,Ckβ13})N_{1}(\beta)=\min(\lfloor\beta^{-\frac{1}{3}}\rfloor,\sup\{N:\forall k\leq N,C_{k}\leq\beta^{-\frac{1}{3}}\}).

Then,

|Gβ,f,g0,Niβk=1N1((fg)(𝒟k)(fg)(𝒟k))|k=1N1Ckβ2β43=o(β).\Big{|}G_{\beta,f,g}^{0,N}-i\beta\sum_{k=1}^{N_{1}}((fg)(\mathcal{D}_{k})-(fg)(\mathcal{D}_{-k}))\Big{|}\leq\sum_{k=1}^{N_{1}}C_{k}\beta^{2}\leq\beta^{\frac{4}{3}}=o(\beta).

Besides, N1β0+N_{1}\underset{\beta\to 0}{\longrightarrow}+\infty, and Theorem 1 implies that

k=1N1((fg)(𝒟k)(fg)(𝒟k))β0𝐧X(z)f(z)g(z)dz.\sum_{k=1}^{N_{1}}((fg)(\mathcal{D}_{k})-(fg)(\mathcal{D}_{-k}))\underset{\beta\to 0}{\longrightarrow}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X}(z)f(z)g(z){\mathop{}\!\mathrm{d}}z.

Therefore,

Gβ,f,g0,N=iβ𝐧X(z)f(z)g(z)dz+o(β).G_{\beta,f,g}^{0,N}=i\beta\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int{\mathbf{n}}_{X}(z)f(z)g(z){\mathop{}\!\mathrm{d}}z+o(\beta). (6)

We now look at the tail part of Gβ,f,gG_{\beta,f,g}. Let δ>0\delta>0 and CC (random) be such that for all N0N\neq 0 and ϕ𝒞bϵ\phi\in\mathcal{C}^{\epsilon}_{b},

|ϕ(𝒟N)12π|N|01ϕ(Xu)du|Cϕ𝒞bϵN1δ.\Big{|}\phi(\mathcal{D}_{N})-\frac{1}{2\pi|N|}\int_{0}^{1}\phi(X_{u}){\mathop{}\!\mathrm{d}}u\Big{|}\leq C\|\phi\|_{\mathcal{C}_{b}^{\epsilon}}N^{-1-\delta}.

Recall that the existence of such a couple (δ,C)(\delta,C) is provided by Lemma 3.3. Let N2=N2(β)N_{2}=N_{2}(\beta) be any integer-valued function such that βN2β0+\beta N_{2}\underset{\beta\to 0}{\longrightarrow}+\infty and βN21δβ00\beta N_{2}^{1-\delta}\underset{\beta\to 0}{\longrightarrow}0.

For all ϕ,ψ𝒞bϵ\phi,\psi\in\mathcal{C}^{\epsilon}_{b}, |ϕψ|𝒞ϵ|ϕ|𝒞ϵψ+ϕ|ψ|𝒞ϵ|\phi\psi|_{\mathcal{C}^{\epsilon}}\leq|\phi|_{\mathcal{C}^{\epsilon}}\|\psi\|_{\infty}+\|\phi\|_{\infty}|\psi|_{\mathcal{C}^{\epsilon}}. We deduce that for all kk and β\beta,

ϕk,β\displaystyle\|\phi_{k,\beta}\|_{\infty} eiβkf1eiβfgβfg,\displaystyle\leq\|e^{i\beta kf}\|_{\infty}\|1-e^{i\beta f}\|_{\infty}\|g\|_{\infty}\leq\beta\|f\|_{\infty}\|g\|_{\infty},
|ϕk,β|𝒞ϵ\displaystyle|\phi_{k,\beta}|_{\mathcal{C}^{\epsilon}} |eiβkf|𝒞ϵ1eiβfg+eiβkf|1eiβf|𝒞ϵg+eiβkf1eiβf|g|𝒞ϵ\displaystyle\leq|e^{i\beta kf}|_{\mathcal{C}^{\epsilon}}\|1-e^{i\beta f}\|_{\infty}\|g\|_{\infty}+\|e^{i\beta kf}\|_{\infty}|1-e^{i\beta f}|_{\mathcal{C}^{\epsilon}}\|g\|_{\infty}+\|e^{i\beta kf}\|_{\infty}\|1-e^{i\beta f}\|_{\infty}|g|_{\mathcal{C}^{\epsilon}}
kβ2|f|𝒞ϵfg+β|f|𝒞ϵg+βf|g|𝒞ϵ,\displaystyle\leq k\beta^{2}|f|_{\mathcal{C}^{\epsilon}}\|f\|_{\infty}\|g\|_{\infty}+\beta|f|_{\mathcal{C}^{\epsilon}}\|g\|_{\infty}+\beta\|f\|_{\infty}|g|_{\mathcal{C}^{\epsilon}},

so that

ϕk,β𝒞bϵβ(1+kβ)(1+f𝒞bϵ)f𝒞bϵg𝒞bϵ.\|\phi_{k,\beta}\|_{\mathcal{C}^{\epsilon}_{b}}\leq\beta(1+k\beta)(1+\|f\|_{\mathcal{C}^{\epsilon}_{b}})\|f\|_{\mathcal{C}^{\epsilon}_{b}}\|g\|_{\mathcal{C}^{\epsilon}_{b}}.

We deduce that, for all k>0k>0,

|ϕk,β(𝒟k)+ϕk,β(𝒟k)12πk01(ϕk,β(Xu)+ϕk,β(Xu))du|2C(1+f𝒞ϵ)f𝒞ϵg𝒞ϵβ(1+kβ)k1δ,\Big{|}\phi_{k,\beta}(\mathcal{D}_{k})+\phi_{-k,\beta}(\mathcal{D}_{-k})-\frac{1}{2\pi k}\int_{0}^{1}(\phi_{k,\beta}(X_{u})+\phi_{-k,\beta}(X_{u})){\mathop{}\!\mathrm{d}}u\Big{|}\leq 2C(1+\|f\|_{\mathcal{C}^{\epsilon}})\|f\|_{\mathcal{C}^{\epsilon}}\|g\|_{\mathcal{C}^{\epsilon}}\beta(1+k\beta)k^{-1-\delta},

and there exists constants C=C(f,g),C′′=C′′(f,g)C^{\prime}=C^{\prime}(f,g),\ C^{\prime\prime}=C^{\prime\prime}(f,g) such that for all N2N1N_{2}\geq N_{1},

|Gβ,f,gN1,N2\displaystyle\Big{|}G_{\beta,f,g}^{N_{1},N_{2}}- 12πk=N1+1N21k01(ϕk,β(Xu)+ϕk,β(Xu))du|\displaystyle\frac{1}{2\pi}\sum_{k=N_{1}+1}^{N_{2}}\frac{1}{k}\int_{0}^{1}(\phi_{k,\beta}(X_{u})+\phi_{-k,\beta}(X_{u})){\mathop{}\!\mathrm{d}}u\Big{|}
Ck=N1+1N2β(1+kβ)k1δC′′β(N1δ+βN21δ)=o(β).\displaystyle\leq C^{\prime}\sum_{k=N_{1}+1}^{N_{2}}\beta(1+k\beta)k^{-1-\delta}\leq C^{\prime\prime}\beta(N_{1}^{-\delta}+\beta N_{2}^{1-\delta})=o(\beta).

The remaining part of the analysis is standard calculus. Set

ψk,β=eiβkfsgn(k)iβfg.\psi_{k,\beta}=e^{i\beta kf}\operatorname{sgn}(k)i\beta fg.

Then, for βf\beta\leq\|f\|_{\infty},

|k=N1+1N2ϕk,βψk,βk|\displaystyle\Big{|}\sum_{k=N_{1}+1}^{N_{2}}\frac{\phi_{k,\beta}-\psi_{k,\beta}}{k}\Big{|} =|g||k=N1+1N21keiβkf(1esgn(k)iβfsgn(k)iβf)|\displaystyle=|g|\Big{|}\sum_{k=N_{1}+1}^{N_{2}}\frac{1}{k}e^{i\beta kf}(1-e^{-\operatorname{sgn}(k)i\beta f}-\operatorname{sgn}(k)i\beta f)\Big{|}
|g|k=N1+1N21kβ2f22Cf,g|log(β)|β2=o(β).\displaystyle\leq|g|\sum_{k=N_{1}+1}^{N_{2}}\frac{1}{k}\frac{\beta^{2}f^{2}}{2}\leq C_{f,g}|\log(\beta)|\beta^{2}=o(\beta).

It follows that

Gβ,f,gN1,N2\displaystyle G_{\beta,f,g}^{N_{1},N_{2}} =12πk=N1+1N21k01(ψk,β(Xu)+ψk,β(Xu))du+o(β)\displaystyle=\frac{1}{2\pi}\sum_{k=N_{1}+1}^{N_{2}}\frac{1}{k}\int_{0}^{1}(\psi_{k,\beta}(X_{u})+\psi_{-k,\beta}(X_{u})){\mathop{}\!\mathrm{d}}u+o(\beta)
=βπk=N1+1N201f(Xu)g(Xu)sin(kβf(Xu))kdu+o(β)\displaystyle=-\frac{\beta}{\pi}\sum_{k=N_{1}+1}^{N_{2}}\int_{0}^{1}f(X_{u})g(X_{u})\frac{\sin(k\beta f(X_{u}))}{k}{\mathop{}\!\mathrm{d}}u+o(\beta)
=βπk=1N201f(Xu)g(Xu)sin(kβf(Xu))kdu+o(β).\displaystyle=-\frac{\beta}{\pi}\sum_{k=1}^{N_{2}}\int_{0}^{1}f(X_{u})g(X_{u})\frac{\sin(k\beta f(X_{u}))}{k}{\mathop{}\!\mathrm{d}}u+o(\beta).

The last line follows from the fact that

|βπk=1N101f(Xu)g(Xu)sin(kβf(Xu))kdu|f2gβ2N1=o(β).\Big{|}\frac{\beta}{\pi}\sum_{k=1}^{N_{1}}\int_{0}^{1}f(X_{u})g(X_{u})\frac{\sin(k\beta f(X_{u}))}{k}{\mathop{}\!\mathrm{d}}u\Big{|}\leq\|f\|_{\infty}^{2}\|g\|_{\infty}\beta^{2}N_{1}=o(\beta).

For s0s\leq 0, let

Φ(s)={01f(Xu)g(Xu)sin(sf(Xu))sdufor s001f(Xu)2g(Xu)dufor s=0,\Phi(s)=\left\{\begin{array}[]{ll}\int_{0}^{1}f(X_{u})g(X_{u})\frac{\sin(sf(X_{u}))}{s}{\mathop{}\!\mathrm{d}}u&\mbox{for }s\neq 0\\ \int_{0}^{1}f(X_{u})^{2}g(X_{u}){\mathop{}\!\mathrm{d}}u&\mbox{for }s=0,\end{array}\right.

so that Φ\Phi is continuous on [0,)[0,\infty) and

Gβ,f,gN1,N2=β2πk=1N2Φ(βk)+o(β).G_{\beta,f,g}^{N_{1},N_{2}}=-\frac{\beta^{2}}{\pi}\sum_{k=1}^{N_{2}}\Phi(\beta k)+o(\beta). (7)

For all R>0R>0,

|βk=1Rβ1Φ(βk)0RΦ(s)ds|βf2g+ωΦ,[0,R](β),\Big{|}\beta\sum_{k=1}^{\lfloor R\beta^{-1}\rfloor}\Phi(\beta k)-\int_{0}^{R}\Phi(s){\mathop{}\!\mathrm{d}}s\Big{|}\leq\beta\|f\|_{\infty}^{2}\|g\|_{\infty}+\omega_{\Phi,[0,R]}(\beta),

where ωΦ,[0,R](β)=sups,t[0,R]|Φ(s)Φ(t)|\omega_{\Phi,[0,R]}(\beta)=\sup_{s,t\in[0,R]}|\Phi(s)-\Phi(t)| is the continuity modulus of Φ\Phi.

Since β+ωΦ,[0,R](β)0\beta+\omega_{\Phi,[0,R]}(\beta)\to 0 for all R>0R>0, there exists a function RβR_{\beta} such that RβR_{\beta}\to\infty as β0\beta\to 0 and β+ωΦ,[0,Rβ](β)0\beta+\omega_{\Phi,[0,R_{\beta}]}(\beta)\to 0. We fix such a function, and set N2=β22δ(β1Rβ)N_{2}=\beta^{-\frac{2}{2-\delta}}\wedge(\beta^{-1}R_{\beta}). This way, we do have βN2β0+\beta N_{2}\underset{\beta\to 0}{\longrightarrow}+\infty and βN21δβ00\beta N_{2}^{1-\delta}\underset{\beta\to 0}{\longrightarrow}0.

We obtain

|βk=1N2Φ(βk)0β1N2Φ(s)ds|=o(1).\Big{|}\beta\sum_{k=1}^{N_{2}}\Phi(\beta k)-\int_{0}^{\beta^{-1}N_{2}}\Phi(s){\mathop{}\!\mathrm{d}}s\Big{|}=o(1). (8)

To estimate this last integral, there is two things we must be careful about. First, because of the sinc\operatorname{sinc} function in the definition of Φ\Phi, the function Φ\Phi is not integrable on [0,+)[0,+\infty) so we cannot naively replace the bound β1N2\beta^{-1}N_{2} with its limit. Secondly, when manipulating the integral, we must be extra careful at the vicinity of f(Xu)=0f(X_{u})=0.

Recall that for x0x\neq 0, limC0Csin(sx)sds=sgn(x)π2\lim_{C\to\infty}\int_{0}^{C}\frac{\sin(sx)}{s}{\mathop{}\!\mathrm{d}}s=\operatorname{sgn}(x)\frac{\pi}{2}. Performing an integration by part, we deduce that for all xx and C>0C>0,

|0Csin(sx)sdssgn(x)π2|\displaystyle\Big{|}\int_{0}^{C}\frac{\sin(sx)}{s}{\mathop{}\!\mathrm{d}}s-\operatorname{sgn}(x)\frac{\pi}{2}\Big{|} =|limCCCsin(sx)sds|\displaystyle=\Big{|}\lim_{C^{\prime}\to\infty}\int_{C}^{C^{\prime}}\frac{\sin(sx)}{s}{\mathop{}\!\mathrm{d}}s\Big{|}
=|cos(Cx)CxlimCCCcos(sx)s2xds|\displaystyle=\Big{|}\frac{\cos(Cx)}{Cx}-\lim_{C^{\prime}\to\infty}\int_{C}^{C^{\prime}}\frac{\cos(sx)}{s^{2}x}{\mathop{}\!\mathrm{d}}s\Big{|}
2C|x|.\displaystyle\leq\frac{2}{C|x|}.

It follows that

|0β1N2Φ(s)ds\displaystyle\Big{|}\int_{0}^{\beta^{-1}N_{2}}\Phi(s){\mathop{}\!\mathrm{d}}s π201|f(Xu)|g(Xu)du|\displaystyle-\frac{\pi}{2}\int_{0}^{1}|f(X_{u})|g(X_{u}){\mathop{}\!\mathrm{d}}u\Big{|}
=|01f(Xu)g(Xu)(0β1N2sin(sf(Xu))sdssgn(f(Xu))π2)du|\displaystyle=\Big{|}\int_{0}^{1}f(X_{u})g(X_{u})\Big{(}\int_{0}^{\beta^{-1}N_{2}}\frac{\sin(sf(X_{u}))}{s}{\mathop{}\!\mathrm{d}}s-\operatorname{sgn}(f(X_{u}))\frac{\pi}{2}\Big{)}{\mathop{}\!\mathrm{d}}u\Big{|}
01|f(Xu)|g(Xu)2β1N2|f(Xu)|du\displaystyle\leq\int_{0}^{1}|f(X_{u})|g(X_{u})\frac{2}{{\beta^{-1}N_{2}}|f(X_{u})|}{\mathop{}\!\mathrm{d}}u
=O(βN21)=o(1).\displaystyle=O(\beta N_{2}^{-1})=o(1). (9)

Combining (7), (8) and (9), we obtain

Gβ,f,gN1,N2=β201|f(Xu)|g(Xu)du+o(β).G^{N_{1},N_{2}}_{\beta,f,g}=-\frac{\beta}{2}\int_{0}^{1}|f(X_{u})|g(X_{u}){\mathop{}\!\mathrm{d}}u+o(\beta). (10)

We finally look at the end part of Gβ,f,gG_{\beta,f,g}. Since the 𝒞ϵ\mathcal{C}^{\epsilon} norm of ϕk,β\phi_{k,\beta} becomes arbitrarily large as kk goes to infinity, one cannot directly rely on Lemma 3.3. For a positive integer jj, we decompose Gβ,f,gj2N2,(j+1)2N2G^{j^{2}N_{2},(j+1)^{2}N_{2}}_{\beta,f,g} into

Gβ,f,gj2N2,(j+1)2N2\displaystyle G^{j^{2}N_{2},(j+1)^{2}N_{2}}_{\beta,f,g} =k=j2N2+1(j+1)2N2(ϕk,β(𝒟(j+1)2N2)ϕk,β(𝒟(j+1)2N2))Hβ,f,gj\displaystyle=\underbrace{\sum_{k=j^{2}N_{2}+1}^{(j+1)^{2}N_{2}}(\phi_{k,\beta}(\mathcal{D}_{(j+1)^{2}N_{2}})-\phi_{-k,\beta}(\mathcal{D}_{-(j+1)^{2}N_{2}}))}_{H^{j}_{\beta,f,g}}
+k=j2N2+1(j+1)2N2(ϕk,β(𝒟k)ϕk,β(𝒟(j+1)2N2)ϕk,β(𝒟k)+ϕk,β(𝒟(j+1)2N2)Kβ,f,gj.\displaystyle\hskip 56.9055pt+\underbrace{\sum_{k=j^{2}N_{2}+1}^{(j+1)^{2}N_{2}}(\phi_{k,\beta}(\mathcal{D}_{k})-\phi_{k,\beta}(\mathcal{D}_{(j+1)^{2}N_{2}})-\phi_{k,\beta}(\mathcal{D}_{-k})+\phi_{-k,\beta}(\mathcal{D}_{-(j+1)^{2}N_{2}})}_{K^{j}_{\beta,f,g}}.

We have

|k=j2N2+1(j+1)2N2ϕk,β(𝒟(j+1)2N2)|\displaystyle\Big{|}\sum_{k=j^{2}N_{2}+1}^{(j+1)^{2}N_{2}}\phi_{k,\beta}(\mathcal{D}_{(j+1)^{2}N_{2}})\Big{|} =|𝒟(j+1)2N2k=j2N2+1(j+1)2N2eiβkf(z)(1eiβf(z))g(z)dz|\displaystyle=\Big{|}\int_{\mathcal{D}_{(j+1)^{2}N_{2}}}\sum_{k=j^{2}N_{2}+1}^{(j+1)^{2}N_{2}}e^{-i\beta kf(z)}(1-e^{-i\beta f(z)})g(z){\mathop{}\!\mathrm{d}}z\Big{|}
=|𝒟(j+1)2N2eiβ(j2N2+1)f(z)(1eiβ((j+1)2N2j2N2)f(z))g(z)dz|\displaystyle=\Big{|}\int_{\mathcal{D}_{(j+1)^{2}N_{2}}}e^{-i\beta(j^{2}N_{2}+1)f(z)}(1-e^{-i\beta((j+1)^{2}N_{2}-j^{2}N_{2})f(z)})g(z){\mathop{}\!\mathrm{d}}z\Big{|}
𝒟(j+1)2N22|g(z)|dz\displaystyle\leq\int_{\mathcal{D}_{(j+1)^{2}N_{2}}}2|g(z)|{\mathop{}\!\mathrm{d}}z
2gD(j+1)2N2.\displaystyle\leq 2\|g\|_{\infty}D_{(j+1)^{2}N_{2}}.

Using again Lemma 3.3 with f=1f=1, we deduce that almost surely, there exists CC such that for all NN, DNCND_{N}\leq\frac{C}{N}. It follows that

|Hβ,f,gj|4Cg(j+1)2N2,|H^{j}_{\beta,f,g}|\leq\frac{4C\|g\|_{\infty}}{(j+1)^{2}N_{2}},

which yields

|j=1Hβ,f,gj|4CgN2j=21j2=o(β).\Big{|}\sum_{j=1}^{\infty}H^{j}_{\beta,f,g}\Big{|}\leq\frac{4C\|g\|_{\infty}}{N_{2}}\sum_{j=2}^{\infty}\frac{1}{j^{2}}=o(\beta).

As for Kβ,f,gjK^{j}_{\beta,f,g}, using the fact that the sequences (𝒟k)k1(\mathcal{D}_{k})_{k\geq 1} and (𝒟k)k1(\mathcal{D}_{-k})_{k\geq 1} are nested, we have

Kβ,f,gj\displaystyle K^{j}_{\beta,f,g} =k=j2N2+1(j+1)2N2|ϕx,β|(DkD(j+1)2N2+DkD(j+1)2N2)\displaystyle=\sum_{k=j^{2}N_{2}+1}^{(j+1)^{2}N_{2}}|\phi_{x,\beta}|(D_{k}-D_{(j+1)^{2}N_{2}}+D_{-k}-D_{-(j+1)^{2}N_{2}})
k=j2N2+1(j+1)2N2βfg(DkD(j+1)2N2+DkD(j+1)2N2).\displaystyle\leq\sum_{k=j^{2}N_{2}+1}^{(j+1)^{2}N_{2}}\beta\|f\|_{\infty}\|g\|_{\infty}(D_{k}-D_{(j+1)^{2}N_{2}}+D_{-k}-D_{-(j+1)^{2}N_{2}}).

Let C,δ>0C,\delta>0 such that for all N0N\neq 0,

|DN12π|N||CN1δ.\big{|}D_{N}-\frac{1}{2\pi|N|}\big{|}\leq CN^{-1-\delta}.

Then, for all k{j2N2+1,,(j+1)2N2}k\in\{j^{2}N_{2}+1,\dots,(j+1)^{2}N_{2}\},

0DkD(j+1)2N212πk12π(j+1)2N2+2Ck1δC(1j3N22+(j2N2)1δ).0\leq D_{k}-D_{(j+1)^{2}N_{2}}\leq\frac{1}{2\pi k}-\frac{1}{2\pi(j+1)^{2}N_{2}}+2Ck^{-1-\delta}\leq C^{\prime}\big{(}\frac{1}{j^{3}N_{2}^{2}}+(j^{2}N_{2})^{-1-\delta}\big{)}.

We deduce

|Kβ,f,gj|C′′fgN21j2,|K^{j}_{\beta,f,g}|\leq C^{\prime\prime}\|f\|_{\infty}\|g\|_{\infty}N_{2}^{-1}j^{-2},

and it follows that

j=1|Kβ,f,gj|=o(β).\sum_{j=1}^{\infty}|K^{j}_{\beta,f,g}|=o(\beta).

Finally, we have

|Gβ,f,gN2,|j=1|Gβ,f,gj2N2,(j+1)2N2|j=1|Kβ,f,gj|+j=1|Hβ,f,gj|=o(β).|G^{N_{2},\infty}_{\beta,f,g}|\leq\sum_{j=1}^{\infty}|G^{j^{2}N_{2},(j+1)^{2}N_{2}}_{\beta,f,g}|\leq\sum_{j=1}^{\infty}|K^{j}_{\beta,f,g}|+\sum_{j=1}^{\infty}|H^{j}_{\beta,f,g}|=o(\beta). (11)

We conclude the proof by putting together (6), (10) and (11). ∎

6. Funding

I am pleased to acknowledge support from the ERC Advanced Grant 740900 (LogCorRM), and later from the EPSRC grant EP/W006227/1 .

References

  • [1] Robert M. Blumenthal and Ronald Getoor. Some theorems on stable processes. Transactions of the American Mathematical Society, 95:263–273, 1960.
  • [2] Sudip Chakravarty and Albert Schmid. Weak localization: The quasiclassical theory of electrons in a random potential. Physics Reports, 140(4):193–236, 1986.
  • [3] Robert Chen and Larry A. Shepp. On the sum of symmetric random variables. Amer. Statist., 37(3):237, 1983.
  • [4] Jean Desbois, Cyril Furtlehner, and Stéphane Ouvry. Random Magnetic Impurities and the delta Impurity Problem. Journal de Physique I, 6:641–648, 1996. 13 pages, latex, 1 figure upon request.
  • [5] Jean Luc Desbois, Cyril Furtlehner, and Stéphane Ouvry. Random magnetic impurities and the landau problem. Nuclear Physics, 453:759–776, 1995.
  • [6] Oliver Johnson and Richard Samworth. Central limit theorem and convergence to stable laws in Mallows distance. Bernoulli, 11(5):829–845, 2005.
  • [7] Niclas Lindvall, Abhay Shivayogimath, and A. Yurgens. Measurements of weak localization of graphene in inhomogeneous magnetic fields. JETP Letters, 102:367–371, 09 2015.
  • [8] J. Rammer and A. L. Shelankov. Weak localization in inhomogeneous magnetic fields. Phys. Rev. B, 36:3135–3146, Aug 1987.
  • [9] Isao Sauzedde. Planar brownian motion winds evenly along its trajectory, 2021. arXiv:2102.12372.
  • [10] Isao Sauzedde. Winding and intersection of brownian motions, 2021. arXiv:2112.01645.
  • [11] Isao Sauzedde. Lévy area without approximation. Annales de l’Institut Henri Poincaré, Probabilités et Statistiques, 58(4):2165 – 2200, 2022.
  • [12] Wendelin Werner. Rate of explosion of the Amperean area of the planar Brownian loop. In Séminaire de Probabilités XXVIII, pages 153–163. Berlin: Springer, 1994.
  • [13] Wendelin Werner. Formule de Green, lacet brownien plan et aire de Lévy. Stochastic Process. Appl., 57(2):225–245, 1995.