Brownian windings, Stochastic Green’s formula and inhomogeneous magnetic impurities
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 60G171. Introduction
1.1. Stochastic Green’s formula
For a smooth loop and a point outside the range of , let be the winding index of around . For any smooth differential -form , the Green formula states that
where is the exterior derivative of . In other words, for two smooth functions ,
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 along . However, the index function fails from being integrable on the vicinity of [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 .
The first goal of this paper is to extend such a formula to any differential -form with regularity .
For an integer and a positive integer , let be equal to either or (the following theorem holds for both choice).
Theorem 1.
Let be a Brownian motion, and let be the winding function associated with the loop obtained by concatenation of with the straight line segment between its endpoints. Then, almost surely, for all and all ,
converges as .
Furthermore, if with is such that , almost surely,
where the stochastic integral in the right hand side is to be understood in the sense of Stratonovich.
Corollary 2.
For all and in , 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 , since we want to think of it as to the integral of with respect to the measure .
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 with intensity in the plane, and the Aharonov–Bohm effect is described by a phase shift .
In [4], the authors study the limit with constant, and derive a formula for the phase shift averaged over both and .
For an integrable function , is a Monte–Carlo estimation for , and therefore
However, as it is noticed in [5], for a Brownian loop ,
which is due to the lack of integrability of the function .
As we proved in [11], the Monte–Carlo method fails in this situation: it is true that -almost surely, converges in distribution (with respect to ) as , but the limit is not deterministic –or should we say, not measurable with respect to . It is instead equal to the sum of with a centered Cauchy distribution independent from . From this result, one can rigorously prove the formula obtained first in [5] for
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 given by
for a non homogeneous magnetic field and a non homogeneous density of impurities.
Theorem 3.
Let , with . For , let be Poisson process on with intensity , and let be either a Brownian motion or a Brownian bridge with duration , independent from . Then, -almost surely,
where is the expectation over (conditional on ).
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 , with . For , let be Poisson process on with intensity , and be either a Brownian motion or a Brownian bridge with duration , independent from . Let also be a standard Cauchy process. Then, for all , -almost surely, the -uple
converges in distribution toward where
Remark 5.
Given , there always exists a differential -form with regularity such that , so that can always be written as a stochastic integral.
Since all the results hold -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 norms.
2. Notations
2.1. Differential forms and integrals
For , we define as the set of functions such that the semi-norm
is finite. We also define , which we endow with the norm
For a differential -form and , we write if for all .
Given a curve , we write
where these integrals are to be understood either as classical integrals or as Stratonovich integrals, depending on the regularity of . No Itô integral will be involved in this paper, and all the stochastic integrals are to be understood in the sense of Stratonovich.
For , we identify the -form with the signed measure , where is the Lebesgue measure on .
For a bounded set and , we use the unconventional notation
and for the Lebesgue measure of .
2.2. Winding
Given a curve on , that is a continuous function from to for some , we write for the concatenation of with a straight line segment from to . Although the parameterisation of this line segment does not matter in the following, we will assume it is parameterized by at constant speed, unless is a loop (that is, a curve with ), in which case we set .
Given a curve and a point outside the range of , we write for the winding number of around .
For a relative integer , we define
For , we also define
and
We also write (resp. ) for the Lebesgue measure of (resp. ), and we drop the superscript when there is no doubt about the curve we are talking about.
For a real number and a positive integer , we set
Once we have shown that the limit
almost surely exists for all , we will write for this limit.
For a locally finite set of points , we define as the sum . If we are also given a function on , we define as the weighted sum
2.3. Cauchy variables
The Cauchy distribution with position parameter and scale parameter is the probability distribution on which has a density with respect to the Lebesgue measure given at by
A Cauchy random variable with position parameter and scale parameter is a random variable distributed according to . In ordre to unify some results, we will also write for a random variable which is actually deterministic and equal to .
Following [6, Definition 5.2]111As opposed to [6], we include the trivial case in our definition. , we will say that a random variable on lies in the strong domain of attraction of a Cauchy distribution if there exists such that
It then follows from Lemma 5.1 and Theorem 1.2 in [6] that follows a central limit theorem: if are i.i.d. copies of , then there exists a unique such that
Notice that the same assumptions with are not sufficient for such a central limit theorem to hold.
The parameters and such that are uniquely defined. We call them respectively the position parameter of , and the scale parameter of .222When 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
Lemma 3.1 (Lemma 5.2 in [11] ).
Assume belongs to the strong attraction domain of a Cauchy distribution. Then, its position parameter is equal to
When and lie in the strong attraction domain of Cauchy distributions, or even when they are Cauchy random variables, but they are not independent, does not necessarily belong to the strong attraction domain of a Cauchy distribution. What might be even more surprising is that, even if , , and are Cauchy random variables, can differ from (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 and be random variables which each lie in the strong attraction domain of a Cauchy distribution. Assume that there exists such that, for all , ,
Then, lies in the strong attraction domain of a Cauchy distribution, and .
The following lemma should be compared with the definition of the strong domain of attraction, where the random variable is given by , with fixed and a random point distributed according to (when ), where is a convex set containing .
Lemma 3.3 (Lemma 5 in [9]).
Let be a planar Brownian motion. For all , there exists such that almost surely, there exists such that for all bounded continuous function , for all ,
where is the continuity modulus of , i.e. .
From symmetry of the Brownian motion, Lemma 3.3 also holds with replaced with .
We will also need some control.
Lemma 3.4 ( Theorem 6.2 in [11] ).
For all and , there exists a constant such that for all ,
Finally, the following lemma will be used to check the condition inside Lemma 3.2.
Lemma 3.5 (Theorem 1 in [10]).
Let be two independent Brownian motions, starting from equal or different points in the plane. Then, almost surely converges as .
A few more results will be used, but will be easier to formulate later.
4. Stokes formula
In this section, is a standard Brownian motion under .
4.1. Existence of a limit
We will first prove the first part of Theorem 1:
Lemma 4.1.
Let . -almost surely, for all , the limits
exist and are equal. Almost surely, the application from to is continuous.
Proof.
We fix . Let be such that Lemma 3.3 holds, and let be the full probability event on which and Lemma 3.3 holds both for the sequence and the sequence , with a corresponding random constant .
On , for all , with , for all ,
(1) |
Thus, on , the sum
is absolutely convergent. By applying an Abel summation, we obtain
so that the right-hand side is convergent on the event .
The only thing that remains to be shown is the almost sure continuity of the application . Since it is clearly linear, it suffices to show that it is almost surely a bounded operator. By (1),
is bounded by for a random constant which depends on and , but not on nor . Thus, , 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 with the Stratonovich integral , when .
To this end, we decompose the trajectory into several pieces. First, we denote by the dyadic piecewise-linear approximation of with pieces: for , and ,
For , we also set , the restriction of to the interval . Finally, set the almost sure limit
By Lemma 4.1, scale invariance, and translation invariance of the Brownian motion, almost surely, is well -defined for all , for all , for all .333Since we use the translation invariance, the function is replaced with the random function . This is not an issue, because Lemma 4.1 holds almost surely for all , and not the other way around.
Let us first sketch the strategy of our proof. First, notice that for all which does not belong to nor to ,
which essentially comes from the additivity of the winding index, with respect to the concatenation of loops. Thus, it is reasonable to expect that
By applying the standard Stokes’ formula on the last integral, we get
As goes to infinity, we will see that the contribution from the small pieces (i.e. the sum over ) vanishes, whilst the integral along converges toward the Stratonovich integral , 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 , and such that .
Lemma 4.2.
For all , almost surely,
(2) |
Lemma 4.3.
As goes to infinity,
converges almost surely toward zero.
Lemma 4.4.
As goes to infinity, converges almost surely toward .
Of course, the conclusion that almost surely,
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
applied together with the Stokes formula for . However, neither nor are integrable, we have to deal with the cut-offs that allow to define and the : in general, for a finite ,
Proof of Lemma 4.2.
From linearity with respect to , we can and we do assume . In the event that that the restriction of to is identically vanishing, the result is trivial, and we thus assume that
is strictly positive.
Let be a random point in those distribution conditional on admits a density with respect to the Lebesgue measure, given by
Then, -almost surely, -almost surely,
Notice that, for , for equal to either , or to one of the , or to , it holds that
Thus, Lemma 3.3 ensures that -almost surely, the random variable belong to the strong attraction domain of a Cauchy distribution for either or . As for , is bounded by and therefore also belong to the strong attraction domain of a (degenerate, ) Cauchy distribution.
Let us check that, -almost surely, we can apply Lemma 3.2 to the set of variables
First, for , for ,
Besides, for , ,
for a random constant . The last equality follows from Lemma 3.5, applied to the independent Brownian motions
Notice that the constant depends upon and , but we can replace it with so that it only depends on . Furthermore, since there is only countably many couples , the previous inequality holds almost surely for all simultaneously.
Thus, we can indeed apply Lemma 3.2 to deduce that the, -almost surely, the position parameters add up:
(3) |
Furthermore, since is bounded, is quickly checked to be equal , that is
4.4. Contribution from the small loops
We now prove that almost surely,
We will first need the following result, which should be compared with Lemma 3.3.
Lemma 4.5.
Let and . There exists a constant and such that for all and all ,
Proof.
The proof is largely inspired from [9].
Let , which we will later take to be a function of . For , let be the restriction of to the interval . Let be the piecewise linear approximation of with pieces,
For , let
For outside , we have
It follows444See Section 3.2 in [11] for more details. that, for all such that ,
and therefore
We set , , , , , and we assume that is large enough for the inequality to hold. We also set to ease notations.
Using the fact that is contained inside the convex hull of , hence in the ball centered at with radius , we deduce that is bounded above by on . Thus,
Writing for the positive part of , to the power , and using the triangle inequality in , we obtain
We now use the asymptotic equivalence and , as well as Lemma 3.4, and the following estimations ([11, Lemma 2.4]): for all , there exists a constant such that for all ,
We end up with
for an arbitrary but fixed . The conditions on and ensures that all the exponents of are smaller than , so that there exists and such that
The negative part is treated in a similar way, and the lemma follows. ∎
Corollary 4.6.
Let and . There exists a constant such that for all , .
Proof.
Let and be the constants of Lemma 4.5. Then, for all and ,
By triangle inequality in ,
as expected. ∎
With this estimation in hand, we can now prove Lemma 4.3.
Proof of Lemma 4.3.
For , we define the constant function whose unique value is equal to , and . Since for all , is linear, it suffices to show that both
almost surely converge toward as .
From symmetry, for all , . It follows that, for ,
Besides, from a simple scaling argument,
Notice , which follows for example from the previous corollary.
We deduce that
This convergence rate is sufficient to conclude to the almost sure convergence: for all ,
In order to deal with the sum involving , one must be a bit careful about the way we use the translation invariance and scale invariance of the Brownian motion. We set and we define the event
for a fixed . Let be the (random) function defined by
In particular, satisfies the following properties:
-
on , so that, in the event , ,
-
, and ,
-
As a random variable, is measurable with respect to .
Set also , , which is a standard planar Brownian motion started from , independent from . Notice that and , so that
On the event , we have
Using Corollary 4.6 with , we deduce
Thus,
Since this holds for all , we deduce that almost surely converges toward as , 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 identifies the limit
with the Stratonovich integral of along , which is fairly classical. It is for example a direct consequence of the following lemma.
Lemma 4.7.
For a given dissection , and a Brownian motion, let be the piecewise-linear approximation of associated with : for and ,
For , let
Then, almost surely, for all , as ,
Proof.
Let . On the almost sure event , we have
The second convergence is proved in a similar way:
∎
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 is a Brownian loop started from . To deal with the case when is a Brownian bridge from to , one must also take into account the winding function of the triangle between , , and , but this is done in a straightforward way.
From linearity, it suffices to prove the result when restricted to functions . Furthermore, since the result is trivial in the event , we assume .
Let be the restriction of to , its restriction to , and . Then, the distribution of (resp. ) admits a density with respect to the density of a standard planar Brownian motion defined on . Using scale invariance, we can apply Theorem 1 to both and . We deduce that, for , for all , almost surely, for all ,
converges as , and the limits are almost surely equal to respectively and , where is such that .
Now we need to show that almost surely, for all , and add up properly, for which we proceed as in Lemma 4.2, introducing again a random point . Going through the same arguments as in the proof of Lemma 4.2, we see that it suffices to show that, -almost surely,
(4) |
for the four possible couple of signs in front of , and for some .
To prove (4), we further decompose and by setting (resp. , ,) the restriction of to the interval (resp. , , ). Then, where .
We show that almost surely, , the 15 other intersections are treated either similarly. Conditionally on , and are independen. Furthermore, both their distribution, conditional on , have a density with respect to the distribution of a standard Brownian motion with duration , started respectively from and . Thus, it suffices to show that for all , when and are independent Brownian motions started respectively from and . This follows directly from 3.5, with a scaling of . ∎
5. Magnetic impurities
In this section, we fix a function . For all , we define a Poisson process on with intensity , independent from , and a standard Cauchy process, independent from . We write the expectation with respect to , the one with respect to , the expectation with respect to and the expectation on the product space (although none of the variables we consider depend on both and , so truly or , whichever is relevant).
For a function , we define
as well as
Notice that almost surely has a finite -variation for all (see [1, Theorem 4.1]). Since -almost surely, , the integral is well-defined as a Young integral.
The main result from this section is the following
Lemma 5.1.
Let be continuous and bounded functions. Assume that takes non-negative values. Let
Then, -almost surely, as ,
(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 is integrable against the intensity measure of , we can use Campbell’s theorem, which gives
where .
Besides, conditional on , is a centered Cauchy random variable with scale parameter , whilst is deterministic. It follows that
Furthermore, since both and are linear in , one can use the Cramér-Wold device to deduce Corollary 4 from its special case . By Lévy’s continuity theorem, this specific case is equivalent to the statement that -almost surely, for all ,
From our previous computation, this amount to show that almost surely, for all ,
which follows again from Lemma 5.1. ∎
Proof of Lemma 5.1.
From symmetry, we can assume . Performing an Abel summation, we obtain
where
The two terms in (5) comes from two different parts in this last sum: the term comes from the bulk of the sum, that is the part with of the order of . The second term comes from the tail of the sum, or more precisely from the part of the sum when is of the order of . We will split the sum into several parts. For with , we set
For and which will be set later on, we decompose into three parts,
As , both and will slowly diverge toward . In particular, . The reason why we need to treat the end part in a separate way is that its convergence toward is not absolute, in the sense that the
does not converge toward zero as , 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 , as , uniformly in ,
and it follows that
For , let be such that for all ,
and set .
We now look at the tail part of . Let and (random) be such that for all and ,
Recall that the existence of such a couple is provided by Lemma 3.3. Let be any integer-valued function such that and .
For all , . We deduce that for all and ,
so that
We deduce that, for all ,
and there exists constants such that for all ,
The remaining part of the analysis is standard calculus. Set
Then, for ,
It follows that
The last line follows from the fact that
For , let
so that is continuous on and
(7) |
For all ,
where is the continuity modulus of .
Since for all , there exists a function such that as and . We fix such a function, and set . This way, we do have and .
We obtain
(8) |
To estimate this last integral, there is two things we must be careful about. First, because of the function in the definition of , the function is not integrable on so we cannot naively replace the bound with its limit. Secondly, when manipulating the integral, we must be extra careful at the vicinity of .
Recall that for , . Performing an integration by part, we deduce that for all and ,
It follows that
(9) |
We finally look at the end part of . Since the norm of becomes arbitrarily large as goes to infinity, one cannot directly rely on Lemma 3.3. For a positive integer , we decompose into
We have
Using again Lemma 3.3 with , we deduce that almost surely, there exists such that for all , . It follows that
which yields
As for , using the fact that the sequences and are nested, we have
Let such that for all ,
Then, for all ,
We deduce
and it follows that
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.