A new noise, based on vortex structures in 2D (point vortices) and 3D (vortex filaments), is introduced. It is defined as the scaling limit of a jump process, which explores vortex structures, and it can be defined in any domain, also with boundary. The link with fractional Gaussian fields and Kraichnan noise is discussed. The vortex noise is finally shown to be suitable for the investigation of the eddy dissipation produced by small scale turbulence.

The theory of Stochastic Partial Differential Equations (SPDEs) is nowadays very well developed (see, for instance, Refs. 8, 20, 29, and 31, with many contributions on fluid dynamics models, like).3,7,11,12,24,33 However, with the exception of the literature making use of Kraichnan noise, which is motivated in fluid dynamics by its invariance and scaling properties, in most cases, there is no discussion about the origin of noise and its form, in connection with the fact that it is part of a fluid dynamic model. The purpose of this work is to introduce an example of noise based on vortex structures, both in 2D (point vortices) and 3D (vortex filaments). We discuss its motivations and interest for the understanding of fluid properties.

Some preliminary forms in 2D have been introduced in Refs. 13 and 19, but the noise defined here is different and goes much beyond, in particular, because we treat the 3D case on the basis of the theory of random vortex filaments (see Sec. III B).

Usually, in general or theoretical works on SPDEs, the noise is either specified by means of its covariance operator or by means of a finite or countable sum of space-functions multiplied by independent Brownian motions. Here, we start from a different viewpoint. Motivated by the emergence of vortex structures in turbulent fluids, we idealize their production/emergence process by means of a sequence of vortex impulses, mathematically structured using a jump process taking values in a set of vortex structures. This is described in Sec. II. A suitable scaling limit of this jump process gives rise to a Gaussian noise in a suitable Hilbert space. Different examples of such noise depend on different choices of the vortex structures and their statistics, at the level of the jump process. A heuristic picture then emerges of a process that fluctuates very rapidly between the elements of a family of vortex structures. The realizations of this noise are made of vortex structures, which idealize those observed in turbulent fluids—point vortices in 2D and vortex filaments in 3D.

This noise is motivated by turbulent fluids. In the physical literature, the most common noises related to turbulence are the Fractional Gaussian Field (FGF) and Kraichnan noise (see, for instance, Refs. 1, 5, 9, 10, 2123, and 26). In Sec. IV, we show that on a torus in two and three dimensions, the vortex noise covers FGF and Kraichnan noise by a special choice of the statistical properties of the regularization parameter and the vortex intensity. The vortex noise is thus a flexible ensemble—it may cover also multifractal formalisms (see also Ref. 14)—and its realizations are the limit, as described in Secs. II and III, of localized-in-space vortex structures similar to those observed in turbulent fluids.

Finally, another main motivation for this investigation has been the recent results on eddy dissipation, showing that a transport type noise depending in a suitable way on a scaling parameter, in a transport-diffusion equation, in the scaling limit gives rise to an additional diffusion operator.13,17 These results require that the covariance function of the noise, computed along the diagonal, Qx,x, is large, but the operator norm of the covariance is small. We check when the vortex noise satisfies these conditions. Heuristically speaking, they are satisfied when, in the scaling limit, the vortex structures defining the noise are more and more concentrated at small scales. This confirms the belief that eddy diffusion is a consequence of turbulence but only when it is suitably small scale.

When a fluid moves through the small obstacles of a boundary (hills, trees, and houses for the lower surface wind, mountains for the lower atmospheric layer, coast irregularities for the sea, and vegetation for a river) or it moves through small obstacles in the middle of the domain (like islands in the sea), vortices are created by these obstacles, sometimes with a regular rhythm (von Kármán vortices) or sometimes more irregularly. In principle, these vortices are the deterministic consequence of the dynamical interaction between the fluid and structure, but in very many applications, we never write the details of those obstacles when a larger scale investigation is done. Hence, it is reasonable to re-introduce the appearance of these vortices, so important for turbulence, in the form of an external perturbation of the equations of motion.

Assume that the velocity field at time t is ut,x. We may idealize the modification of ut,x due to the emergence of a new vortex near an obstacle as an event occuring in a very short time around time t so that we have a jump,
where σx is presumably localized in space and corresponds to a vortex structure. Continuum mechanics does not make jumps; we idealize a fast change due to an instability as a jump for a cleaner mathematical description.
We may develop the previous idea in two directions. The simplest one is suitable for investigations, such as the effect of turbulence on passive scalars,5 where a simple model of random velocity field is chosen: we consider a stepwise constant velocity field with jumps such as those described above; later on, we shall take a suitable scaling limit and get a Gaussian velocity field, delta correlated in time, with space correlation of very flexible form. A more elaborate proposal is to consider the Navier–Stokes equations with an impulsive force given by a process with jumps,
Here, K is an index set, and for each kK, we denote by t1k<t2k< the sequence of jump times of class k and by σk the vortex structure (described at the level of velocity field) arisen at time tik. This way the fluid moves according to the free Navier–Stokes equations between two consecutive jumps times. In Sec. II B, we formalize the noise kKiδttikσk, or more precisely, similarly to what it is done for white noise and Brownian motion, we formalize the time integral of this distributional process,
(1)
In this first heuristic formulation, it is natural to introduce an index set K, but below, we shall avoid this.

Given an open domain DRd, d = 2, 3, denote by Cc,solD,Rd the space of smooth solenoidal vector fields with compact support in D and denote by H the closure of Cc,solD,Rd in L2D,Rd. One can prove, under some regularity of the boundary, that uH is an L2D,Rd-vector field, with distributional divergence equal to zero, tangent to the boundary.32 The norm uH is given by uH2=Dux2dx.

The following scheme is taken from th work of Métivier,28 first three chapters. The main tightness and convergence results for martingales, as described in Ref. 28, are due to Rebolledo.30 

Let P be a Borel probability measure on H. Assume that
(2)
for some nondecreasing φ:R+R+ that grows faster than quadratic, i.e.,
(3)
Denote by QP the trace class covariance operator defined as
Assume that P has zero average,
(4)
We may also define, a.s. in x,yD, the covariance (matrix-valued) function,
Indeed, Hhx2Pdh< for almost every xD, thanks to Fubini–Tonelli theorem, since HDhx2dxPdh<.
Consider the continuous time jump Markov process in H with law of jumps,
[vH, ABH], namely, with the infinitesimal generator,
for all bounded continuous functions F:HR. Here, τ > 0 is the average interarrival between jumps. Denote by Wt0 the corresponding Markov process with the initial condition W00=0. The Dynkin formula
gives us the decomposition in a finite variation plus a martingale term. Consider first the case when F1v=v (here we do not write down classical details, namely, that the computation should be done for a continuous bounded cutoff of each component v,ei, where ei is a complete orthonormal system; see Ref. 28, p. 14). One has
because mP = 0. Hence, Wt0=MtF1, namely, the process Wt0 is a martingale. Let us compute its Hilbert-space-valued Meyer process W0t. We use the function F2v=vv (again one has to do the computation first for a cutoff of the functions v,eiv,ej),
Therefore, Wt0Wt0=tτQP+MtF2. The Meyer process W0t is thus (see the definition in Ref. 28, pp. 8–12)
Let us now parameterize and rescale the previous process. We take average interarrival between jumps given by
and we reduce by 1N the size of jumps by considering a probability measure PN on H with zero average mP=HhPNdh=0 and covariance QPN given by
Consider the associated process WtN, a martingale with the Meyer process

Definition 1.

Given QP, denote by Wtt0 a Brownian motion on H with incremental covariance QP.

Theorem 2.

The process WtNt0 converges in law to Wtt0, uniformly on every compact set of time, as processes with values in H.

Proof.
Using the classical theorem of tightness for martingales [cf. Ref. 28 (Chap. 2) and Ref. 30], we have that the family of laws of the processes WNN is tight in the Skorohod space (because the family of laws of WN is tight), and every convergent subsequence has limit given by the law of a martingale Wt with W0 = 0 and Meyer process,
If we establish that W has continuous paths, then it is a Brownian motion with incremental covariance QP. One can prove that
(5)
where ΔsWNH is the size of the jump (if any) at time s (WN is càdlàg). Since the set sups0,TΔswNH>ϵ is open in the Skorohod topology, from the Portmanteau theorem, we get
for every ϵ > 0, and hence, W is continuous. To show (5), denote by {si}i=0NT[0,T] the Poisson (τN1) arrival times, and then, we have that
where we used that given the Poisson arrival times, the laws of each jump size ΔsiWNH is independent of it, and identically distributed as what we simply denote by ΔWNH. By the elementary inequality (1 − y)n ≥ 1 − ny, for any y ∈ [0, 1] and nN and Markov’s inequality, we have that
which is finite by (2) and converges to zero as N → ∞ by (3).■

This is a side section, which, however, may help the intuition [see also (1)]: we reformulate the jump process Wt0 as a Poisson Point Process (PPP). On a probability space Ω,F,P, let P be a PPP on [0, ∞) × H with intensity measure λLebP, where λLeb is Lebesgue measure scaled by λ > 0 and P is the probability measure introduced in Subsections II A–II C. Heuristically,
where ti,σi is an i.i.d. sequence with ti “uniformly distributed on [0, ∞),” σi distributed according to P, and ti and σi independent of each other. Define the vector valued random field, defined on Ω,F,P,
Compared to (2), we may think that K in that formula was a finite set, and we have simply reordered the jump times tik in a single sequence ti and we have renamed the jump velocity fields. This definition is slightly heuristic because it makes use of the representation as infinite sum, which is true only in a suitable limit sense; a rigorous definition of Wt,x is
However, in the sequel, for the sake of interpretability, we shall always use the heuristic expressions.

The intuition is that eddies σix are chosen at random with distribution P, with exponential inter-arrival times of rate λ. Condition (4) asks, heuristically speaking, that both an eddy and its opposite are equally likely to be chosen.

Rescale Wt0x as
Let us compute the expectation and the covariance function of this process. One has [E denotes the Mathematical expectation on Ω,F,P]
from the independences and condition (4). Moreover,
having used the independence when ij and property (4) again; hence,

Proposition 3.
Hence,

Proof.
We note that
where ηλ(·) denotes a Poisson process on R+ with intensity λ.

This is another way of seeing the link between the noise with jumps and the covariance of the limit Brownian motion.■

The mathematical object discussed in Sec. II B and C, although initially motivated by vortex structures, was completely general: given any probability measure P on H with covariance QP, the previous construction and results apply and defines a Brownian motion Wt in H with covariance operator QP. Note that P is not necessarily Gaussian: P and W1 have both covariance QP, but only W1 needs to be Gaussian. In a sense, we “realize” approximately samples of the Brownian motion Wt by means of samples of a possibly “nonlinear” (non-Gaussian) process WtN.

In this section, we give our two main examples of the measure P, highly non Gaussian. It is inspired by vortex structures.

Common to both descriptions are a few objects. First, given δ > 0, we define
Second, we have a filtered probability space Ω,F,Ft,P and several F0-measurable r.v.’s: (a) X0 with law p0dx supported on Dδ, which will play the role of the center of the vortex in 2D and the initial position of the vortex filament in 3D; (b) Γ, real valued, with the physical meaning of circulation, with
(c) L, positive valued, randomizing the size of the mollification, with the property
(d) U, positive valued, randomizing the length of the vortex filament. Moreover, in 3D, we also have (e) a Brownian motion on Ω,F,Ft,P with values in R3. In the 2D case, we just take F=F0 and do not need the filtration.

For sake of simplicity of exposition, we shall always assume that X0, Γ, L, and U are independent, but most of the results can be extended to more general cases.

The last common element of the theory is a smooth symmetric probability density θ supported in the ball B0,1 and its rescaled mollifiers,
(6)
with support in B0,.
In 2D, by a point vortex, we mean a vorticity field of delta Dirac type, δx0; its use in 2D fluid mechanics is manifold (see, for instance, Ref. 27). If the vorticity is assumed distributional and equal to δx0, with x0 in the interior of D, then the so-called stream function ψD,x0 is given by the solution of
and the associated velocity vector field is given by
where f=2f,1f. One has
where hD,x0 is a smooth function, solution of the problem
In the sequel, as it is customary, we shall denote uD,x0x simply by Kx,x0. Hence,
(7)
where x = (x2, −x1).
Recall that θ (6), as → 0 is an approximation of the Dirac delta function. Expressions of the form θ(xx0) are idealized smoothed point vortices, at the vorticity level, and the associated velocity field is
With these preliminaries, let us define P.

Definition 4.
In the 2D case, the probability measure P on the space H is the law of the H-valued r.v.,
(8)

For future reference, the spatial covariance matrix of the vortex noise in 2D is given by
(9)

Proposition 5.
The random vector field of Definition 4 takes values in H. If
for some p > 2, then it satisfies (2) and (3). Moreover, it satisfies (4).

Proof.
Fixing any p > 2, we compute by independence between X0, Γ, and Hölder’s inequality [with p1 = p/2, p2 = p/(p − 2) such that 1/p1 + 1/p2 = 1]
where recall that
Per fixed xD, we perform the following analysis. Since X0Dδ and |yX0| ≤ L < δ/2 for any y contributing to the above integral, we have yDδ/2. Therefore, the part xhD,y(x) of the kernel K(x, y) is smooth as a function of xD for every yDδ/2. Due to continuous dependence of hD,y(x) on boundary conditions, hence on the variable y, the following constant is finite:
The contribution of xhD,y(x) to the above integral hence is finite, i.e.,
where we used that DθL(yX0)dy=1 for any realization of X0. It suffices now to focus on the other part of the kernel (2π)1xyxy2. We have that
where we use the fact that the integral of |y″|−1 over a unit ball centered anywhere in R2 is maximized when the center is the origin, and nonrandom constant Cp,θ is independent of x. Hence, we get that
Finally, it satisfies (4),
because the second expectation is finite and the first one is equal to zero, by assumption.■
The case when L = 0 is outside the previous definition and result. The velocity field Kx,x0 is not of class H. Nevertheless, it is of class LpD,R2 for p < 2 or of class HsD,R2 for s > 0. Therefore, we may consider the random field,
taking values in these spaces and call P its law. We shall see below that it satisfies certain special properties.

In 3D, by vortex filament we mean a distributional vector valued field (a “current,” in the language of Calculus of Variations18), given by
where Xt is a function or a process such that the previous expression is well defined. We have already introduced a possibly relevant stopping time τ because it may help to cope with the presence of a boundary. Stochastic currents have been introduced and investigated in some works.2,4,15,16 We do not need, strictly speaking, that theory here since we shall always deal with mollified objects, except in one section where we explain what is necessary. In this work, we shall always assume that Xt has the law of a Brownian motion, but it is interesting to investigate also other processes, for instance, directed polymers, such as in Ref. 25.
The following construction of a vortex filament in 3D is due to Ref. 14 (which we slightly modify). Let (Γ,U,)R+3 be a triple whose joint distribution is given by some probability measure ν(, du, dℓ) (assumed to be a product measure for simplicity). Let (Xt)t0 denote a 3D Brownian motion starting with X0 distributed with a probability density p0(x) supported in Dδ, where p0(x) ∈ [pmin, pmax] ⊂ (0, ∞). We call W its law, which we assume to be independent of ν(·). Define the first exit time from Dδ of (Xt) by
We consider random vorticity fields defined as
Let Ax be the vector potential defined path by path by the solution of the equation
and extend A = 0 outside of D, when necessary. Then, the associated velocity is given by
Concerning the Biot–Savart kernel, here we have
where hD,x0 is a smooth function, solution of the problem
As usual, we shall denote curlψD,x0x simply by Kx,x0, which now is vector valued and its action on a generic vector v is given by
(10)

Definition 6.
In the 3D case, the probability measure P on the space H is the law of the H-valued r.v.,
(11)

Remark 7.

We use the killed BM, not the normally reflected BM, in the definition of the filament because the latter is not a local martingale, only a semimartingale due to the boundary push term, which leads to difficulties in integration against dXt.

For future reference, the spatial covariance matrix of the vortex noise in 3D is given by
(12)

Proposition 8.
The random vector field of Definition 4 takes values in H. If
for some p > 2, then it satisfies (2) and (3). Moreover, it satisfies (4).

Proof.
Fix any p > 2, then we compute
Fixing any realization of (Γ, U, L) according to measure ν, we take expectation with respect to the Wiener measure W first. By Hölder’s inequality and p/2 > 1 and Burkholder–Davis–Gundy inequality, we compute
Since XtτDδ, we have that any y that contributes to the above integral is supported in yDδ/2; hence, xhD,y(x) part of the kernel K(x, y) is uniformly bounded, i.e.,
Hence, its contribution in the above integral can be computed, as for any xD,
using that ∫θL(yXt)dy = 1 for every possible realization of XtτDδ.
It suffices to focus on the other part of the kernel (4π)1xy|xy|3. We can do an explicit calculation: by Hölder’s inequality and then a change of variables, we have that for any xD,
where Cp,θ is a non-random constant independent of x. Indeed, we used the geometric fact that the integral of the function |y″|−2 over a unit ball centered at anywhere in R3 is maximized when the center is the origin.
Thus, we can conclude that
with the finiteness of the RHS providing a sufficient condition.
Finally, it satisfies (4),
because the second expectation is finite and the first one is equal to zero, by assumption.■

In this section, we analyze the covariance operators of our vortex noises constructed above in 2D and 3D and show that our vortex noises are instances of Fractional Gaussian Fields,26 which is a broad class of Gaussian generalized random fields that includes Gaussian Free Field (GFF) and Kraichnan noise. We show that by choosing the statistical parameters of our model suitably, we can reproduce a large class of FGF. It may also reproduce multifractal vector fields, which was the main motivation of study in Ref. 14.

For simplicity, our fields are defined on the torus Td, d = 2, 3.

In the scalar case and on the torus Td=Rd/Zd, the classical d-dimensional FGF of index sR is the Gaussian field with covariance Δs, where Δ is the Laplacian in on Td (see Ref. 26). The case s = 1 is called Gaussian Free Field (GFF). Similarly, let us introduce a Gaussian measure on solenoidal vector fields. Let H be the space of mean zero periodic L2 solenoidal vector fields. The Stokes operator is defined as
(no projection of L2Td,Rd to H is needed here, opposite to the case of a bounded domain with Dirichlet boundary conditions). The Laplacian Δv is computed componentwise. The operator A is invertible in H (see Ref. 32). With these definitions at hand, we call Solenoidal Fractional Gaussian Field (SFGF) of index sR the Gaussian measure with covariance As. The case s = 1 will be called Solenoidal Gaussian Free Field (SGFF).

Let us first consider the 2D case, and recall the definition of the noise based on point vortices (8).

The covariance operator of our noise is given by
Call Qvortexx,x its covariance function (matrix-valued) such that
It is clear (and proved below) that it is homogeneous,
for a matrix function Qvortexx. In the sequel, we denote by Z0d the set Zd\0.

Proposition 9.
Assume θ symmetric and X0 independent of Γ,L and uniformly distributed. Then,
(13)

Proof.
We may rewrite
Therefore,
By the Parseval theorem,
recalling that
and calling Pk=Ikk|k|2 is the projection on the orthogonal to k. Therefore,
Since θ̂k=θ̂k, we get the result.■

Corollary 10.
In addition, assume θ is a smooth function with θ̂k=θ̂k, let fL be the probability density of L, and assume Γ is a function of L: Γ=γL. Assume
for some C > 0 and
Call
which is a finite constant. Then,
This is the covariance function of a SFGF of index,

Proof.
Since θ is smooth, θ̂r has a fast decay, which makes θ̂rrα integrable at infinity for every α; it is also integrable at zero because α > −1. From the assumptions,

Note that α > −1 corresponds to■
so the SGFF (s = 1) is a (just excluded) limit case.
Recall that the solenoidal Kraichnan model with scaling parameter ζ is defined, on the torus Td, by the covariance function
We see thus that the vortex noise, in dimension d = 2 (see Sec. IV B for d = 3), covers the Kraichnan model with scaling parameter,
(any positive ζ is covered).

The space-scale of the vortices is free in the previous results. If we restrict ourselves to small vortices, namely, we take fLr=0 for r>k01, we get the following corollary:

Corollary 11.
Under the same assumptions of the previous corollary except for
for some C, k0 > 0 and α > −1, we get
where
for some constant C′ > 0.

Proof.
As above,
The first limit property is obvious. Moreover (using also θ̂1),
and hence,

We thus see that, up to lower order terms, the vortex model with cutoff corresponds to the Kraichnan model with infrared cutoff k0 [cf. Ref. 10, Eq. (2.3)].

Finally, we remark that the model has the flexibility of multifractality. To explain it in the simplest possible case, assume
Then, we get
Clearly, one can do the same with a continuously distributed multifractality in place of the finite sum (we void to introduce additional notations to explain this point).

Remark 12.

An intriguing but extremely difficult question (we thank an anonymous referee for it) is whether we may infer the value of the scaling exponent ζ of the Kraichnan model, or a multifractal version of it, from the similarity with the vortex noise. It was the main aim of the outstanding book,6 which—as admitted by the author—remained open at the time of the book and it is still open now. Two examples of attempts in this direction have been Refs. 14 and 25; in the latter work, a multifractal formalism based on vortex filaments was developed. However, it must be stressed that no one of these works deduced K41 or other scalings from vortex models; they could only reproduce scalings chosen a priori.

Next, we turn to the 3D case, and recall the definition of the noise based on vortex filaments (11). The covariance of the noise is given by
where
For simplicity, we set from now on the time-horizon U = 1, and assume that the 3D Brownian motion (Xt) starts from uniform distribution on T3, and hence, for any time t > 0, the distribution of Xt remains uniform. (Xt) is also independent of (Γ, L). Using vector identity, we may rewrite
For the 3D kernel K (10), we still have the property that Kx,a=Kxa=Kax.

Our first result is that in 3D, the vortex noise has the same covariance structure as in the 2D case.

Proposition 13.
Assume θ symmetric and (Xt) independent of Γ,L and starts from uniform distribution on T3. Then,

Proof.
where we take conditional expectation with respect to (Xt) first using its time-stationarity and uniform distribution, whereas the randomness of (Γ, L) remains.
By Parseval theorem and vector identities, we may rewrite
By properties of the triple cross product, we have that
and hence,
recalling that in 3D,
Thus, we may conclude that
where Pk=Ik|k|k|k| is the projector on the orthogonal to k. This yields, in turn, that the covariance matrix of the noise is given by■

This formula agrees with formula (13) obtained for 2D; hence, Corollary 10 applies in 3D without change (except for summation over kZ03).

Our result in 3D covers Kraichnan noise with parameter
We can also restrict the vortices to small scales by introducing a cutoff k0, as in Corollary 11. Here, we need to restrict to α > 0 in its statement so that the remainder Rk0(x) is of lower order,
Regarding eddy diffusion enhancement in domains with boundary, we recall the following theorem proved in Ref. 13 (Theorems 1.1 and 1.3). Here, we have a passive scalar θ driven by the white-in-time, correlated-in-space noise tW produced by our vortex structures, where W(t, x) is the limit Gaussian process obtained via the invariance principle in Theorem 2,
◦ denotes Stratonovich integration, and scalar κ > 0. We denote the smallest eigenvalue of the matrix Q(x, x) by
and the squared operator norm Q1/2L2(D)L2(D)2 by

Theorem 14

(Ref. 13, Theorems 1.1 and 1.3).

  • For any θ0H measurable and any t ≥ 0, we have that
    where λD,κ,Q is the first eigenvalue of the elliptic operatorAQ for
  • There exists a constant CD,d > 0 such that
    for every Q such that

In view of this theorem, our aim is to show that the noises based on vortex structures in 2D and 3D that we constructed in Sec. III, for small L, enjoy the property that they have small ϵQ and large q(x, x), simultaneously, once the other parameters of the model are tuned properly. Here, we assume that Γ, U, L, X are independent.

For technical reasons, we demonstrate this only for the torus D=Td, d = 2, 3, in this section. The same conclusions should be true for any regular domains D, but the corrector part of the Green function is difficult to handle; hence, we prefer to state in the simple case of torus. Note in this case, we do not have a boundary, and hence, Dδ=D, δ = 0, and we can put the stopping time τ = ∞ in the 3D case.

The following theorem applies to any realization of L. For fixed > 0, we shall use [recall (8)]
Therefore, for ξR2, we have
while for vH,
In the next statement, we set σ2=E(Γ2).

Theorem 15.
(i) There exists a finite constant C such that for every vH and ∈ (0, 1),

  • (ii)For every xT2, let qx0 be the largest number such that for any vR2 and ∈ (0, 1),
    Then, there exists some positive constant c such that

Remark 16.

We can choose σ2=E(Γ2) to be small and then choose small enough such that σ2|log | is large to fulfill the conditions in Theorem 14.

Proof.
Since D=T2, the function xhDx,y is bounded above uniformly and does not affect the computations on Kx,y, which will be based only on the term 12πxyxy2. Thus, we use the approximation for all xT2, a.s.,
Let CK be the random variable defined as
Under our approximation, we have
and hence, CK is finite a.s. and even uniformly bounded above. Then,
Let C̃K be the deterministic constant defined as
We have proved
Concerning the size of C̃K, under the assumptions that p0 has a bounded density, we have
since
Therefore,
This quantity is small if E(Γ2) is small.

Concerning vTQx,xv, vR2, using again the simplified asymptotics, we have
Given any xT2 and unit vector vR2, there is a cone Cx,vT2 (a set of the form x + rw, r0,r0, w=1, w · eα for some e=1 and α0,1) such that
and
Moreover, assume p0dx0 is bounded below by pminLeb for some constant pmin > 0. We then have
Taking > 0 very small, reduce the cone Cx,v to the set
of points x0 such that
We then have
and thus,
The last inequality is because the quantity x0θ*12x0 is rotationally invariant; hence, the integral C0,vθ*12x0dx0 does not depend on v. Since |C(0, v)| ≥ η, we have that
Let us investigate the problem of the scaling in of the quantity θ*12xdx. Given the mollifier θx=2θ1x that we assume the best possible one (non-negative, smooth, symmetric), let us introduce the smooth symmetric pdf, compactly supported in B(0, 2),
Then,
Below we shall use the formula,
true because
(recall θ is symmetric). After these preliminaries, we have
Now, we have to understand first the behavior of
We can prove that for z1,
Indeed, since |yz| ≤ |y| + |z|,
Then, 1y1yzdyθ2zdz can be bounded below by
where without loss of generality
This yields that
for some c > 0 and any ∈ (0, 1).

Recall that we take D=T3; hence, the computation below can be based solely on the 14πxyxy3× part of the kernel K(x, y) (10), with the other part from xhD,x0x× uniformly bounded. We also set τ = ∞. The following theorem applies to any realization of L. We shall use the notation Q(x, y) and Q for fixed , similarly to what is done in the 2D case, while recalling (11).

In the next statement, we set σ2=E(Γ2).

Theorem 17.

(i) There exists a constant C < ∞ such that for every vH and ∈ (0, 1),

  • (ii) There exists a constant c > 0 such that for all xT3, vR3, and ∈ (0, 1),

Remark 18.

We can choose the distribution of (Γ, U) such that E(U)σ2 is small and then choose small enough such that E(U)σ21 is large to fulfill the conditions in Theorem 14.

Proof.
Taking any vH, we consider
For any fixed realization of (Γ, U), we take expectation over W first
where the last step is due to Itô isometry. We further bound it above by moving the norm inside the integral,
where the random constant Cu is
for some deterministic finite constant CT3 (integrate first dx then dy). Set
Recall that X0 has density p0(x), which is bounded above uniformly by pmax. Since the heat semigroup is an L-contraction, the density of Xt at any later time t is bounded above by pmax, and thus, we have
(integrating first dz then dy) for some deterministic finite constant CT3. We conclude with
Taking now any unit vector vR3, for any xT3, we consider the quantity
We again fix any realization of (Γ, U, ) and take expectation over W first,
where the last step is due to Itô isometry.
Since D=T3 is compact, the density of Xt, denoted pt(z), converges to the uniform distribution, and hence, it is not hard to see that there exists some pmin > 0 independent of t such that
Then, we can continue to bound below WΓ2vu(x)2 by
For any xT3, there exist a cone C(x, v) and a ball B = B(x*, /2) ⊂ C(x, v) of radius /2 with center x* with |xx*| = 2 such that provided zB, we have all the y that contribute to the above integral be contained in B(x*, 3/2) and /2 ≤ |xy| ≤ 7/2, and on the other hand, the orientation of the cone is chosen such that v × (xy) are roughly in the same direction for all the y. This implies that for some absolute constant c > 0 and any zB,
Thus, we have that upon squaring and using |B| ≍ 3,
This completes the proof.■

We thank an anonymous referee for the contribution to Sec. IV, which was prepared after the advice to compare better our model with the FGF.

The research of F.F. is funded by the European Union (ERC, NoisyFluid, Grant No. 101053472). Views and opinions expressed are, however, those of the authors only and do not necessarily reflect those of the European Union or the European Research Council. Neither the European Union nor the granting authority can be held responsible for them.

The authors have no conflicts to disclose.

Franco Flandoli: Formal analysis (equal); Investigation (equal); Writing – original draft (equal). Ruojun Huang: Formal analysis (equal); Investigation (equal); Writing – original draft (equal).

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

1.
Apolinário
,
G. B.
,
Beck
,
G.
,
Chevillard
,
L.
,
Gallagher
,
I.
, and
Grande
,
R.
, “
A linear stochastic model of turbulent cascades and fractional fields
,” arXiv:2301.00780 (
2023
).
2.
Bessaih
,
H.
,
Coghi
,
M.
, and
Flandoli
,
F.
, “
Mean field limit of interacting filaments and vector valued non-linear PDEs
,”
J. Stat. Phys.
166
(
5
),
1276
1309
(
2017
).
3.
Breit
,
D.
,
Feireisl
,
E.
, and
Hofmanová
,
M.
,
Stochastically Forced Compressible Fluid Flows
(
De Gruyter
,
Berlin
,
2018
).
4.
Capasso
,
V.
and
Flandoli
,
F.
, “
On stochastic distributions and currents
,”
Math. Mech. Complex Syst.
4
(
3–4
),
373
406
(
2016
).
5.
Chaves
,
M.
,
Gawedzki
,
K.
,
Horvai
,
P.
,
Kupiainen
,
A.
, and
Vergassola
,
M.
, “
Lagrangian dispersion in Gaussian self-similar velocity ensembles
,”
J. Stat. Phys.
113
,
643
692
(
2003
).
6.
Chorin
,
A. J.
,
Vorticity and Turbulence
,
Applied Mathematical Sciences Vol. 103
(
Springer
,
Berlin
,
1994
).
7.
Chow
,
P.-L.
, “
Stochastic partial differential equations in turbulence related problems
,”
Probab. Anal. Relat. Top.
1
,
1
43
(
1978
).
8.
Da Prato
,
G.
and
Zabczyk
,
J.
,
Stochastic Equations in Infinite Dimensions
(
Cambridge University Press
,
Cambridge
,
1992
).
9.
Eyink
,
G. L.
and
Xin
,
J.
, “
Existence and uniqueness of L2-solutions at zero-diffusivity in the Kraichnan model of a passive scalar
,” arXiv:chao-dyn/9605008 (
1996
).
10.
Eyink
,
G. L.
and
Xin
,
J.
, “
Self-similar decay in the Kraichnan model of a passive scalar
,”
J. Stat. Phys.
100
(
3–4
),
679
741
(
2000
).
11.
Flandoli
,
F.
, “
An introduction to 3D stochastic fluid dynamics
,” in
SPDE in Hydrodynamic: Recent Progress and Prospects
, edited by
Da Prato
,
G.
and
Röckner
,
M.
(
Springer
,
Berlin
,
2008
), pp.
51
150
.
12.
Flandoli
,
F.
,
Random Perturbation of PDEs and Fluid Dynamic Models
,
Lecture Notes in Mathematics Vol. 2015
(
Springer
,
Berlin
,
2011
).
13.
Flandoli
,
F.
,
Galeati
,
L.
, and
Luo
,
D.
, “
Eddy heat exchange at the boundary under white noise turbulence
,”
Philos. Trans. R. Soc. A
380
,
20210096
(
2022
).
14.
Flandoli
,
F.
and
Gubinelli
,
M.
, “
Statistics of a vortex filament model
,”
Electron. J. Probab.
10
,
865
900
(
2005
).
15.
Flandoli
,
F.
,
Gubinelli
,
M.
,
Giaquinta
,
M.
, and
Tortorelli
,
V. M.
, “
Stochastic currents
,”
Stochastic Process. Appl.
115
(
9
),
1583
1601
(
2005
).
16.
Flandoli
,
F.
,
Gubinelli
,
M.
, and
Russo
,
F.
, “
On the regularity of stochastic currents, fractional Brownian motion and applications to a turbulence model
,”
Ann. Inst. Henri Poincare
45
(
2
),
545
576
(
2009
).
17.
Galeati
,
L.
, “
On the convergence of stochastic transport equations to a deterministic parabolic one
,”
Stochastics Partial Differ. Equations: Anal. Comput.
8
(
4
),
833
868
(
2020
).
18.
Giaquinta
,
M.
,
Modica
,
G.
, and
Souček
,
J.
,
Cartesian Currents in the Calculus of Variations I: Cartesian Currents
(
Springer-Verlag
,
Berlin
,
1998
).
19.
Grotto
,
F.
, “
Stationary solutions of damped stochastic 2-dimensional Euler’s equation
,”
Electron. J. Probab.
25
,
1
24
(
2020
).
20.
Hytönen
,
T.
,
van Neerven
,
J.
,
Veraar
,
M.
, and
Weis
,
L.
,
Analysis in Banach Spaces. Volume I: Martingales and Littlewood-Paley Theory
(
Springer
,
Berlin
,
2016
).
21.
Kraichnan
,
R. H.
, “
Inertial ranges in two-dimensional turbulence
,”
Phys. Fluids
10
(
7
),
1417
1423
(
1967
).
22.
Kraichnan
,
R. H.
, “
Small-scale structure of a scalar field convected by turbulence
,”
Phys. Fluids
11
,
945
953
(
1968
).
23.
Kraichnan
,
R. H.
, “
Anomalous scaling of a randomly advected passive scalar
,”
Phys. Rev. Lett.
72
,
1016
(
1994
).
24.
Kuksin
,
S. B.
and
Shirikyan
,
A.
,
Mathematics of Two-Dimensional Turbulence
(
Cambridge University Press
,
Cambridge
,
2012
).
25.
Lions
,
P.-L.
and
Majda
,
A.
, “
Equilibrium statistical theory for nearly parallel vortex filaments
,”
Commun. Pure Appl. Math.
53
,
76
142
(
2000
).
26.
Lodhia
,
A.
,
Sheffield
,
S.
,
Sun
,
X.
, and
Watson
,
S. S.
, “
Fractional Gaussian fields: A survey
,”
Probab. Surv.
13
,
1
56
(
2016
).
27.
Marchioro
,
C.
and
Pulvirenti
,
M.
,
Mathematical Theory of Incompressible Nonviscous Fluids
,
Applied Mathematical Sciences Vol. 96
(
Springer-Verlag
,
New York
,
1994
).
28.
Métivier
,
M.
,
Stochastic Partial Differential Equations in Infinite Dimensional Spaces
(
Quaderni Scuola Normale Superiore
,
Pisa
,
1988
).
29.
Prévôt
,
C.
and
Röckner
,
M.
,
A Concise Course on Stochastic Partial Differential Equations
,
Lecture Notes in Mathematics Vol. 1905
(
Springer
,
Berlin
,
2007
).
30.
Rebolledo
,
R.
, “
La méthode des martingales appliquée à la convergence en loi des processus
,”
Mem. Soc. Math. France
62
,
130
(
1979
).
31.
Rozovsky
,
B. L.
and
Lototsky
,
S.
,
Stochastic Evolution Systems, Linear Theory and Applications to Non-linear Filtering
(
Springer
,
Berlin
,
2018
).
32.
Temam
,
R.
,
Navier-Stokes Equations
(
North-Holland Publishing Company
,
1977
).
33.
Vishik
,
M. J.
and
Fursikov
,
A. V.
,
Mathematical Problems of Statistical Hydromechanics
(
Kluwer
,
Boston
,
1988
).