Processing math: 100%
Research article Special Issues

Sub-exponential mixing of generalized cellular flows with bounded palenstrophy

  • We study the mixing properties of a passive scalar advected by an incompressible flow. We consider a class of cellular flows (more general than the class in [Crippa-Schulze M3AS 2017]) and show that, under the constraint that the palenstrophy is bounded uniformly in time, the mixing scale of the passive scalar cannot decay exponentially.

    Citation: Gianluca Crippa, Christian Schulze. Sub-exponential mixing of generalized cellular flows with bounded palenstrophy[J]. Mathematics in Engineering, 2023, 5(1): 1-12. doi: 10.3934/mine.2023006

    Related Papers:

    [1] Francesca Tedeschi, Giulio G. Giusteri, Leonid Yelash, Mária Lukáčová-Medvid'ová . A multi-scale method for complex flows of non-Newtonian fluids. Mathematics in Engineering, 2022, 4(6): 1-22. doi: 10.3934/mine.2022050
    [2] Giuseppe Procopio, Massimiliano Giona . Bitensorial formulation of the singularity method for Stokes flows. Mathematics in Engineering, 2023, 5(2): 1-34. doi: 10.3934/mine.2023046
    [3] Paola F. Antonietti, Chiara Facciolà, Marco Verani . Unified analysis of discontinuous Galerkin approximations of flows in fractured porous media on polygonal and polyhedral grids. Mathematics in Engineering, 2020, 2(2): 340-385. doi: 10.3934/mine.2020017
    [4] Yangyang Qiao, Qing Li, Steinar Evje . On the numerical discretization of a tumor progression model driven by competing migration mechanisms. Mathematics in Engineering, 2022, 4(6): 1-24. doi: 10.3934/mine.2022046
    [5] Andrea Manzoni, Alfio Quarteroni, Sandro Salsa . A saddle point approach to an optimal boundary control problem for steady Navier-Stokes equations. Mathematics in Engineering, 2019, 1(2): 252-280. doi: 10.3934/mine.2019.2.252
    [6] Ivan Fumagalli . Discontinuous Galerkin method for a three-dimensional coupled fluid-poroelastic model with applications to brain fluid mechanics. Mathematics in Engineering, 2025, 7(2): 130-161. doi: 10.3934/mine.2025006
    [7] Jason Howell, Katelynn Huneycutt, Justin T. Webster, Spencer Wilder . A thorough look at the (in)stability of piston-theoretic beams. Mathematics in Engineering, 2019, 1(3): 614-647. doi: 10.3934/mine.2019.3.614
    [8] Mouhamed Moustapha Fall, Veronica Felli, Alberto Ferrero, Alassane Niang . Asymptotic expansions and unique continuation at Dirichlet-Neumann boundary junctions for planar elliptic equations. Mathematics in Engineering, 2019, 1(1): 84-117. doi: 10.3934/Mine.2018.1.84
    [9] Yoshikazu Giga, Hirotoshi Kuroda, Michał Łasica . The fourth-order total variation flow in $ \mathbb{R}^n $. Mathematics in Engineering, 2023, 5(6): 1-45. doi: 10.3934/mine.2023091
    [10] Weimin Sheng, Shucan Xia . Interior curvature bounds for a type of mixed Hessian quotient equations. Mathematics in Engineering, 2023, 5(2): 1-27. doi: 10.3934/mine.2023040
  • We study the mixing properties of a passive scalar advected by an incompressible flow. We consider a class of cellular flows (more general than the class in [Crippa-Schulze M3AS 2017]) and show that, under the constraint that the palenstrophy is bounded uniformly in time, the mixing scale of the passive scalar cannot decay exponentially.



    We are interested in the mixing properties of passive scalars advected by incompressible flows. How well a scalar is mixed by a flow is an important problem in fluid mechanics, with applications to atmospheric and oceanographic science, biology, and chemistry, just to name a few. In many situations the advected scalar has a negligible feedback on the underlying flow, and mathematically this results in the fact that the scalar solves a linear continuity equation with a given velocity field (see (1.1)).

    Let us now specify our mathematical setting. We consider a binary passive scalar ρ{+1,1} advected by a time-dependent, divergence-free velocity field u on the two dimensional open unit square Q=(12,12)2R2, with u=0 on Q. Given an initial condition ˉρ, the scalar ρ satisfies the Cauchy problem for the continuity equation with velocity field u, that is, it solves

    {tρ+div(uρ)=0onR+×R2ρ(0,)=ˉρonR2. (1.1)

    Outside of Q, both the velocity field u and the solution ρ are identically zero for all times. Without loss of generality we can assume that ˉρ satisfies Qˉρ=0 (otherwise we can simply consider ˜ρ=ρˉρ) and we observe that such a condition will be satisfied by the solution ρ(t,) at all times. The results presented in this paper hold in all dimensions d2 as well with the necessary changes in the scaling analysis. For the construction of examples, d=2 is the most restrictive case.

    Mixing scales. In order to mathematically study the problem of optimal mixing we first need to specify how we measure the "degree of mixedness" of the solution ρ at each time. The resulting "mixing scales" vary greatly in the literature and depend not only on the field of mathematics, but also on the specific PDE under consideration. The presence of diffusion, for instance, changes the mixing process drastically, and the L2 norm is often a suitable mixing scale in the diffusive setting. However, the L2 norm is conserved in the inviscid equation (1.1) and therefore it is not a good measure of the mixing scale. One needs instead to find a quantification of the (possible) rate of weak convergence to zero of the solution. Two notions of mixing scale in such a case are available in the literature, namely the geometric and the functional mixing scale.

    Definition 1.1 (Geometric mixing scale [3]). Given an accuracy parameter 0<κ<1, the geometric mixing scale of ρ(t,) is the infimum of all ϵ>0 such that for every xR2 it holds

    1ρ(t,)|B(x,ϵ)ρ(t,y)dy|<κ. (1.2)

    Note that in the binary case ρ{+1,1} the above condition is equivalent to

    1κ2<|{ρ(t,)=+1}B(x,ϵ)||B(x,ϵ)|<1+κ2. (1.3)

    We denote by G(ρ(t,)) such infimum.

    Definition 1.2 (Functional mixing scale [8,11,13]). The functional mixing scale of ρ(t,) is

    ρ(t,)˙H1(R2)=sup{R2ρ(t,x)ξ(x)dx:ξL2(R2)1}. (1.4)

    Notation. It is often convenient to write mix(ρ(t,)) to denote any of the two mixing scales, in case a statement holds for both of them.

    It is worth mentioning that the two mixing scales are strongly related, but not equivalent, see the examples in [9].

    Optimal mixing. The central question in optimal mixing is how fast the passive scalar can be mixed under suitable (energetic) constraints on the velocity field u, that is, how fast (as a function of the time) the mixing scales can decay to zero. Answering this question entails two parts: on the one hand, one establishes lower bounds on the decay of the mixing scales, and on the other hand one exhibits (explicit) examples of velocity fields that mix optimally under such a constraint, that is, examples saturating the lower bounds.

    To illustrate one particular, yet very relevant, example, let us focus on the case of a fixed enstrophy constraint, that means u(t,)L2C for all t0. Under this constraint, Crippa and De Lellis [4] proved an exponential lower bound for the geometric mixing scale. This result was later extended by Iyer, Kiselev and Xu [7] and by Seis [12] to the functional mixing scale, so that we have

    G(ρ(t,))cexp(ct)andρ(t,)˙H1cexp(ct), (1.5)

    where the constant c depends on suptu(t,)L2 and on the initial datum ˉρ.

    For future use, we record here that the proof of the lower bounds in (1.5) is based on a key lemma from [4] showing Lusin-Lipschitz regularity of the flow of a Sobolev velocity field, in the form of a quantitative bound:

    Theorem 1.3 (Crippa and De Lellis). Let X:[0,T]×R2R2 be the flow associated to a velocity field in L1([0,T];W1,p(R2)), with p>1. Then, for every ϵ>0 and every R>0, we can find a set KB(0,R) such that |B(0,R)K|ϵ and for any 0tT we have

    Lip(X(t,)|K)exp(cpt0u(s,)Lpdsϵ1/p). (1.6)

    The above result somewhat extends the Cauchy-Lipschitz theory to the Sobolev framework.

    The exponential lower bounds in (1.5) turn out to be sharp. Indeed, Alberti, Crippa and Mazzucato [1,2] and Yao and Zlatoš [14] constructed explicit examples of entrophy-constrained velocity fields exhibiting exponential mixing. Both examples are of cellular type, a special structure that plays an important role in our analysis and that we describe in the next paragraph.

    Cellular structure. The rough idea of the cellular structure is sketched in Figure 1. We start with a binary initial condition ρ0 with zero average on the 2-dimensional square Q. In a first step we subdivide Q into 4 disjoint sub-squares D1,,D4 of equal size. We then construct a velocity field u0 in such a way that the solution ρ(1,) has zero average on each of the sub-squares, which means that the tracer gets equally distributed among D1,,D4, as schematically visualized in Figure 1.

    Figure 1.  First two steps of a cellular flow.

    We continue inductively by subdividing each Di into four equal sub-squares and distributing the tracer equally among those sub-squares with tracer movements localized in Di respectively, and so forth. In a sense, in each step the problem of mixing on a large area is transferred into the problem of separately mixing on four smaller areas.

    Using a cellular type structure for the construction of mixing flows has many advantages. In addition to the convenient inductive nature of the construction, a crucial benefit of this structure is that it allows to keep track of both the geometric and the functional mixing scale at all times by simple scaling arguments, as we will see later in this paper.

    Fixed palenstrophy constraint. In this paper we deal with a fixed palenstrophy constraint, by which we mean a uniform-in-time bound on the ˙Ws,p norm of the velocity field u, where s>1 and 1<p (with a slight abuse of terminology, since more usually the term "palenstrophy" denotes the ˙H2 norm of the velocity field u). We refer to [9,10] for more details on the concept of palenstrophy and its significance in the context of mixing of passive scalars. Under a fixed palenstrophy constraint we immediately inherit the exponential lower bounds (1.5) for both mixing scales from the fixed enstrophy case due to the (fractional) Poincaré inequality

    u(t,)˙Ws,p(R2)Cs,pu(t,)˙W1,p(R2). (1.7)

    Elgindi and Zlatoš [6] proved that this lower bound is sharp for a certain range of s,p>1. The authors prove that a variation of Baker's map (a discrete map well-known in the theory of dynamical systems) is associated to a flow. The action of the velocity field u on the time interval [0,1] is sketched in Figure 2. The velocity field is then periodically extended and therefore it is not of cellular type, since the same velocity field is repeated periodically in time without going to smaller scales.

    Figure 2.  Action of the velocity field u by Elgindi and Zlatoš for 0t1.

    It is in fact believed that the exponential bound is sharp in the full range s,p>1, but no analytic examples are available (see also the numerical simulations in [8]).

    We further notice that it is possible to rescale (in time) the cellular flows in [2] and [14] in such a way that they satisfy the fixed palenstrophy constraint. However, by doing so, the rescaled flows mix at a slower, polynomial rate. It is therefore natural to ask whether such "reduced mixing ability" is specific to these particular examples (which are not only cellular, but in fact have a self-similar character), or if it applies to all cellular examples under this constraint and is due to the increased localization of the action of the flow.

    We partially answered this question in [5], where we considered cellular flows based on "building blocks" that do not mix "too well" at every iteration. To be more specific, consider again the sketch in Figure 1. Note that, in every iteration, the side-length of the cells decays by a fixed factor 1/2. In [5] we considered cellular flows for which also the mixing scale decays at most by a factor 1/2 in every iteration, as it is the case in Figure 1. As a consequence, in the context of [5], after every iteration the ratio between the mixing scale and the side length of the cells is constant. Note that the cellular examples in [2] and [14] follow this rule and therefore fall within the setting of [5]. Under this assumption, we proved that both mixing scales cannot decay faster than polynomially under a fixed palenstrophy constraint.

    The class considered in [5] is meaningful as it contains several explicit examples of mixers (for instance, self-similar and quasi self-similar ones, see again [2]). However, the requirement on the constant ratio between the cell size and the mixing scale is quite restrictive and it is not evident whether the sub-exponential lower bound in [5] is due to this condition or, more generally, it holds as a consequence of the increased localization of the velocity field. The natural questions that arise following this line of thinking are: What happens if we allow a cellular velocity field to mix arbitrarily well (with respect to the cell size) at every iteration? Is this restriction (weaker than the one in [5]) strong enough to exclude exponential mixing under fixed palentstrophy?

    In the present article we simplify both the setup and the proof in [5] and completely drop the restriction preventing the building blocks from mixing "too well", still showing that exponential mixing cannot happen. We describe in detail the generalized cellular structure we are able to deal with in Section 2, while in Section 3 we state and prove our Main Theorem. In order to make the comparison with the statement in [5] more explicit, we underline that in Definition 2.6(ⅱ) we assume that, after n iterations, the solution is mixed at scale λn+σ(n), where the increasing function σ:NN may be unbounded. The result in [5] corresponds to the case σ(n)=0, as we comment in Remark 2.7(ⅲ). The case of an unbounded function σ corresponds to a mixing scale which, after n steps, is strictly finer than λn.

    Fix λ>0 such that 1/λ is an integer greater or equal than two. We define the tiling of Q=(12,12)2 corresponding to the parameter λ as follows.

    Definition 2.1. We denote by Tλ the tiling of Q with squares of side λ, consisting of the 1/λ2 open squares of the form

    {(x,y)Q:12+kλ<x<12+(k+1)λand12+hλ<y<12+(h+1)λ}

    for k,h=0,,1/λ1.

    Let (Tn)nN be an arbitrary (increasing) sequence of times. We define the generalized cellular structure in the following way. For times Tnt<Tn+1, we assume that in any tile QTλn we can write the velocity field u and corresponding solution ρ as

    u(t,x)=λnTn+1TnuQ,n(tTnTn+1Tn,xrQλn)andρ(t,x)=ρQ,n(tTnTn+1Tn,xrQλn),

    where (uQ,n,ρQ,n) solves (1.1) for 0t1 and rQ denotes the center of QTλn. Furthermore, we assume that ρQ,n(0,) is both mixed and un-mixed at scale λσ(n), where (σ(n))nNN is an increasing sequence of positive integers. We specify our notion of mixed and un-mixed in Definition 2.4, after a short digression on some background material.

    We adapt the notion of un-mixed from [5]. The idea is to call a binary function ρ un-mixed at scale r if there exists a ball of radius r that violates the geometric mixing scale condition in (1.3) significantly. For this we fix a parameter ˉγ(0,1). This parameter will be fixed for the rest of this section and for this reason explicit dependencies of constants on ˉγ will often not be stated. In the following, 0<κ<1 is an accuracy parameter as in Definition 1.1.

    Definition 2.2 (Characteristic length scale). Let a set JE of positive measure be given. We denote by C=C(E,J) the set of all admissible balls B(x,r)E such that

    |JB(x,r)||B(x,r)|>11κ2ˉγ. (2.1)

    We define the characteristic length scale of the set J with respect to E as

    LSE(J)=sup{r>0 such that there exists B(x,r)C}.

    Remark 2.3. This scale determines the largest radius of a ball which violates significantly (that is, with the uniform gap ˉγ) condition (1.3) of the geometric mixing scale for a binary tracer ρ{+1,1} and will serve as our measurement of how un-mixed the tracer is. To see this, let us set ˉγ=1 and

    J={ρ(t,)=+1}. (2.2)

    Note that (2.1) becomes

    |{ρ(t,)=+1}B(x,r)||B(x,r)|>1+κ2, (2.3)

    and hence B(x,r) violates condition (1.3) of the geometric mixing scale. For ˉγ(0,1), it violates the condition significantly. Notice that the notion of characteristic length scale becomes weaker as the parameter ˉγ approaches the value 1.

    Concerning our notion of being mixed at scale λk, we simply require that the solution ρ has zero average on all cells QTλk. We summarize our terminology in the following definition:

    Definition 2.4 (Mixed and un-mixed). Let λ>0 be a tiling parameter and α(0,1) a universal constant. We say that a function ρ is both mixed and un-mixed at scale λk if

    (i) Qρ(x)dx=0 for all QTλk and

    (ii) On every QTλk, the function ρ has the form

    ρ|Q={1onAQ1onAcQ

    where LSQ(AQ)αλk.

    The following elementary lemma shows that, if a function is mixed at scale λ according to Definition 2.4(ⅰ), then both the geometric and the functional mixing scale are smaller than Cλk. See the Appendix of [5] for its proof.

    Lemma 2.5. Let ρ be a bounded function such that

    Qρdy=0 (2.4)

    for every tile QTλ. Then there exist constants C1=C1(κ) and C2=C2(ρ) such that

    (i) G(ρ)C1λ,

    (ii) ρ˙H1C2λ.

    In the following definition we summarize our notion of generalized cellular structure.

    Definition 2.6 (Generalized cellular structure). We say that a pair of a velocity field u with corresponding solution ρ to the continuity equation (1.1) is of (λ,α,ˉγ,s,p,(Tn)n,(σ(n))n)-generalized cellular type, or simply of generalized cellular type when parameters are fixed, if on any time interval Tnt<Tn+1 and for any cell QTλn we can write u and ρ as

    u(t,x)=λnTn+1TnuQ,n(tTnTn+1Tn,xrQλn)andρ(x,t)=ρQ,n(tTnTn+1Tn,xrQλn), (2.5)

    where rQ is the center of the tile Q and (uQ,n,ρQ,n) is a solution to (1.1) on [0,1]×Q, such that

    (i) uQ,nLt(˙Ws,px) on the time interval 0t1 and is divergence free, u0=0 on Q and zero outside of Q,

    (ii) ρQ,n(0,) is both mixed and un-mixed at scale λσ(n).

    Remark 2.7. Note that Definition 2.6 implies the following facts.

    (i) At time Tn, the solution ρ is both mixed and un-mixed at scale λn+σ(n) (see Figure 3).

    Figure 3.  Example of ρ at iteration step n.

    (ii) Since (u,ρ) solves (1.1) and exploiting (2.5), it follows that ρQ,n(1,) is both mixed and un-mixed at scale λ1+σ(n+1).

    (iii) Definition 2.6 is a generalization of the cellular structure in [5], which corresponds to the case σ(n)=0 for all nN.

    In this section we state and prove our main theorem, which asserts that under a fixed palenstrophy constraint it is not possible to mix exponentially fast by using a generalized cellular structure.

    Main Theorem. Let (u,ρ) be a solution of generalized cellular type (as in Definition 2.6) of the continuity equation (1.1) such that u(t,)˙Ws,p is bounded uniformly in time, for some s,p>1. Then the decay of both the geometric and the functional mixing scale is slower than exponential.

    Proof. We divide the proof in two steps.

    Minimal cost estimate. Let us fix nN and a cell QTλn. Let kN to be chosen later. Note that by Definition 2.6, by concatenating on the time intervals [Tn,Tn+1], [Tn+1,Tn+2], , [Tn+k1,Tn+k] solutions as in (2.5), on [Tn,Tn+k]×Q the solution (u,ρ) has the form

    u(t,x)=λnTn+kTnuQ,n,k(tTnTn+kTn,xrQλn)andρ(x,t)=ρQ,n,k(tTnTn+kTn,xrQλn),

    where ρQ,n,k(0,):QR is both mixed and un-mixed at scale λσ(n) and ρQ,n,k(1,) is both mixed and un-mixed at scale λk+σ(n+k). We will prove the following minimal cost estimate

    c1log(¯c2mix(ρQ,n,k(0,))mix(ρQ,n,k(1,)))c1log(c2λk+σ(n+k)σ(n))=Mn,k10uQ,n,k(t,)Lp(Q)dt (3.1)

    for k large enough. The main ingredient for this step is the regularity result by Crippa and De Lellis [4] in Theorem 1.3.

    Let A={xsuchthatρQ,n,k(0,x)=1}. Since ρQ,n,k(0,) is un-mixed at scale λσ(n), for every QTλσ(n) there exists a ball BQ=BQ(x,r)Q, such that rωλσ(n)α (where ω(0,1) will be chosen later depending on ˉγ) and

    |ABQ||BQ|>11κ2ˉγ1ˉγ2, (3.2)

    with α the constant in Definition 2.4. For every QTλσ(n) we consider the ball ˜BQ=B(x,σ)BQ with the same center as BQ and such that |˜BQ|=((3+ˉγ)/4)|BQ|. Thus

    |A˜BQ||ABQ||BQ˜BQ|(3ˉγ4)|BQ|=(3ˉγ4)πr2(3ˉγ4)πω2λ2σ(n)α2. (3.3)

    We now define

    B=QTλσ(n)BQand˜B=QTλσ(n)˜BQ. (3.4)

    Since the balls BQ are pairwise disjoint, by (3.2) it follows that

    |AB||B|>11κ2ˉγ1ˉγ2, (3.5)

    and by (3.3) we have that

    |A˜B|(3ˉγ4)πω2α2. (3.6)

    Furthermore, note that there exists a constant 1>C=C(ˉγ)>0 (in fact, C(ˉγ)=1(3+ˉγ)/4) such that

    dist(A˜B,Bc)C(ˉγ)rωC(ˉγ)λσ(n)α, (3.7)

    which we need later on in the proof.

    We now focus on ρQ,n,k at time t=1. We set Φ=X(1,), where X:[0,1]×QQ is the flow associated to uQ,n,k. For notational purposes, we denote by δ=λk+σ(n+k). Let {Qi}1/δ2i=1 denote the sub-squares in the tiling Tδ. For i=1,,1/δ2, denote

    Ai:=Φ(A˜B)Qi.

    Note that the Ai's are disjoint for different i's and by (3.6) we have that

    1/δ2i=1|Ai|=|A˜B|(3ˉγ4)ω2α2π. (3.8)

    Let Gi:=Φ(Ac)Qi. We further decompose each Gi into Ggoodi and Gbadi, where

    Ggoodi:=Φ(Ac)QiΦ(Bc)andGbadi:=Φ(Ac)QiΦ(B).

    Recall that (3.5) yields that

    1/δ2i=1|Gbadi|=|AcB|ˉγ2α2π, (3.9)

    which we will use later in the proof.

    By Theorem 1.3, for any η>0, there exists EQ with |E|<η, such that

    Lip(Φ1|Ec)exp(Cη1/p10uQ,n,k(s,)Lpds). (3.10)

    Let us fix

    η=(1ˉγ)316α2π (3.11)

    and let E be the corresponding set as in (3.10). For such E, we claim that there exists some index i{1,,1/δ2}, such that both AiE and GgoodiE are nonempty. Once this is proved, taking any points ˜xAiE and yGgoodiE yields

    Φ1(˜x)A˜BandΦ1(y)AcBc.

    Note that by (3.7) this implies that |Φ1(˜x)Φ1(y)|Cλσ(n)α (where C depends on ˉγ and ω). On the other hand, since ˜x,yQi we have that |˜xy|2δ=2λk+σ(n+k). Hence, by Theorem 1.3, this yields

    Cαλk+σ(n+k)σ(n)|Φ1(˜x)Φ1(y)||˜xy|Lip(Φ1|Ec)exp(Cη1/p10uQ,n,k(s,)Lpds).

    Choosing k large enough, such that the left hand side is larger than one, by the above inequality and by our particular choice of η in (3.11) there exist positive constants c1,c2 depending on ˉγ and α only (which are in turn fixed parameters) such that

    Mn,k=c1log(c2λk+σ(n+k)σ(n))10uQ,n,k(s,)Lpds. (3.12)

    Noting that, by Remark 2.7(ⅰ), there exist constants C1 and C2 such that

    mix(ρQ,n,k(0,))C1λσ(n)andmixρQ,n,k(1,))C2λk+σ(n+k), (3.13)

    by (3.12) and setting ¯c2=c2C2/C1 we obtain the minimal cost estimate (3.1).

    It only remains to prove the claim. For each i=1,,1/δ2, note that

    min{|AiE|,|GgoodiE|}min{|Ai|,|Gi|}|Gbadi||QiE|=|Ai||Gbadi||QiE|. (3.14)

    Summing (3.14) up for i=1,,1/δ2, by (3.8), (3.9) and (3.11) and setting ω2=(1+ˉγ)/2 this yields

    1/δ2i=1min{|AiE|,|GgoodiE|}|A˜B||AcB||E|[(3ˉγ4)ω2ˉγ2(1ˉγ)316]α2π[34ω23ˉγ4(1ˉγ)316]α2π =[1ˉγ]316α2π>0,

    hence the claim is proved.

    Scaling computation. We compute sxu(t,)Lp(Q) on the time interval [Tn,Tn+k), where k=k(n) is as in the previous step. For notational convenience we denote |[Tn,Tn+k]|=τn,k. Remember that for any tile QTλn the velocity field u has the form

    u(t,x)|Qλnτn,kuQ,n,k(tTnτn,k,xrQλn).

    We will perform the following step for the case sN. The computation for the case sRN is analogous, however due to the non-locality of the fractional derivative the first equality in (3.15) changes to "". For more details, we refer to [5], where this step is performed in detail for the case sRN. Using the change of variables, the Poincaré inequality, and the Jensen inequality we get

    Tn+kTnsu(t,)pLp(Q)dt=Tn+kTnQTλnQ|sx[λnτn,kuQ,n,k(tTnτn,k,xrQλn)]|pdxdt=1λ(s1)npτn,kτpn,kQTλn10Q|(sxuQ,n,k)(t,xrQλn)|pdxdt=τn,k(λs1)npτpn,kλ2nQTλn10Q|(sxuQ,n,k)(t,x)|pdxdtPCp,sτn,k(λs1)npτpn,kλ2nQTλn10Q|(xuQ,n,k)(t,x)|pdxdtJCp,sτn,k(λs1)npτpn,kMpn,k, (3.15)

    where Mn,k is as in (3.1), relative to the basic building blocks uQ,n,k used on [Tn,Tn+k]. Now since |[Tn,Tn+1]|=τn,k, together with (3.15) we get that

    sxu(t,)pLp(Q)Lt(Tn,Tn+1)(1λs1)np1τpn,kCp,sMpn,k. (3.16)

    Taking the p-th root of the above inequality, using the minimal cost estimate (3.12) and the fact that the Ws,p norm is uniformly bounded in time, this implies that

    C(1λs1)n1τn,kCp,sc1log(c2λk+σ(n+k)σ(n)). (3.17)

    Note that there exist constants C1 and C2 such that

    mix(ρ(Tn,))C1λn+σ(n)andmix(ρ(Tn+k,))C2λn+k+σ(n+k), (3.18)

    which implies

    mix(ρ(Tn,))mix(ρ(Tn+k,))Cλn+σ(n)λn+k+σ(n+k)=Cλk+σ(n+k)σ(n). (3.19)

    Using (3.19) in (3.17) and denoting β=λs1(0,1) this yields

    exp[Cβn(Tn+kTn)]Cmix(ρ(Tn,))mix(ρ(Tn+k,)). (3.20)

    By Theorem 1.3 and the fact that the velocity field u is bounded in W1,p uniformly in time we have

    mix(ρ(Tn))γ1exp(α1Tn) (3.21)

    for all nN. Now let us assume that the mixing scale of ρ decays exponentially in time, i.e. that there exists α2>0, such that

    mix(ρ(Tn))γ2exp(α2Tn) (3.22)

    for all nN. By (3.21) we must have α1α2. Then, by (3.20), (3.21) and (3.22) it follows that

    γ2exp(α2Tn+k)mix(ρ(Tn+k,))Cmix(ρ(Tn,))exp[Cβn(Tn+kTn)]Cexp[Cβn(Tn+kTn)]γ1exp(α1Tn)=Cγ1exp[(Cβn(Tn+kTn)+α1Tn)] (3.23)

    and hence

    ˉCexp[(α2Cβn)Tn+k(α1Cβn)Tn] (3.24)

    for a positive constant ˉC. This is however a contradiction: if we choose n large enough in such a way that both (α2Cβn)>0 and (α1Cβn)>0, we can then let k go to infinity (which in turn lets Tn+k go to infinity) and we then see that the right hand side of (3.24) goes to infinity.

    This research has been partially supported by the ERC Starting Grant 676675 FLIRT.

    The authors declare no conflict of interest.



    [1] G. Alberti, G. Crippa, A. Mazzucato, Exponential self-similar mixing and loss of regularity for continuity equations, C. R. Math., 352 (2014), 901–906. http://dx.doi.org/10.1016/j.crma.2014.08.021 doi: 10.1016/j.crma.2014.08.021
    [2] G. Alberti, G. Crippa, A. Mazzucato, Exponential self-similar mixing by incompressible flows, J. Amer. Math. Soc., 32 (2019), 445–490. http://dx.doi.org/10.1090/jams/913 doi: 10.1090/jams/913
    [3] A. Bressan, A lemma and a conjecture on the cost of rearrangements, Rend. Sem. Mat. Univ. Padova, 110 (2003), 97–102.
    [4] G. Crippa, C. De Lellis, Estimates and regularity results for the DiPerna Lions flow, J. Reine Angew. Math., 616 (2008), 15–46. http://dx.doi.org/10.1515/CRELLE.2008.016 doi: 10.1515/CRELLE.2008.016
    [5] G. Crippa, C. Schulze, Cellular mixing with bounded palenstrophy, Math. Models Methods Appl. Sci., 27 (2017), 2297–2320. http://dx.doi.org/10.1142/S0218202517500452 doi: 10.1142/S0218202517500452
    [6] T. M. Elgindi, A. Zlatoš, Universal mixers in all dimensions, Adv. Math., 356 (2019), 106807. http://dx.doi.org/10.1016/j.aim.2019.106807 doi: 10.1016/j.aim.2019.106807
    [7] G. Iyer, A. Kiselev, X. Xu, Lower bounds on the mix norm of passive scalars advected by incompressible enstrophy-constrained flows, Nonlinearity, 27 (2014), 973–985. http://dx.doi.org/10.1088/0951-7715/27/5/973 doi: 10.1088/0951-7715/27/5/973
    [8] Z. Lin, J.-L. Thiffeault, C. R. Doering, Optimal stirring strategies for passive scalar mixing, J. Fluid Mech., 675 (2011), 465–476. http://dx.doi.org/10.1017/S0022112011000292 doi: 10.1017/S0022112011000292
    [9] E. Lunasin, Z. Lin, A. Novikov, A. Mazzucato, C. R. Doering, Optimal mixing and optimal stirring for fixed energy, fixed power, or fixed palenstrophy flows, J. Math. Phys., 53 (2012), 115611. http://dx.doi.org/10.1063/1.4752098 doi: 10.1063/1.4752098
    [10] E. Lunasin, Z. Lin, A. Novikov, A. Mazzucato, C. R. Doering, Erratum: Optimal mixing and optimal stirring for fixed energy, fixed power, or fixed palenstrophy flows, J. Math. Phys., 54 (2013), 079903. http://dx.doi.org/10.1063/1.4816334 doi: 10.1063/1.4816334
    [11] G. Mathew, I. Mezic, L. Petzold, A multiscale measure for mixing, Physica D, 211 (2005), 23–46. http://dx.doi.org/10.1016/j.physd.2005.07.017 doi: 10.1016/j.physd.2005.07.017
    [12] C. Seis, Maximal mixing by incompressible fluid flows, Nonlinearity, 26 (2013), 3279–3289. http://dx.doi.org/10.1088/0951-7715/26/12/3279 doi: 10.1088/0951-7715/26/12/3279
    [13] J.-L. Thiffeault, Using multiscale norms to quantify mixing and transport, Nonlinearity, 25 (2012), R1–R44. http://dx.doi.org/10.1088/0951-7715/25/2/R1 doi: 10.1088/0951-7715/25/2/R1
    [14] Y. Yao, A. Zlatoš, Mixing and un-mixing by incompressible flows, J. Eur. Math. Soc., 19 (2014), 1911–1948. http://dx.doi.org/10.4171/JEMS/709 doi: 10.4171/JEMS/709
  • This article has been cited by:

    1. Roberta Bianchini, Chiara Saffirio, Fluid instabilities, waves and non-equilibrium dynamics of interacting particles: a short overview, 2022, 5, 2640-3501, 1, 10.3934/mine.2023033
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1894) PDF downloads(147) Cited by(1)

Figures and Tables

Figures(3)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog