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

Dissipation enhancement of cellular flows in general advection diffusion equations

Yu Feng Yu Feng, Beijing International Center for Mathematical Research, Peking University, No. 5 Yiheyuan Road Haidian District, Beijing, P.R.China 100871 fengyu@bicmr.pku.edu.cn  and  Xiaoqian Xu Xiaoqian Xu: Zu Chongzhi Center for Mathematics and Computational Sciences, Duke Kunshan University, China xiaoqian.xu@dukekunshan.edu.cn
Abstract.

The main contribution of this paper is twofold: (1) Recently, Iyer, Xu, and Zlatoš studied the dissipation enhancement by cellular flows based on standard advection-diffusion equations via a stochastic method. We generalize their results to advective hyper-diffusion equations and advective nonlinear diffusion equations. (2) We prove there exist smooth incompressible flows that are relaxation enhancing to hyper-diffusion but not to standard diffusion.

1. Introduction

Let 𝕋d=[0,1]d\mathbb{T}^{d}=[0,1]^{d} denotes the dd dimensional torus, L02L_{0}^{2} be the mean-zero subspace of L2(𝕋d)L^{2}(\mathbb{T}^{d}) with the inner product ,\langle\cdot,\cdot\rangle, and uu be a divergence-free vector field. Then we consider the following equation on 𝕋d\mathbb{T}^{d}:

(1.1) tθ+uθ+γ(Δ)αθ=0,\partial_{t}\theta+u\cdot\nabla\theta+\gamma\left(-\Delta\right)^{\alpha}\theta=0,

with α,γ>0\alpha,\gamma>0, and initial data θ(0,x)=θ0L02\theta(0,x)=\theta_{0}\in L_{0}^{2}. In particular, for α=1\alpha=1, it gives the standard advection-diffusion equation

(1.2) tθ+uθγΔθ=0.\partial_{t}\theta+u\cdot\nabla\theta-\gamma\Delta\theta=0.

For α=2\alpha=2, (1.1) becomes the advective hyper-diffusion equation

(1.3) tθ+uθ+γΔ2θ=0.\partial_{t}\theta+u\cdot\nabla\theta+\gamma\Delta^{2}\theta=0.

The dissipation enhancement effect has been studied extensively in recent decades. Most of these research is carried out based on (1.2), and can be generalized to (1.1) for α>1\alpha>1 without any difficulty. However, the results obtained from the maximal principle or stochastic process associated with (1.2) cannot be generalized to (1.1) trivially (for example [2, 12, 8, 9]). In this paper, we show the dissipation enhancement result obtained via a stochastic method in [8] can be generalized to (1.1) and nonlinear diffusion equations, in which the stochastic structure is lost.

To begin with, we define the dissipation time associate to (1.1).

Definition 1.1.

Let 𝒮s,tu,α\mathcal{S}_{s,t}^{u,\alpha} be the solution operator to (1.1) on (0,)×𝕋d(0,\infty)\times\mathbb{T}^{d}, that is, for any function θs(x)L02\theta_{s}(x)\in L^{2}_{0}, Ss,tu,αθs(x)S_{s,t}^{u,\alpha}\theta_{s}(x) solves (1.1) with initial data θ(s,x)=θs(x)\theta(s,x)=\theta_{s}(x) on 𝕋d\mathbb{T}^{d}. The dissipation time of uu is defined by

(1.4) τα(u,γ)=definf{t0\nonscript|\nonscript𝒮s,s+tu,αL02L0212 for all s0}.\tau_{\alpha}^{*}(u,\gamma)\stackrel{{\scriptstyle\scriptscriptstyle\textup{def}}}{{=}}\inf\Big{\{}t\geqslant 0\nonscript\>\Big{|}\nonscript\>\mathopen{}\allowbreak\lVert\mathcal{S}_{s,s+t}^{u,\alpha}\rVert_{L_{0}^{2}\to L_{0}^{2}}\leqslant\tfrac{1}{2}\text{ for all $s\geqslant 0$}\Big{\}}\,.

Recent researches indicate that the flows with small dissipation time can be used to suppress the formation of singularities in the nonlinear PDEs (see [8, 3, 7, 5]). In the second-order differential equations, such as the Keller-Segel equations and the ignition type reaction-diffusion equations, one needs to provide flows such that τ1(u,γ)\tau_{1}^{*}(u,\gamma) is small enough [8]. And for the fourth order differential equations, such as Cahn-Hilliard equation [3], Kuramoto-Sivashinsky equation [7] and thin-film equation [5], one need to provide flows such that τ2\tau_{2}^{*} is sufficiently small (see Section 4.1 for more details). Therefore, it is meaningful to construct flows with arbitrarily small dissipation times for either α=1\alpha=1 or 22.

In the seminal paper [1], the main result interpret that given a time-independent velocity field uu, τ1(Au,γ)\tau^{*}_{1}(Au,\gamma) tends to zero as AA tends to infinity if and only if the operator uu\cdot\nabla has no eigenfunctions in H1(𝕋d)H^{1}(\mathbb{T}^{d}) rather than constants. A typical class of such flows is the weakly mixing flows, for which uu\cdot\nabla only has continuous spectrums. The proof of [1] relies on the so-called RAGE theorem, as a consequence, contains no quantitative information on the appearance of a faster time scale induced by uu. After that, the authors in [6, 16] obtain explicit bounds on τ1(u,γ)\tau_{1}^{*}(u,\gamma), by requiring the mixing rate of a mixing flow uu, associated to the underlying transport equation (see [13]). Here the intuition is based on the fact that the mixing flows create small-scale structures (or high frequencies). And the faster this process happens, the faster the energy of solution to (1.1) is dissipated and, as a result, the smaller τ1(u,γ)\tau_{1}^{*}(u,\gamma) becomes. However, known examples of such flows are either complicated or not regular enough.

Fortunately, many smooth flows are not mixing but have small dissipation times, thus enough to suppress singularities in nonlinear PDEs. For such flows, the dissipation times do not tend to zero as one increases the amplitude of the flow. One typical class of such flows is cellular flows (see [8, 10] for applications). A prototypical example of cellular flow in two dimensions is given by

(1.5) u(x)=sin(2πx1)sin(2πx2).u(x)=\nabla^{\perp}\sin(2\pi x_{1})\sin(2\pi x_{2}).

Similar to the example above, all two-dimensional cellular flows have closed trajectories and hence are not mixing. In [8], based on equation (1.2), the authors proved that the dissipation time, τ1\tau^{*}_{1}, of such flows could be made arbitrarily small by rescaling the spacial scale and the flow amplitude at the same time. Their proof is based on establishing the relationship between the dissipation time τ1(u,1)\tau_{1}(u,1) and the effective diffusivity, denote as D(u)D(u) (see Section 2 for a more precise definition). For convenience to the reader, we cite their theorem here.

Theorem 1.2.

For each mm\in\mathbb{N}, let umW1,(𝕋d)u_{m}\in W^{1,\infty}(\mathbb{T}^{d}) be a mean zero divergence free velocity field that is symmetric (condition (2.1) holds) in all coordinates. If the effective diffusivity D(um)D(u_{m}) satisfies limmD(um)=\lim\limits_{m\rightarrow\infty}D(u_{m})=\infty, then there exists lml_{m}\in\mathbb{N}, such that the rescaled velocity fields vm(x)=deflmum(lmx)v_{m}(x)\stackrel{{\scriptstyle\scriptscriptstyle\textup{def}}}{{=}}-l_{m}u_{m}(l_{m}x) satisfy

(1.6) limmτ1(vm,γ)=0.\lim_{m\rightarrow\infty}\tau_{1}^{*}(v_{m},\gamma)=0.

In particular, for d{2,3}d\in\left\{2,3\right\} and any σ>0\sigma>0, there exists a smooth cellular flow uu on 𝕋d\mathbb{T}^{d} such that τ1(u,γ)σ\tau_{1}^{*}(u,\gamma)\leqslant\sigma.

Remark 1.3.

Based on the example (5.9)(5.9) in [8], for instance, when d=2d=2, one can choose vm=m2+αu(mx)v_{m}=m^{2+\alpha}u(mx), with α=15113\alpha=\frac{15}{113}, where u(x)u(x) is given by (1.5), so that (1.6) holds. Similar examples exist in three dimensions. In the recent paper [9], by constructing a successful coupling, Iyer and Zhou improve this result by providing quantitative bounds on the dissipation enhancement in terms of the flow amplitude, cell size, and diffusivity.

Both the proofs in [8] and [9] rely on the stochastic processes associated to equation (1.2), and therefore cannot be generalized to (1.1) directly. On the other hand, in [3] the authors derived an upper bound of τ2(u,γ)\tau_{2}^{*}(u,\gamma) in terms of τ1(u,γ)\tau_{1}^{*}(u,\gamma) for any smooth divergence-free vector field. We cite their result here.

Lemma 1.4.

There exists an explicit dimensional constant CC such that for every divergence free uL([0,);C2(𝕋d))u\in L^{\infty}([0,\infty);C^{2}(\mathbb{T}^{d})), and every γ>0\gamma>0, we have

(1.7) τ2(u,γ)Cτ1(u,γ)(1+uC2τ1(u,γ)).\tau_{2}^{*}(u,\gamma)\leqslant C\tau_{1}^{*}(u,\gamma)(1+\lVert u\rVert_{C^{2}}\tau_{1}^{*}(u,\gamma)).

Once velocity fields with small τ1(u,γ)\tau_{1}^{*}(u,\gamma) are known, one can use Lemma 1.4 to produce velocity fields for which τ2(u,γ)\tau_{2}^{*}(u,\gamma) is small. For instance, if uu is mixing at a sufficiently fast rate, then the results of Wei[15], Coti Zelati et al.[16], Feng and Iyer[6] along with Lemma 1.4 can be used to produce velocity fields for which τ2(u,γ)\tau_{2}^{*}(u,\gamma) is arbitrarily small. However, Lemma 1.4 is not tight enough to produce arbitrary small τ2(u,γ)\tau_{2}^{*}(u,\gamma) for cellular flows. Indeed, with the τ1\tau_{1}^{*} bound of cellular flows in [8] and [9], the right-hand side of (1.7) blows up as uC2\lVert u\rVert_{C^{2}} increases.

There are two main contributions of this paper. Firstly, we show that cellular flows indeed can be used to produce τα(u,γ)\tau_{\alpha}^{*}(u,\gamma) arbitrarily small for any α1\alpha\geqslant 1, that is Theorem 1.2 remains true if we replace τ1\tau_{1}^{*} by τα\tau_{\alpha}^{*} with α>1\alpha>1, the formal conclusion is established in Theorem 3.1. As a byproduct, we also show that cellular flows can be used to enhance dissipation in nonlinear diffusion equations, such as porous medium equations and p-Laplacian equations (see Theorem 4.7 and Theorem 4.7). In the second part of this paper, we prove that there exist smooth incompressible flows that are relaxation enhancing with respect to the advective hyper-diffusion equation but not to the standard advection-diffusion equation (Proposition 5.2).

2. Preliminary

We devote this section to introduce the notations and establish some elementary estimates.

For a function f(x)f(x) in defined in 𝕋d\mathbb{T}^{d}, we define the homogeneous L2L^{2}-Sobolev norm H˙s\dot{H}^{s} (s0s\neq 0) to be

fH˙s=(|n|0|f^(n)|2|n|2s)12,\|f\|_{\dot{H}^{s}}=\left(\sum\limits_{|n|\neq 0}|\hat{f}(n)|^{2}|n|^{2s}\right)^{\frac{1}{2}},

where f^(n)\hat{f}(n) is the nthn^{\text{th}} Fourier coefficient of ff.

We say that a flow uu is symmetric in xnx_{n} if we have

(2.1) u(Rnx)=Rnu(x)for all x𝕋d.u(R_{n}x)=R_{n}u(x)\qquad\text{for all }x\in\mathbb{T}^{d}.

Here Rnv=defv2vnenR_{n}v\stackrel{{\scriptstyle\scriptscriptstyle\textup{def}}}{{=}}v-2v_{n}e_{n} for v=n=1dvnendv=\sum\limits_{n=1}^{d}v_{n}e_{n}\in\mathbb{R}^{d} is the reflection in the nthn^{\text{th}} coordinate. A periodic flow that is symmetric in all dd coordinates is said to have a cellular structure and we call it a cellular flow.

Consider (1.2) on d\mathbb{R}^{d}, with initial data θ0\theta_{0}, and let uu be time-independent, mean-zero, divergence-free, and Lipschitz continuous vector fields. For simplicity, we assume γ=1\gamma=1 from now on. Define the stochastic process Xtx(ω)X_{t}^{x}(\omega):

(2.2) dXtx=2dBtu(Xtx)dt,X0x=x,dX_{t}^{x}=\sqrt{2}dB_{t}-u(X_{t}^{x})dt,\quad X_{0}^{x}=x,

where BtB_{t} is a normalized Brownian motion on d\mathbb{R}^{d} with B0=0B_{0}=0, defined on some probability space (Ω,,Ω)(\Omega,\mathcal{B}_{\infty},\mathbb{P}_{\Omega}). Let Gt(x,y)G_{t}(x,y) denotes the fundamental solution to (1.2) and 𝔼Ω\mathbb{E}_{\Omega} the expectation with respect to ωΩ\omega\in\Omega, then the corresponding solution to (1.2) can be written as θ(t,x)=dGt(x,y)θ0(y)𝑑y=𝔼Ω(θ0(Xtx))\theta(t,x)=\int_{\mathbb{R}^{d}}G_{t}(x,y)\theta_{0}(y)dy=\mathbb{E}_{\Omega}(\theta_{0}(X_{t}^{x})). Then we define the effective diffusivity of uu in the direction ee by

(2.3) De(u)=deflimt𝔼Ω(|(Xtxx)e|22t)(1),D_{e}(u)\stackrel{{\scriptstyle\scriptscriptstyle\textup{def}}}{{=}}\lim_{t\rightarrow\infty}\mathbb{E}_{\Omega}\left(\frac{\lvert\left(X_{t}^{x}-x\right)\cdot e\rvert^{2}}{2t}\right)\qquad(\geqslant 1),

with the limit being independent of xdx\in\mathbb{R}^{d}. In addition, let

(2.4) D(u)=defmin{De1(u),,Ded(u)}1D(u)\stackrel{{\scriptstyle\scriptscriptstyle\textup{def}}}{{=}}\min\left\{D_{e_{1}}(u),\dots,D_{e_{d}}(u)\right\}\geqslant 1

denotes the minimum of effective diffusivities in all the coordinate directions (see [17] for more details). Intuitively, De(u)D_{e}(u) is the average distance a point can go along with the process (2.2).

Let m=defiu(mx)\mathcal{L}_{m}\stackrel{{\scriptstyle\scriptscriptstyle\textup{def}}}{{=}}iu(mx)\cdot\nabla with uu divergence free, note that m\mathcal{L}_{m} is a self-adjoint operator with respect to ,\langle\cdot,\cdot\rangle. Let PcP_{c} and PpP_{p} be the spectral projection on its continuous and discrete spectral subspace correspondingly. On the other hand, we denote by 0<λ1λ20<\lambda_{1}\leqslant\lambda_{2}\leqslant\dots the eigenvalues of the operator Δ-\Delta on the torus, and PNP_{N} be the orthogonal projection on the subspace spanned by its first NN eigen-modes, and by S={θL2:θL2=1}S=\left\{\theta\in L^{2}:\lVert\theta\rVert_{L^{2}}=1\right\} the unit sphere in L2L^{2}. For the convection field with an amplitude A>0A>0, equation (1.1) can be represented in the following way:

(2.5) ddtθmA(t)=iAmθmA(t)(Δ)αθmA(t),\frac{d}{dt}\theta_{m}^{A}(t)=iA\mathcal{L}_{m}\theta_{m}^{A}(t)-(-\Delta)^{\alpha}\theta_{m}^{A}(t),

with initial date θmA(0)=θ0\theta_{m}^{A}(0)=\theta_{0}. Here we assume α1\alpha\geqslant 1. For convenient, we rescale (2.5) in time by the factor ε1=A\varepsilon^{-1}=A, thus we obtain the following equivalent reformulation:

(2.6) ddtθmε(t)=imθmε(t)ε(Δ)αθmε(t),θmε(0)=θ0.\frac{d}{dt}\theta_{m}^{\varepsilon}(t)=i\mathcal{L}_{m}\theta_{m}^{\varepsilon}(t)-\varepsilon(-\Delta)^{\alpha}\theta_{m}^{\varepsilon}(t),\qquad\theta_{m}^{\varepsilon}(0)=\theta_{0}.

Multiplying θmε(t)\theta^{\varepsilon}_{m}(t) and integration by parts in xx, one can easily get

(2.7) ddtθmεL22=2εθmεH˙α2,\frac{d}{dt}\lVert\theta_{m}^{\varepsilon}\rVert_{L^{2}}^{2}=-2\varepsilon\lVert\theta_{m}^{\varepsilon}\rVert_{\dot{H}^{\alpha}}^{2},

and we have the following lemma on the decay of θmεL2\|\theta_{m}^{\varepsilon}\|_{L^{2}}.

Lemma 2.1.

For any fixed m>0m>0, suppose that for all times t(a,b)t\in(a,b) we have θmε(t)H˙α2Nθmε(t)L22\lVert\theta_{m}^{\varepsilon}(t)\rVert_{\dot{H}^{\alpha}}^{2}\geqslant N\lVert\theta_{m}^{\varepsilon}(t)\rVert_{L^{2}}^{2}. Then the following decay estimate holds

(2.8) θmε(b)L22e2εN(ba)θmε(a)L22.\lVert\theta_{m}^{\varepsilon}(b)\rVert_{L^{2}}^{2}\leqslant e^{-2\varepsilon N(b-a)}\lVert\theta_{m}^{\varepsilon}(a)\rVert_{L^{2}}^{2}.

We also need an estimate on the growth of the difference between the viscous and inviscid problem in terms of L2L^{2} norm. First notice that for any smooth incompressible flow u(mx)u(mx), and define m\mathcal{L}_{m} as before, then for any ϕH˙α\phi\in\dot{H}^{\alpha}, α1\alpha\geqslant 1 and t>0t>0 the following estimates hold:

(2.9) mϕL2Cu(m)LϕH˙αandeimtϕH˙αFm(t)ϕH˙α\lVert\mathcal{L}_{m}\phi\rVert_{L^{2}}\leqslant C\|u(m\cdot)\|_{L^{\infty}}\lVert\phi\rVert_{\dot{H}^{\alpha}}\quad\text{and}\quad\lVert e^{i\mathcal{L}_{m}t}\phi\rVert_{\dot{H}^{\alpha}}\leqslant F_{m}(t)\lVert\phi\rVert_{\dot{H}^{\alpha}}

with both the constant CC and the function Fm(t)<F_{m}(t)<\infty independent of ϕ\phi and Fm(t)Lloc2(0,)F_{m}(t)\in L_{\text{loc}}^{2}(0,\infty). Here Fm(t)F_{m}(t) depends on u(m)Hα\|u(m\cdot)\|_{H^{\alpha}}.

Lemma 2.2.

Fix any m>0m>0, let θm0(t)\theta_{m}^{0}(t), θmε(t)\theta_{m}^{\varepsilon}(t) be solution of

(2.10) ddtθm0(t)=imθm0(t),ddtθmε(t)=(imε(Δ)α)θmε(t),\frac{d}{dt}\theta_{m}^{0}(t)=i\mathcal{L}_{m}\theta_{m}^{0}(t),\qquad\frac{d}{dt}\theta_{m}^{\varepsilon}(t)=(i\mathcal{L}_{m}-\varepsilon(-\Delta)^{\alpha})\theta_{m}^{\varepsilon}(t),

with θm0(0)=θmε(0)=θ0H˙α\theta_{m}^{0}(0)=\theta_{m}^{\varepsilon}(0)=\theta_{0}\in\dot{H}^{\alpha}. Then

(2.11) ddtθm0(t)θmε(t)L2212εθm0(t)H˙α212εFm2(t)θ0H˙α2.\frac{d}{dt}\lVert\theta_{m}^{0}(t)-\theta_{m}^{\varepsilon}(t)\rVert_{L^{2}}^{2}\leqslant\frac{1}{2}\varepsilon\lVert\theta_{m}^{0}(t)\rVert_{\dot{H}^{\alpha}}^{2}\leqslant\frac{1}{2}\varepsilon F_{m}^{2}(t)\lVert\theta_{0}\rVert_{\dot{H}^{\alpha}}^{2}.

And further,

(2.12) θm0(t)θmε(t)L2212εθ0H˙α20τFm2(t)𝑑t\lVert\theta_{m}^{0}(t)-\theta_{m}^{\varepsilon}(t)\rVert_{L^{2}}^{2}\leqslant\frac{1}{2}\varepsilon\lVert\theta_{0}\rVert_{\dot{H}^{\alpha}}^{2}\int_{0}^{\tau}F_{m}^{2}(t)dt

for any time t<τt<\tau.

The proofs of Lemma 2.1 and Lemma 2.2 are standard, the reader can find them on [1].

3. Discussion of the main theorem

Theorem 3.1.

Let uC(𝕋2)u\in C^{\infty}(\mathbb{T}^{2}) be a cellular flow given in (1.5). Fix γ>0\gamma>0, α1\alpha\geqslant 1. For any ε>0\varepsilon>0, there exists m0>0m_{0}>0, such that for any mm0m\geqslant m_{0}, there exists an A0>0A_{0}>0, such that τα(Au(m),γ)ε\tau_{\alpha}^{*}(Au(m\cdot),\gamma)\leqslant\varepsilon for any AA0A\geqslant A_{0}.

Remark 3.2.

For d=3d=3, one can also have analogous of (1.5), say, Example 5.10 in [8].

Remark 3.3.

The case α=γ=1\alpha=\gamma=1 has already been proved in [8] via a stochastic method. Therefore, we only remain to show that the smallness of τ1(Au(m),γ)\tau_{1}^{*}(Au(m\cdot),\gamma) implies the smallness of τα(Au(m),γ)\tau_{\alpha}^{*}(Au(m\cdot),\gamma) for any α>1\alpha>1.

Remark 3.4.

Through this chapter, α1\alpha\geqslant 1 is a fixed number. The constant on the estimates may always depend on the value of α\alpha.

To prove the main theorem, we first establish and prove some lemmas.

First, let us consider the eigenfunctions of m\mathcal{L}_{m}. We will show that the eigenfunctions of m\mathcal{L}_{m} will have big H˙α\dot{H}^{\alpha} norm. In fact, for the cellular flow in dimension two, the eigenfunction always exists and is of class HαH^{\alpha}. However, for the sake of simplicity we ignore the existence discussion here.

Let ϕm0\phi_{m}^{0} be any normalized (ϕm0L2=1\lVert\phi_{m}^{0}\rVert_{L^{2}}=1) eigenfunction of m\mathcal{L}_{m} that belongs to H˙α\dot{H}^{\alpha} (with α1\alpha\geqslant 1) associated with eigenvalue EmE_{m}. Then the following lemma shows that for fixed sufficiently large mm, the normalized point spectrums of m\mathcal{L}_{m} have an uniform lower bound in terms of the H˙α\dot{H}^{\alpha} norm, otherwise it contradicts with the result for α=γ=1\alpha=\gamma=1 proved in [8].

Lemma 3.5.

With the same assumption of u(m)u(m\cdot) in Theorem 3.1. Given τ0>0\tau_{0}>0, there exists m0(τ0)m_{0}(\tau_{0}) such that for any fixed m>m0(τ0)m>m_{0}(\tau_{0}) it holds ϕm0H˙12>12τ0\lVert\phi_{m}^{0}\rVert_{\dot{H}^{1}}^{2}>\frac{1}{2\tau_{0}}. If ϕm0\phi_{m}^{0} further belongs to HαH^{\alpha} for some α>1\alpha>1, then ϕm0H˙α2>12λ1α1τ0\lVert\phi_{m}^{0}\rVert_{\dot{H}^{\alpha}}^{2}>\frac{1}{2\lambda_{1}^{\alpha-1}\tau_{0}}.

Proof.

The proof is very similar to the proof of Theorem 1.4 (the easier direction) in [1]. We prove by showing contradiction, assume for arbitrary large m0m_{0} and m>m0m>m_{0}, there exists an eigenvalue EmE_{m} and corresponding eigenfunction ϕm0\phi_{m}^{0} (associate to operator m\mathcal{L}_{m}) such that τ0ϕm0H˙1212\tau_{0}\lVert\phi_{m}^{0}\rVert_{\dot{H}^{1}}^{2}\leqslant\frac{1}{2}. Consider the solution θmA(t)\theta^{A}_{m}(t) to (1.2), with convection term Au(mx)Au(mx)\cdot\nabla and initial data ϕm0L02H˙1\phi_{m}^{0}\in L^{2}_{0}\cap\dot{H}^{1}(with ϕm0L2=1\lVert\phi_{m}^{0}\rVert_{L^{2}}=1). Take the L2L^{2} inner product of (1.2) with ϕm0\phi_{m}^{0}, we get

(3.1) ddtθmA(t),ϕm0=iAEmθmA,ϕm0+ΔθmA,ϕm0.\frac{d}{dt}\langle\theta^{A}_{m}(t),\phi_{m}^{0}\rangle=iAE_{m}\langle\theta^{A}_{m},\phi_{m}^{0}\rangle+\langle\Delta\theta^{A}_{m},\phi_{m}^{0}\rangle.

This further yields

(3.2) |ddt(eiAEmtθmA,ϕm0)|12(θmA(t)H˙12+ϕm0H˙12).\Big{|}\frac{d}{dt}\left(e^{-iAE_{m}t}\langle\theta^{A}_{m},\phi_{m}^{0}\rangle\right)\Big{|}\leqslant\frac{1}{2}\left(\lVert\theta^{A}_{m}(t)\rVert_{\dot{H}^{1}}^{2}+\lVert\phi_{m}^{0}\rVert_{\dot{H}^{1}}^{2}\right).

Note that 0θmA(t)H˙12𝑑t12ϕm0L22=12\int_{0}^{\infty}\lVert\theta^{A}_{m}(t)\rVert_{\dot{H}^{1}}^{2}dt\leqslant\frac{1}{2}\lVert\phi_{m}^{0}\rVert_{L^{2}}^{2}=\frac{1}{2} and τ0ϕm0H˙1212\tau_{0}\|\phi_{m}^{0}\|_{\dot{H}^{1}}^{2}\leqslant\frac{1}{2}. Hence,

|(eiAEmτ0θmA,ϕm0)1|\displaystyle\left|\left(e^{-iAE_{m}\tau_{0}}\langle\theta^{A}_{m},\phi_{m}^{0}\rangle\right)-1\right| 0τ0|ddt(eiAEmtθmA,ϕm0)|𝑑t\displaystyle\leqslant\int_{0}^{\tau_{0}}\left|\frac{d}{dt}\left(e^{-iAE_{m}t}\langle\theta^{A}_{m},\phi_{m}^{0}\rangle\right)\right|dt
12(0τ0θmA(t)H˙12𝑑t+τ0ϕm0H˙12)12.\displaystyle\leqslant\frac{1}{2}\left(\int_{0}^{\tau_{0}}\lVert\theta^{A}_{m}(t)\rVert_{\dot{H}^{1}}^{2}dt+\tau_{0}\lVert\phi_{m}^{0}\rVert_{\dot{H}^{1}}^{2}\right)\leqslant\frac{1}{2}.

Therefore for 0<tτ00<t\leqslant\tau_{0}, we have |θmA(t),ϕm0|12\lvert\langle\theta^{A}_{m}(t),\phi_{m}^{0}\rangle\rvert\geqslant\frac{1}{2}, which further implies θmA(τ0)L212ϕm0L2\lVert\theta^{A}_{m}(\tau_{0})\rVert_{L^{2}}\geqslant\frac{1}{2}\lVert\phi_{m}^{0}\rVert_{L^{2}} uniformly in AA. Equivalently, for any m>0m>0 and A>0A>0, the dissipation time τ1(Au(mx))τ0\tau_{1}^{*}(Au(mx))\geqslant\tau_{0}, which contradicts to the Theorem 1.2 and Remark 1.3. This completes the proof of the first part of the lemma. The second part directly follows from the Poincare inequality. ∎

With the help of Lemma 3.5, we can further control from below the growth of H˙α\dot{H}^{\alpha} norm of solutions, to the underlying inviscid problem, corresponding to discrete eigenfunctions.

Lemma 3.6.

With the same assumption of u(mx)u(mx) in Theorem 3.1. Let KS={θL2:θL2=1}K\subset S=\{\theta\in L^{2}:\|\theta\|_{L^{2}}=1\} be a compact set in L2L^{2}. Consider the set K1=def{θK|PpθL212}K_{1}\stackrel{{\scriptstyle\scriptscriptstyle\textup{def}}}{{=}}\left\{\theta\in K|\lVert P_{p}\theta\rVert_{L^{2}}\geqslant\frac{1}{2}\right\}. Then for any B>0B>0, there exists m0(B)m_{0}(B) such that for any fixed m>m0m>m_{0} we can find Np(B,K,m)N_{p}(B,K,m) and Tp(B,K,m)T_{p}(B,K,m) such that for any NNp(B,K,m)N\geqslant N_{p}(B,K,m), TTp(B,K,m)T\geqslant T_{p}(B,K,m) and any θK1H˙α\theta\in K_{1}\cap\dot{H}^{\alpha},

(3.3) 1T0TPNeimtPpθH˙α2𝑑tB.\frac{1}{T}\int_{0}^{T}\lVert P_{N}e^{i\mathcal{L}_{m}t}P_{p}\theta\rVert_{\dot{H}^{\alpha}}^{2}dt\geqslant B.
Proof.

The proof is quite similar to the proof of Lemma 3.3 in [1]. Denote by EjmE_{j}^{m} the eigenvalues of m\mathcal{L}_{m}(distinct, without repetitions) and by QjmQ_{j}^{m} the orthogonal projection on the space spanned by the eigenfunctions corresponding to EjmE_{j}^{m}. Without lost of generality, we assume K1K_{1} is nonempty, otherwise there is nothing to prove. Observe that, by applying Lemma 3.5 with τ0=18Bλ1α1\tau_{0}=\frac{1}{8B\lambda_{1}^{\alpha-1}} there is m0(B)m_{0}(B) such that for any m>m0m>m_{0} we have

(3.4) jQjmθH˙α2>4B.\sum_{j}\lVert Q_{j}^{m}\theta\rVert_{\dot{H}^{\alpha}}^{2}>4B.

Therefore by compactness of KK, there exists Np=Np(B,K,m)N_{p}=N_{p}(B,K,m), uniform in jj, such that for N>Np(B,K,m)N>N_{p}(B,K,m)

jPNQjmθH˙α22B.\sum_{j}\lVert P_{N}Q_{j}^{m}\theta\rVert_{\dot{H}^{\alpha}}^{2}\geqslant 2B.

The remains proof is identical the same as the proof of Lemma 3.3 in [1], with 1\lVert\cdot\rVert_{1} (defined in [1]) replaced by H˙α\lVert\cdot\rVert_{\dot{H}^{\alpha}}. ∎

On the other hand, for the continuous spectrum of m\mathcal{L}_{m} we can control it by the following RAGE type theorem.

Lemma 3.7.

Let KSK\subset S be a compact set. For any N,σ,m>0N,\sigma,m>0, there exists Tc(N,σ,m,K)T_{c}(N,\sigma,m,K) such that for all TTc(N,σ,m,K)T\geqslant T_{c}(N,\sigma,m,K) and any θK\theta\in K,

(3.5) 1T0TPNeimtPcθL22𝑑tσ.\frac{1}{T}\int_{0}^{T}\lVert P_{N}e^{i\mathcal{L}_{m}t}P_{c}\theta\rVert_{L^{2}}^{2}dt\leqslant\sigma.

The proof of Lemma 3.7 can be found in [1].

Now, we turn to the main theorem. We may decomposite the proof into two parts:

(1). For the iniviscid problem, by a decomposition of the initial value θ0\theta_{0} into continuous spectrum and discrete spectrum of m\mathcal{L}_{m}, one can use Lemma 3.6 and 3.7 lower bound of high frequencies in the sense of time average.

(2). By Lemma 3.5 and rescaling of time, one can control the difference of the solutions to the inviscid problem and the viscid problem, and based on Lemma 2.1 to get the desired decay estimate.

This argument is similar to the proof of the Theorem 1.4 in [1]. For completion, we provide the whole proof in Appendix.

Remark 3.8.

The work in [9] can be generalized to hyper-diffusion in the same manner. However, due to applying the RAGE theorem in the proof of Theorem 3.1 (or Lemma 3.7), the quantitative estimates in [9] will be lost while translating to the hyper-diffusion scenario.

4. Applications

4.1. Applications to reaction hyper-diffusion equations

In [8], the authors provide two applications that rely on the smallness of τ1(u)\tau_{1}^{*}(u), one is the suppression of blow-up in the Keller-Segel system, the other one is the quenching in models of combustion. Here we collect several results that rely on the smallness of τ2(u)\tau_{2}^{*}(u). In this chapter, we will use f¯\bar{f} to denote the mean of function ff.

  1. (1)

    Advective Cahn-Hilliard equation (see [3])

    (4.1) ct+uc+γΔ2c=Δ(c3c).c_{t}+u\cdot\nabla c+\gamma\Delta^{2}c=\Delta(c^{3}-c).
    Theorem 4.1.

    Let d{2,3}d\in\{2,3\}, uL([0,);W1,(𝕋d))u\in L^{\infty}([0,\infty);W^{1,\infty}(\mathbb{T}^{d})), and cc be the strong solution of (4.1) with initial data c0H2(𝕋d)c_{0}\in H^{2}(\mathbb{T}^{d}).

    1. (a)

      When d=2d=2, for any β>1,μ>0\beta>1,\mu>0, there exists a time

      T0=T0(c0c¯L2,c¯,β,γ,μ)T_{0}=T_{0}(\lVert c_{0}-\bar{c}\rVert_{L^{2}},\bar{c},\beta,\gamma,\mu)

      such that if τ2(u,γ)<T0\tau_{2}^{*}(u,\gamma)<T_{0}, then for every t0t\geqslant 0, we have

      (4.2) c(t)c¯L2βeμtc0c¯L2.\lVert c(t)-\bar{c}\rVert_{L^{2}}\leqslant\beta e^{-\mu t}\lVert c_{0}-\bar{c}\rVert_{L^{2}}.
    2. (b)

      When d=3d=3, for any β>1,μ>0\beta>1,\mu>0, there exists a time

      T1=T1(c0c¯L2,c¯,β,γ,μ)T_{1}=T_{1}(\lVert c_{0}-\bar{c}\rVert_{L^{2}},\bar{c},\beta,\gamma,\mu)

      such that if

      (1+uL)1/2τ2(u,γ)<T1,(1+\lVert\nabla u\rVert_{L^{\infty}})^{1/2}\tau_{2}^{*}(u,\gamma)<T_{1},

      then (4.2) still holds for every t0t\geqslant 0.

  2. (2)

    Advective Kuramoto-Sivashinsky equation (see [7])

    (4.3) ϕt+uϕ+Δ2ϕ=12|ϕ|2Δϕ.\phi_{t}+u\cdot\nabla\phi+\Delta^{2}\phi=-\frac{1}{2}\lvert\nabla\phi\rvert^{2}-\Delta\phi.
    Theorem 4.2.

    Let d=2d=2, uL((0,);W1,(𝕋2))u\in L^{\infty}((0,\infty);W^{1,\infty}(\mathbb{T}^{2})) and ϕ\phi be the mild solution of (4.3) with initial data ϕ0L2(𝕋2)\phi_{0}\in L^{2}(\mathbb{T}^{2}) (see Definition 2.1 in [7] for the mild solution). Let ϕ0ϕ¯L2=B>0\lVert\phi_{0}-\bar{\phi}\rVert_{L^{2}}=B>0. There exists a time T2(B)T_{2}(B) such that if τ2(u,1)<T2(B)\tau_{2}^{*}(u,1)<T_{2}(B), then (4.3) admits a global mild solution.

  3. (3)

    Advective thin-film equation (see [5])

    (4.4) th+uh+Δ2h=(|h|p2h).\partial_{t}h+u\cdot\nabla h+\Delta^{2}h=-\nabla\cdot(\lvert\nabla h\rvert^{p-2}\nabla h).
    Theorem 4.3.

    For 2<p<32<p<3, uL([0,),W1,(𝕋d))u\in L^{\infty}\left([0,\infty),W^{1,\infty}(\mathbb{T}^{d})\right), and μ>0\mu>0. Let hh be the mild solution of (4.4) with initial data h(0)=h0L02h(0)=h_{0}\in L_{0}^{2}. There exists a threshold value

    T1=T1(h0L2,μ,p)T_{1}=T_{1}(\lVert h_{0}\rVert_{L^{2}},\mu,p)

    such that if

    (uL(τ2(u,1))54+(τ2(u,1))34)T1(h0L2,μ,p),\left(\lVert u\rVert_{L^{\infty}}\left(\tau_{2}^{*}(u,1)\right)^{\frac{5}{4}}+\left(\tau_{2}^{*}(u,1)\right)^{\frac{3}{4}}\right)\leqslant T_{1}(\lVert h_{0}\rVert_{L^{2}},\mu,p),

    then there exists a constant β>0\beta>0, such that for any t>0t>0 it holds

    h(t)L2βeμth0L2.\lVert h(t)\rVert_{L^{2}}\leqslant\beta e^{-\mu t}\lVert h_{0}\rVert_{L^{2}}.
Remark 4.4.

Theorem 3.1 can be used to show the existence of cellular flows satisfy the conditions for the two-dimensional Cahn-Hilliard equation (Theorem 4.1 part (a)) and the Advective Kuramoto-Sivashinsky equation (Theorem 4.2). However, both three dimensional Cahn-Hilliard equation (Theorem 4.1 part (b)) and the Advective thin-film equation (Theorem 4.3) further requires the smallness of uHβτ2(u,1)\lVert u\rVert_{H}^{\beta}\tau_{2}^{*}(u,1) for some 0<β<10<\beta<1 and HH to be some Sobolev space. The existence of time-independent flow that satisfies such conditions is not clear and remains to be further investigated.

4.2. Applications to nonlinear diffusion equations

In this section, we show how to generalize the dissipation enhancement results of cellular flows to nonlinear diffusion equations.

Firstly, we consider the so-called porous medium equations on 𝕋2\mathbb{T}^{2}:

(4.5) tθ+uθνΔ(θq)=0,θ(x,0)=θ0(x),\partial_{t}\theta+u\cdot\nabla\theta-\nu\Delta(\theta^{q})=0,\quad\theta(x,0)=\theta_{0}(x),

where q>1,ν>0q>1,\nu>0, and uu is a divergence-free vector field. In [11], the authors showed that for time-periodic flows, if its unitary evolution operator (see equations (1.3)-(1.5) in [11]) has no eigenfunctions in H1H^{1}, then the time-periodic flow is relaxation-enhancing to (4.5). Our conclusion is similar to that in [11], but we consider uu is a time independent cellular flow. Consider the same class of initial data as in [11], that is: 0<hθ0h10<h\leqslant\theta_{0}\leqslant h^{-1}. Then the classical theory (see, e.g., [14]) yields there exists a unique solution correspond to θ0\theta_{0} provided uC(𝕋2)u\in C^{\infty}(\mathbb{T}^{2}). Further more, the maximum principle guarantees hθh1h\leqslant\theta\leqslant h^{-1}. Also observe that the average θ¯=θ¯0\bar{\theta}=\bar{\theta}_{0} is still preserved by (4.5). We have following dissipation enhancement kind result for cellular flows.

Theorem 4.5.

Consider equation (4.5) with 0<hθ0h10<h\leqslant\theta_{0}\leqslant h^{-1} and θ0θ¯0L2=1\lVert\theta_{0}-\bar{\theta}_{0}\rVert_{L^{2}}=1. Then for any τ>0\tau>0, there exist cellular flows with proper cell size and amplitude (depend on τ\tau, hh, and ν\nu) such that

(4.6) θ(x,τ)θ¯L212.\lVert\theta(x,\tau)-\bar{\theta}\rVert_{L^{2}}\leqslant\frac{1}{2}.
Remark 4.6.

Note that to compare with the linear case (that is, Theorem 3.1), the main difference is that the cellular flows in Theorem 4.5 depend on the initial data θ0\theta_{0}, or more precisely on hh.

Proof.

Firstly, observe that for any 0<hψh10<h\leqslant\psi\leqslant h^{-1}, the expression 𝕋2ψq1|ψ|2𝑑x\int_{\mathbb{T}^{2}}\psi^{q-1}\lvert\nabla\psi\rvert^{2}dx is equivalent to the H1(𝕋2)H^{1}(\mathbb{T}^{2}) norm. Then the proof is almost the same as the linear case, except we need to establish a parallel estimate for Lemma 2.2. For any cellular flow with cell size mm: um=u(mx)u_{m}=u(mx), let θmν\theta_{m}^{\nu} denote the corresponding solution to (4.5) and θm0\theta_{m}^{0} denotes the solution to the inviscid problem as in Lemma 2.2. The authors in [11] showed that

ddtθmν(t)θm0(t)L222ν𝕋2Δ(θmν)q(θmνθm0)𝑑x\displaystyle\frac{d}{dt}\lVert\theta_{m}^{\nu}(t)-\theta_{m}^{0}(t)\rVert_{L^{2}}^{2}\leqslant 2\nu\int_{\mathbb{T}^{2}}\Delta(\theta_{m}^{\nu})^{q}(\theta_{m}^{\nu}-\theta_{m}^{0})dx
2νq(𝕋2(θmν)q1|θmν|2𝑑x)1/2(𝕋2(θmν)q1|θm0|2𝑑x)1/2\displaystyle\leqslant 2\nu q\left(\int_{\mathbb{T}^{2}}(\theta_{m}^{\nu})^{q-1}\lvert\nabla\theta_{m}^{\nu}\rvert^{2}dx\right)^{1/2}\left(\int_{\mathbb{T}^{2}}(\theta_{m}^{\nu})^{q-1}\lvert\nabla\theta_{m}^{0}\rvert^{2}dx\right)^{1/2}
2νq𝕋2(θmν)q1|θmν|2𝑑x\displaystyle\qquad-2\nu q\int_{\mathbb{T}^{2}}(\theta_{m}^{\nu})^{q-1}\lvert\nabla\theta_{m}^{\nu}\rvert^{2}dx
νq2𝕋2(θmν)q1|θm0|2𝑑x\displaystyle\leqslant\frac{\nu q}{2}\int_{\mathbb{T}^{2}}(\theta_{m}^{\nu})^{q-1}\lvert\nabla\theta_{m}^{0}\rvert^{2}dx
νqh1q2Fm2(t)θ0L22.\displaystyle\leqslant\frac{\nu qh^{1-q}}{2}F_{m}^{2}(t)\lVert\nabla\theta_{0}\rVert_{L^{2}}^{2}.

Here Fm(t)F_{m}(t) is as in Lemma 2.2. The remaining proof is almost the same as the linear case, only with some estimates may depend on hh, so we omit them. ∎

The second nonlinear diffusion equation we consider is the advective p-Laplacian equation:

(4.7) tϑ+uϑν(|ϑ|p2ϑ)=0,ϑ(x,0)=ϑ0(x),\partial_{t}\vartheta+u\cdot\nabla\vartheta-\nu\nabla\cdot\left(\lvert\nabla\vartheta\rvert^{p-2}\nabla\vartheta\right)=0,\quad\vartheta(x,0)=\vartheta_{0}(x),

where p>2p>2 and ν>0\nu>0. In the recent paper [4], the author studied dissipation enhancement effect of time-dependent mixing flows to the weak solutions of (4.7). For simplicity, consider ϑ0L02(𝕋2)\vartheta_{0}\in L_{0}^{2}(\mathbb{T}^{2}), then (4.7) possesses a unique weak solution ϑL02(𝕋2)\vartheta\in L_{0}^{2}(\mathbb{T}^{2}) (see Section 3 in [4]). With the same spirit as the porous medium equation, the result in [4] can be easily generalized to the cellular flow scenario. Let uu be a cellular flow, we have:

Theorem 4.7.

Consider equation (4.7) with ϑ0L02(𝕋2)\vartheta_{0}\in L_{0}^{2}(\mathbb{T}^{2}) and ϑ0L2=1\lVert\vartheta_{0}\rVert_{L^{2}}=1. Then for any τ>0\tau>0, there exist cellular flows with proper cell size and amplitude (depend on τ,ϑ0L2\tau,\lVert\vartheta_{0}\rVert_{L^{2}}, and ν\nu) such that

(4.8) ϑ(x,τ)L212.\lVert\vartheta(x,\tau)\rVert_{L^{2}}\leqslant\frac{1}{2}.
Proof.

Similar to the porous medium case, it is sufficient for us to observe the fact that

(4.9) ψLpCψL2\lVert\nabla\psi\rVert_{L^{p}}\geqslant C\lVert\nabla\psi\rVert_{L^{2}}

for any p2p\geqslant 2, according to the Hölder inequality. Then we just need to re-estimate the difference between (4.7) and the underlying inviscid problem. This estimate was established in Lemma 4.1 of [4]. For the convenience of readers, we cite the estimate here. For a cellular flow with cell size mm, let ϑmν\vartheta_{m}^{\nu} denote the solution to (4.7) and ϑm0\vartheta_{m}^{0} denote the inviscid problem. Then the estimate writes

(4.10) ϑmν(t)ϑm0(t)L22CpumLe2umLtφ0Lpp,\lVert\vartheta_{m}^{\nu}(t)-\vartheta_{m}^{0}(t)\rVert_{L^{2}}^{2}\leqslant\frac{C_{p}}{\lVert\nabla u_{m}\rVert_{L^{\infty}}}e^{2\lVert\nabla u_{m}\rVert_{L^{\infty}}t}\lVert\nabla\varphi_{0}\rVert_{L^{p}}^{p},

where CpC_{p} is a constant depending on the parameter pp. Combine the fact (4.9) and estimate (4.10), with the guide of [4] one can easily complete the proof of Theorem 4.7 by modifying the proof of the linear diffusion case. ∎

5. Relaxation enhancing flows for hyper-diffusion

In this chapter, we will construct a relaxation enhancing flow for hyper-diffusion but not for standard diffusion.

To begin with, we recall that a number α\alpha\in\mathbb{R} is called β\beta-Diophatine if there exists a constant CC such that for each k{0}k\in\mathbb{Z}\setminus\{0\} we have

infp|αk+p|C|k|1+β.\inf_{p\in\mathbb{Z}}|\alpha\cdot k+p|\geqslant\frac{C}{|k|^{1+\beta}}.

And the number α\alpha is Liouvillean if it is not Diophantine for any β>0\beta>0. More specifically, for any n(>1)n(>1) and constant C>0C>0, one can find qn{0}q_{n}\in\mathbb{Z}\setminus\{0\} and pnp_{n}\in\mathbb{Z}, such that

(5.1) |αqn+pn|C|qn|n.|\alpha\cdot q_{n}+p_{n}|\leqslant\frac{C}{|q_{n}|^{n}}.

One can understand the Liouvillean numbers are the ones which can be very well approximated by rationals.

Remark 5.1.

There exist lots of such irrational Liouvillean numbers. For example, a well-known one is given by α=n0110n!\alpha=\sum\limits_{n\geqslant 0}\frac{1}{10^{n!}}.

On the other hand, we denote by Φtu\Phi^{u}_{t} the flow on the torus generated by uu, and by UtU^{t} the evolution operator on L2(𝕋2)L^{2}(\mathbb{T}^{2}) generated by Φtu\Phi_{t}^{u}:(Utf)(x)=f(Φtu(x))(U^{t}f)(x)=f(\Phi_{-t}^{u}(x)). By now, we are ready to state our result.

Proposition 5.2.

There exists a smooth incompressible flow u(x,y)u(x,y) on the two-dimensional torus so that the corresponding unitary evolution UtU^{t} has a discrete spectrum on H1(𝕋2)H^{1}(\mathbb{T}^{2}) but none of the eigenfunctions of UtU^{t} are in H2(𝕋2)H^{2}(\mathbb{T}^{2}).

To prove Proposition 5.2, we first prove following auxiliary lemma. Let 𝕊1=[12,12]\mathbb{S}^{1}=[-\frac{1}{2},\frac{1}{2}] be the one-dimensional circle.

Lemma 5.3.

Consider the irrational Liouvillean number α\alpha. There exists a C(𝕊1)C^{\infty}(\mathbb{S}^{1}) mean-zero function Q(ξ)Q(\xi) so that the homology equation

(5.2) R(ξ+α)R(ξ)=Q(ξ)R(\xi+\alpha)-R(\xi)=Q(\xi)

has a measurable solution R(ξ):𝕊1R(\xi):\mathbb{S}^{1}\rightarrow\mathbb{R} such that R(x)R(x) is in H1(𝕊1)H^{1}(\mathbb{S}^{1}) but not in H2(𝕊1)H^{2}(\mathbb{S}^{1}). In addition, the function Rλ(ξ)=eiλR(ξ)R_{\lambda}(\xi)=e^{i\lambda R(\xi)} is in H1(𝕊1)H^{1}(\mathbb{S}^{1}) but not in H2(𝕊1)H^{2}(\mathbb{S}^{1}), for any λ{0}\lambda\in\mathbb{R}\setminus\{0\}.

Proof.

As the fist step, we construct a function R~\tilde{R} in L2(𝕊1)L4(𝕊1)L^{2}(\mathbb{S}^{1})\cap L^{4}(\mathbb{S}^{1}) but not in H1(𝕊1)H^{1}(\mathbb{S}^{1}) satisfying (5.2) with a mean zero function Q~(ξ)C(𝕊1)\tilde{Q}(\xi)\in C^{\infty}(\mathbb{S}^{1}). Let θ(t)\theta(t) be a CC^{\infty} bump function in 𝕊1\mathbb{S}^{1}, with θ(t)\theta(t) equals to 11 when 116t116-\frac{1}{16}\leqslant t\leqslant\frac{1}{16}, and equals to 0 when 18t12\frac{1}{8}\leqslant t\leqslant\frac{1}{2} or 12t18-\frac{1}{2}\leqslant t\leqslant-\frac{1}{8}, and smooth everywhere with |dθdt|20|\frac{d\theta}{dt}|\leqslant 20.

Then for q1q\geqslant 1 we define Q~q(x)=θ(q6x)\tilde{Q}_{q}(x)=\theta(q^{6}x), for x[18q6,18q6]x\in[-\frac{1}{8q^{6}},\frac{1}{8q^{6}}], and 0 elsewhere in [12,12][-\frac{1}{2},\frac{1}{2}]. Note that Q~q\tilde{Q}_{q} is still a CC^{\infty} function on 𝕊1\mathbb{S}^{1}. We define

(5.3) Q~=k(Q~qk2(qkα)Q~qk2),\tilde{Q}=\sum_{k}(\tilde{Q}_{q^{2}_{k}}(\cdot-q_{k}\alpha)-\tilde{Q}_{q^{2}_{k}}),

where the sequence {qk}\{q_{k}\} is chosen as follows. By the definition of Liouvillean number α\alpha in (5.1), fix any C>0C>0, one can choose integers qkq_{k} and pkp_{k} so that |αqkpk|C|qk|k|\alpha\cdot q_{k}-p_{k}|\leqslant\frac{C}{|q_{k}|^{k}} for any k>1k>1. Also note that since α\alpha is irrational, qkq_{k} has infinitely many choices for each kk. Therefore, without loss of generality, we can choose qkkq_{k}\geqslant k and qkq_{k} increasing in kk.
Now, we show Q~\tilde{Q} belongs to CC^{\infty} by checking the right-hand side of (5.3) converges very fast. By chain rule, we have

Q~qHrC(r)q6r.||\tilde{Q}_{q}||_{H^{r}}\leqslant C(r)q^{6r}.

Then by using Fourier expansion, Parseval’s identity and the definition of α\alpha, we get

||Q~qk2(qkα)Q~qk2||Hr\displaystyle||\tilde{Q}_{q^{2}_{k}}(\cdot-q_{k}\alpha)-\tilde{Q}_{q^{2}_{k}}||_{H^{r}} C|n|rQ~^qk2(n)(ei2πnqkα1)l2\displaystyle\leqslant C\||n|^{r}\hat{\tilde{Q}}_{q^{2}_{k}}(n)\left(e^{i2\pi n\cdot q_{k}\alpha}-1\right)\|_{l^{2}}
C|n|rQ~^qk2(n)(ei2πn(qkαpk)1)l2\displaystyle\leqslant C\||n|^{r}\hat{\tilde{Q}}_{q^{2}_{k}}(n)\left(e^{i2\pi n\cdot(q_{k}\alpha-p_{k})}-1\right)\|_{l^{2}}
C(r)|n|rQ~^qk2(n)l2/|qk|k\displaystyle\leqslant C(r)\||n|^{r}\hat{\tilde{Q}}_{q^{2}_{k}}(n)\|_{l^{2}}/|q_{k}|^{k}
C(r)Q~qk2Hr|qk|k.\displaystyle\leqslant C(r)\frac{||\tilde{Q}_{q^{2}_{k}}||_{H^{r}}}{|q_{k}|^{k}}.

By taking the HrH^{r} norm on the both sides of (5.3), and the series on the right-hand side is bounded by Ckqk10Ck1k10<C\cdot\sum\limits_{k}q_{k}^{-10}\leqslant C\cdot\sum\limits_{k}\frac{1}{k^{10}}<\infty for any k12r+10k\geqslant 12r+10. Therefore, Q~\tilde{Q} belongs to HrH^{r} for any rr, which further yields Q~\tilde{Q} is smooth by the Sobolev embedding arguments.
Before defining the function R~\tilde{R}, we define R~q=l=0q1Q~q2(lα)\tilde{R}_{q}=-\sum\limits_{l=0}^{q-1}\tilde{Q}_{q^{2}}(\cdot-l\alpha), then immediately we have

Q~q2(qα)Q~q2()=R~q(α)R~q().\tilde{Q}_{q^{2}}(\cdot-q\alpha)-\tilde{Q}_{q^{2}}(\cdot)=\tilde{R}_{q}(\cdot-\alpha)-\tilde{R}_{q}(\cdot).

Now we define R~=kR~qk\tilde{R}=\sum_{k}\tilde{R}_{q_{k}}, which satisfies R~(α)R~=Q~\tilde{R}(\cdot-\alpha)-\tilde{R}=\tilde{Q}. Let mm be the standard Lebesgue measure. Since we have m(supp(R~q))qm(supp(Q~q2))14q11m(supp(\tilde{R}_{q}))\leqslant q\cdot m(supp(\tilde{Q}_{q^{2}}))\leqslant\frac{1}{4}q^{-11}, so we have qm(supp(R~q))<12\sum_{q}m(supp(\tilde{R}_{q}))<\frac{1}{2}. Which means for any sequence {aq}\{a_{q}\}, qaqR~q\sum_{q}a_{q}\tilde{R}_{q} converges in measure, so is a measurable function in 𝕊1\mathbb{S}^{1}. Moreover, since we have Q~q2L4Cq3||\tilde{Q}_{q^{2}}||_{L^{4}}\leqslant Cq^{-3}, R~qL4qQ~q2L4Cq2||\tilde{R}_{q}||_{L^{4}}\leqslant q||\tilde{Q}_{q^{2}}||_{L^{4}}\leqslant Cq^{-2}, so R~\tilde{R} is in L4L^{4}, hence also in L2L^{2} and is finite almost everywhere. However, let t[116,116]t\in[-\frac{1}{16},\frac{1}{16}], then for any positive integer k0k_{0}, R~qk(t1qk012)Q~qk2(t1qk012)=θ(qk12qk012t)=1\tilde{R}_{q_{k}}(t\cdot\frac{1}{q^{12}_{k_{0}}})\leqslant-\tilde{Q}_{q^{2}_{k}}(t\cdot\frac{1}{q^{12}_{k_{0}}})=-\theta(\frac{q_{k}^{12}}{q_{k_{0}}^{12}}t)=-1, for any kk0k\leqslant k_{0}. This means R~(t1qk012)k0\tilde{R}(t\cdot\frac{1}{q^{12}_{k_{0}}})\leqslant-k_{0}, for any large positive integer k0k_{0} and any t[116,116]t\in[-\frac{1}{16},\frac{1}{16}]. This means for any k0>1k_{0}>1, there is a positive-measure set 𝔐k0𝕊1\mathfrak{M}_{k_{0}}\subset\mathbb{S}^{1} such that R~k0\tilde{R}\leqslant-k_{0} on it. Hence, there cannot be a continuous function R¯\bar{R} on 𝕊1\mathbb{S}^{1} such that R~(x)=R¯(x)\tilde{R}(x)=\bar{R}(x) almost everywhere, otherwise R¯(x)\bar{R}(x) must be unbounded in 𝕊1\mathbb{S}^{1}, hence by Sobolev embedding R~\tilde{R} is not in H1H^{1}. As a consequence, we construct such function R~\tilde{R} in L2L4L^{2}\cap L^{4} but not in H1H^{1}.

Now, without loss of generality, by subtracting constants and swap the sign, we can assume R~\tilde{R} we get is mean zero, and together with Q~\tilde{Q} satisfying (5.2), with the same regularity. Let R(x)=0xR~(y)𝑑yR(x)=\int_{0}^{x}\tilde{R}(y)dy, Q(x)=0xQ~(y)𝑑y+0αR~(y)𝑑yQ(x)=\int_{0}^{x}\tilde{Q}(y)dy+\int_{0}^{\alpha}\tilde{R}(y)dy, as R~(y+α)R~(y)=Q~(y)\tilde{R}(y+\alpha)-\tilde{R}(y)=\tilde{Q}(y), taking integration from 0 to xx, we have R(x+α)R(x)=Q(x)R(x+\alpha)-R(x)=Q(x). It is clear that QQ is in C(𝕊1)C^{\infty}(\mathbb{S}^{1}). By Fubini theorem, one can easily see that if the nn-th Fourier coefficient of R~\tilde{R} is ana_{n}, then the nn-th Fourier coefficient of RR is ann\frac{a_{n}}{n} for n0n\neq 0. Hence, R(x)R(x) is not in H2(𝕊1)H^{2}(\mathbb{S}^{1}). Meanwhile, by Lebesgue differentiation theorem, R(x)R(x) is in H1(𝕊1)W1,4(𝕊1)H^{1}(\mathbb{S}^{1})\cap W^{1,4}(\mathbb{S}^{1}). By chain rule and the fact that RR is in W1,4(𝕊)H1(𝕊)W^{1,4}(\mathbb{S})\cap H^{1}({\mathbb{S}}), we also have Rλ(x)=eiλR(x)R_{\lambda}(x)=e^{i\lambda R(x)} is in H1H^{1} but not in H2H^{2}, for λ0\lambda\neq 0. ∎

The remaining proof of Proposition 5.2 is identically the same as the proof of Proposition 6.2 in [1] (but replace Proposition 6.3 in [1] by Lemma 5.3 above). For completeness, we summarize the idea of the proof in the following two propostions.

Proposition 5.4.

There exists a function F(x,y)F(x,y) on 𝕋2\mathbb{T}^{2}, such that the unitary evolution to the flow u~(x,y)=(αF(x,y),1F(x,y))\tilde{u}(x,y)=(\frac{\alpha}{F(x,y)},\frac{1}{F(x,y)}) has a discrete spectrum on H1(𝕋2)H^{1}(\mathbb{T}^{2}) but none of the eigenfunctions of such evolution are in H2(𝕋2)H^{2}(\mathbb{T}^{2}).

Remark 5.5.

The construction of FF is explicit. For example, as in [1], by adding a constant and rescaling, one can replace the right hand side of (5.2) by Q1Q-1 instead of QQ, such that QQ is positive and has mean 11, and

F(x,y)=b+ψ(y)(Q(xαy)b),F(x,y)=b+\psi(y)(Q(x-\alpha y)-b),

where b=12minQb=\frac{1}{2}\min Q, ψ(y)\psi(y) is a smooth cutoff function supported near 0. And we also emphasize that u~\tilde{u} may not be incompressible.

Proposition 5.6.

For F(x,y)F(x,y) and the corresponding u~(x,y)\tilde{u}(x,y) in Proposition 5.4, there exists an incompressible flow u(x,y)u(x,y) and a measure-preserving map SS from 𝕋2\mathbb{T}^{2} to 𝕋2\mathbb{T}^{2}, such that the evolution of uu is conjugate by SS to the evolution of u~\tilde{u}, and the evolution of uu possesses the same spectrum as that of u~\tilde{u}.

From Proposition 5.6 one can directly get Proposition 5.2.

Acknowledgments

The work of Y.F. is supported by the National Key R&D Program of China, Project Number 2021YFA1001200. The work of X.X. is supported by the National Key R&D Program of China, Project Number 2021YFA1001200, and the NSFC Youth program, grant number 12101278. And the authors would like to thank Andrej Zlatoš, Gautam Iyer, and Jean-Luc Thiffeault for helpful suggestions.

Appendix A Proof of Theorem 3.1

Similar to (2.6), we rescale the time variable of our equation by AA and call ε=1A\varepsilon=\frac{1}{A}. Then it is sufficient for us to show that given τ>0\tau>0, there exists m0m_{0} and corresponding ε0=ε0(m0)>0\varepsilon_{0}=\varepsilon_{0}(m_{0})>0 such that if m>m0m>m_{0} and ε<ε0\varepsilon<\varepsilon_{0}, we have θmε(τ/ε)L2<12\lVert\theta_{m}^{\varepsilon}(\tau/\varepsilon)\rVert_{L^{2}}<\frac{1}{2} for any initial datum θ0\theta_{0} with θ0L2=1\lVert\theta_{0}\rVert_{L^{2}}=1. Notice that (Δ)α(-\Delta)^{\alpha} is an unbounded positive operator with a discrete spectrum. Its eigenvalues Λn=λnα\Lambda_{n}=\lambda_{n}^{\alpha}\rightarrow\infty as nn\rightarrow\infty, here λn\lambda_{n} stands for the eigenvalue of Δ-\Delta. Thus we can choose NN large enough, so that eΛNτ/80<12e^{-\Lambda_{N}\tau/80}<\frac{1}{2}. Define the sets K={θS|θH˙α2ΛN}SK=\left\{\theta\in S\big{|}\lVert\theta\rVert_{\dot{H}^{\alpha}}^{2}\leqslant\Lambda_{N}\right\}\subset S and K1={θK|PpθL221/2}K_{1}=\left\{\theta\in K\big{|}\lVert P_{p}\theta\rVert_{L^{2}}^{2}\geqslant 1/2\right\}, then KK is compact. According to Lemma 3.5, there exists m0m_{0} large enough so that for any fixed m>m0m>m_{0}, we can further choose JNJ\geqslant N and JNp(5ΛN,K,m)J\geqslant N_{p}(5\Lambda_{N},K,m). Define

(A.1) τ1=max{Tp(5ΛN,K,m),Tc(J,ΛN20ΛJ,m,K)},\tau_{1}=\max\left\{T_{p}(5\Lambda_{N},K,m),T_{c}(J,\frac{\Lambda_{N}}{20\Lambda_{J}},m,K)\right\},

where TpT_{p} is from Lemma 3.6, and TcT_{c} from Lemma 3.7. Notice that τ1\tau_{1} only depends on the value of mm and τ\tau. Finally, choose ε0>0\varepsilon_{0}>0 small enough (may depend on mm, τ\tau) so that τ1<τ/2ε0\tau_{1}<\tau/2\varepsilon_{0}, and

(A.2) ε00τ1Fm2(t)𝑑t120ΛJ,\varepsilon_{0}\int_{0}^{\tau_{1}}F_{m}^{2}(t)dt\leqslant\frac{1}{20\Lambda_{J}},

where Fm(t)F_{m}(t) is the function from condition (2.9). Take any ε<ε0\varepsilon<\varepsilon_{0}. If θmε(s)H˙α2ΛNθmε(s)L22\lVert\theta_{m}^{\varepsilon}(s)\rVert_{\dot{H}^{\alpha}}^{2}\geqslant\Lambda_{N}\lVert\theta_{m}^{\varepsilon}(s)\rVert_{L^{2}}^{2} holds for all s[0,τ/ε]s\in[0,\tau/\varepsilon], then according to Lemma 2.1 and the choice of ΛN\Lambda_{N} we directly get θmε(τ/ε)L2e2ΛNτ12\lVert\theta_{m}^{\varepsilon}(\tau/\varepsilon)\rVert_{L^{2}}\leqslant e^{-2\Lambda_{N}\tau}\leqslant\frac{1}{2}, and we are done. Otherwise, let τ0\tau_{0} be the first time in [0,τ/ε][0,\tau/\varepsilon] such that θmε(τ0)H˙α2ΛNθmε(τ0)L22\lVert\theta_{m}^{\varepsilon}(\tau_{0})\rVert_{\dot{H}^{\alpha}}^{2}\leqslant\Lambda_{N}\lVert\theta_{m}^{\varepsilon}(\tau_{0})\rVert_{L^{2}}^{2}. Then we claim that the following decay holds on the time interval [τ0,τ0+τ1][\tau_{0},\tau_{0}+\tau_{1}]:

(A.3) θmε(τ0+τ1)L22eΛNετ1/20θmε(τ0)L22.\lVert\theta_{m}^{\varepsilon}(\tau_{0}+\tau_{1})\rVert_{L^{2}}^{2}\leqslant e^{-\Lambda_{N}\varepsilon\tau_{1}/20}\lVert\theta_{m}^{\varepsilon}(\tau_{0})\rVert_{L^{2}}^{2}.

For simplicity, denote θmε(τ0)=θ0\theta_{m}^{\varepsilon}(\tau_{0})=\theta_{0}. Consider the two equations in (2.10), then by Lemma 2.2 and the choice of ε0\varepsilon_{0}, we have

(A.4) θmε(t)θm0(t)L22ΛN40ΛJθ0L22\lVert\theta_{m}^{\varepsilon}(t)-\theta_{m}^{0}(t)\rVert_{L^{2}}^{2}\leqslant\frac{\Lambda_{N}}{40\Lambda_{J}}\lVert\theta_{0}\rVert_{L^{2}}^{2}

for all t[τ0,τ0+τ1]t\in[\tau_{0},\tau_{0}+\tau_{1}]. Split θm0(t)=θm,c(t)+θm,p(t)\theta_{m}^{0}(t)=\theta_{m,c}(t)+\theta_{m,p}(t), observe that θm,c\theta_{m,c} and θm,p\theta_{m,p} also solve the inviscid problem ddtθ(t)=imθ\frac{d}{dt}\theta(t)=i\mathcal{L}_{m}\theta, but with initial data Pcθ0P_{c}\theta_{0} and Ppθ0P_{p}\theta_{0} at t=τ0t=\tau_{0}, respectively. Now, we consider two cases:
Case I. Assume that Pcθ0L2234θ02\lVert P_{c}\theta_{0}\rVert_{L^{2}}^{2}\geqslant\frac{3}{4}\lVert\theta_{0}\rVert^{2}, equivalently, Ppθ0L2214θ02\lVert P_{p}\theta_{0}\rVert_{L^{2}}^{2}\leqslant\frac{1}{4}\lVert\theta_{0}\rVert^{2}. Note that since θ0/θL2K\theta_{0}/\lVert\theta\rVert_{L^{2}}\in K, thus we can apply Lemma 3.7, by our choice of τ1\tau_{1} in (A.1), we get

(A.5) 1τ1τ0τ0+τ1PNθm,c(t)L22𝑑tΛN20ΛJθ0L22.\frac{1}{\tau_{1}}\int_{\tau_{0}}^{\tau_{0}+\tau_{1}}\lVert P_{N}\theta_{m,c}(t)\rVert_{L^{2}}^{2}dt\leqslant\frac{\Lambda_{N}}{20\Lambda_{J}}\lVert\theta_{0}\rVert_{L^{2}}^{2}.

By elementary inequalities, we further have

(A.6) (IPN)θm0(t)L22\displaystyle\lVert(I-P_{N})\theta_{m}^{0}(t)\rVert_{L^{2}}^{2} 12(IPN)θm,c(t)L22(IPN)θm,p(t)L22\displaystyle\geqslant\frac{1}{2}\lVert(I-P_{N})\theta_{m,c}(t)\rVert_{L^{2}}^{2}-\lVert(I-P_{N})\theta_{m,p}(t)\rVert_{L^{2}}^{2}
(A.7) 12θm,c(t)L2212PNθm,c(t)L22θm,p(t)L22.\displaystyle\geqslant\frac{1}{2}\lVert\theta_{m,c}(t)\rVert_{L^{2}}^{2}-\frac{1}{2}\lVert P_{N}\theta_{m,c}(t)\rVert_{L^{2}}^{2}-\lVert\theta_{m,p}(t)\rVert_{L^{2}}^{2}.

Combine the facts of the free evolution eimte^{i\mathcal{L}_{m}t} is unitary, ΛJΛN\Lambda_{J}\geqslant\Lambda_{N}, assumptions on Pcθ0L2\lVert P_{c}\theta_{0}\rVert_{L^{2}} and Ppθ0L2\lVert P_{p}\theta_{0}\rVert_{L^{2}}, and the estimate (A.5), we have

(A.8) 1τ1τ0τ0+τ1(IPN)θm0(t)L22𝑑t110θ0L22.\frac{1}{\tau_{1}}\int_{\tau_{0}}^{\tau_{0}+\tau_{1}}\lVert(I-P_{N})\theta_{m}^{0}(t)\rVert_{L^{2}}^{2}dt\geqslant\frac{1}{10}\lVert\theta_{0}\rVert_{L^{2}}^{2}.

By using the estimate (A.4), we get

(A.9) 1τ1τ0τ0+τ1(IPN)θmε(t)L22𝑑t140θ0L22.\frac{1}{\tau_{1}}\int_{\tau_{0}}^{\tau_{0}+\tau_{1}}\lVert(I-P_{N})\theta_{m}^{\varepsilon}(t)\rVert_{L^{2}}^{2}dt\geqslant\frac{1}{40}\lVert\theta_{0}\rVert_{L^{2}}^{2}.

The above estimate further implies

(A.10) τ0τ0+τ1θmε(t)H˙α2𝑑tΛJτ140θ0L22.\int_{\tau_{0}}^{\tau_{0}+\tau_{1}}\lVert\theta_{m}^{\varepsilon}(t)\rVert_{\dot{H}^{\alpha}}^{2}dt\geqslant\frac{\Lambda_{J}\tau_{1}}{40}\lVert\theta_{0}\rVert_{L^{2}}^{2}.

Combine the above inequality with identity (2.7) to get

(A.11) θmε(τ0+τ1)L22(1ΛJετ120)θmε(τ0)L22eΛJετ1/20θmε(τ0)L22.\lVert\theta_{m}^{\varepsilon}(\tau_{0}+\tau_{1})\rVert_{L^{2}}^{2}\leqslant\left(1-\frac{\Lambda_{J}\varepsilon\tau_{1}}{20}\right)\lVert\theta_{m}^{\varepsilon}(\tau_{0})\rVert_{L^{2}}^{2}\leqslant e^{-\Lambda_{J}\varepsilon\tau_{1}/20}\lVert\theta_{m}^{\varepsilon}(\tau_{0})\rVert_{L^{2}}^{2}.

This completes the proof of (A.3) in the first case, where we used the fact that ΛJΛN\Lambda_{J}\geqslant\Lambda_{N}.
Case II. Now we assume Ppθ0L2214θ0L22\lVert P_{p}\theta_{0}\rVert_{L^{2}}^{2}\geqslant\frac{1}{4}\lVert\theta_{0}\rVert_{L^{2}}^{2}. In this case θ0/θ0L2K1\theta_{0}/\lVert\theta_{0}\rVert_{L^{2}}\in K_{1}, so that we can apply Lemma 3.6. Thus, by the choice of JJ and τ1\tau_{1},

(A.12) 1τ1τ0τ0+τ1PNθm,p(t)H˙α2𝑑t5ΛNθ0L22.\frac{1}{\tau_{1}}\int_{\tau_{0}}^{\tau_{0}+\tau_{1}}\lVert P_{N}\theta_{m,p}(t)\rVert_{\dot{H}^{\alpha}}^{2}dt\geqslant 5\Lambda_{N}\lVert\theta_{0}\rVert_{L^{2}}^{2}.

Notice that (A.5) still holds because of the choice of τ0\tau_{0} and τ1\tau_{1}, it yields

(A.13) 1τ1τ0τ0+τ1PNθm,c(t)H˙α2𝑑tΛN20θ0L22.\frac{1}{\tau_{1}}\int_{\tau_{0}}^{\tau_{0}+\tau_{1}}\lVert P_{N}\theta_{m,c}(t)\rVert_{\dot{H}^{\alpha}}^{2}dt\leqslant\frac{\Lambda_{N}}{20}\lVert\theta_{0}\rVert_{L^{2}}^{2}.

Combine the estimates (A.12) and (A.13) we get

(A.14) 1τ1τ0τ0+τ1PNθm0(t)H˙α2𝑑t2ΛNθ0L22.\frac{1}{\tau_{1}}\int_{\tau_{0}}^{\tau_{0}+\tau_{1}}\lVert P_{N}\theta_{m}^{0}(t)\rVert_{\dot{H}^{\alpha}}^{2}dt\geqslant 2\Lambda_{N}\lVert\theta_{0}\rVert_{L^{2}}^{2}.

Then (A.4) and (A.14) give

(A.15) τ0τ0+τ1PNθmε(t)H˙α2𝑑tΛNτ12θ0L22\int_{\tau_{0}}^{\tau_{0}+\tau_{1}}\lVert P_{N}\theta_{m}^{\varepsilon}(t)\rVert_{\dot{H}^{\alpha}}^{2}dt\geqslant\frac{\Lambda_{N}\tau_{1}}{2}\lVert\theta_{0}\rVert_{L^{2}}^{2}

where we also used the fact that PNθmεPNθm0H˙α2ΛJθmεθm0L22\lVert P_{N}\theta_{m}^{\varepsilon}-P_{N}\theta_{m}^{0}\rVert_{\dot{H}^{\alpha}}^{2}\leqslant\Lambda_{J}\lVert\theta_{m}^{\varepsilon}-\theta_{m}^{0}\rVert_{L^{2}}^{2}. Then similar to the previous case, (A.15) implies

(A.16) θmε(τ0+τ1)L22eΛNετ1θmε(τ0)L22,\lVert\theta_{m}^{\varepsilon}(\tau_{0}+\tau_{1})\rVert_{L^{2}}^{2}\leqslant e^{-\Lambda_{N}\varepsilon\tau_{1}}\lVert\theta_{m}^{\varepsilon}(\tau_{0})\rVert_{L^{2}}^{2},

which completes the proof of the claim (A.3) in the second case.

Finally, we summarize all the cases. Firstly, we see that if θmε(τ0)H˙α2ΛNθmε(τ0)L22\lVert\theta_{m}^{\varepsilon}(\tau_{0})\rVert_{\dot{H}^{\alpha}}^{2}\leqslant\Lambda_{N}\lVert\theta_{m}^{\varepsilon}(\tau_{0})\rVert_{L^{2}}^{2}, then the decay (A.3) always holds. On the other hand, for any time interval I=[a,b]I=[a,b] such that θmε(t)H˙α2ΛNθmε(t)L22\lVert\theta_{m}^{\varepsilon}(t)\rVert_{\dot{H}^{\alpha}}^{2}\geqslant\Lambda_{N}\lVert\theta_{m}^{\varepsilon}(t)\rVert_{L^{2}}^{2} on II, then Lemma 2.1 yields

(A.17) θmε(b)L22e2ΛNε(ba)θmε(a)L22.\lVert\theta_{m}^{\varepsilon}(b)\rVert_{L^{2}}^{2}\leqslant e^{-2\Lambda_{N}\varepsilon(b-a)}\lVert\theta_{m}^{\varepsilon}(a)\rVert_{L^{2}}^{2}.

Combine the two decay factors obtained from (A.3) and the inequality above, also use the the fact τ1<τ/2ε\tau_{1}<\tau/2\varepsilon, then we can further find τ2[τ/2ε,τ/ε]\tau_{2}\in[\tau/2\varepsilon,\tau/\varepsilon] such that

(A.18) θmε(τ2)L22eΛNετ2/20eΛNτ/40<1/4\lVert\theta_{m}^{\varepsilon}(\tau_{2})\rVert_{L^{2}}^{2}\leqslant e^{-\Lambda_{N}\varepsilon\tau_{2}/20}\leqslant e^{-\Lambda_{N}\tau/40}<1/4

by the choice of ΛN\Lambda_{N}. Then by the monotonicity of the L2L^{2} norm (2.7) we have θmε(τ/ε)L2θmε(τ2)L2<1/2\lVert\theta_{m}^{\varepsilon}(\tau/\varepsilon)\rVert_{L^{2}}\leqslant\lVert\theta_{m}^{\varepsilon}(\tau_{2})\rVert_{L^{2}}<1/2, this completes the proof.

References

  • [1] P. Constantin, A. Kiselev, L. Ryzhik, and A. Zlatoš. Diffusion and mixing in fluid flow. Ann. of Math. (2), 168(2):643–674, 2008.
  • [2] Albert Fannjiang, Alexander Kiselev, and Lenya Ryzhik. Quenching of reaction by cellular flows. Geometric & Functional Analysis GAFA, 16(1):40–69, 2006.
  • [3] Yu Feng, Yuanyuan Feng, Gautam Iyer, and Jean-Luc Thiffeault. Phase separation in the advective cahn–hilliard equation. Journal of NonLinear Science, pages 1–25, 2020.
  • [4] Yu Feng, Bingyang Hu, and Xiaoqian Xu. Dissipation enhancement for a degenerated parabolic equation. arXiv preprint arXiv:2104.12578, 2022.
  • [5] Yu Feng, Bingyang Hu, and Xiaoqian Xu. Suppression of epitaxial thin film growth by mixing. Journal of Differential Equations, 317:561–602, 2022.
  • [6] Yuanyuan Feng and Gautam Iyer. Dissipation enhancement by mixing. Nonlinearity, 32(5):1810–1851, 2019.
  • [7] Yuanyuan Feng and Anna L Mazzucato. Global existence for the two-dimensional kuramoto-sivashinsky equation with advection. Communications in Partial Differential Equations, 47(2):279–306, 2022.
  • [8] Gautam Iyer, Xiaoqian Xu, and Andrej Zlatoš. Convection-induced singularity suppression in the keller-segel and other non-linear pdes. Transactions of the American Mathematical Society, 2021.
  • [9] Gautam Iyer and Hongyi Zhou. Quantifying the dissipation enhancement of cellular flows. arXiv preprint arXiv:2209.11645, 2022.
  • [10] Alexander Kiselev and Leonid Ryzhik. Enhancement of the traveling front speeds in reaction-diffusion equations with advection. Annales de l’Institute Henri Poincaré C, Analyse non linéaire, 18(3):309–358, 2001.
  • [11] Alexander Kiselev, Roman Shterenberg, and Andrej Zlatoš. Relaxation enhancement by time-periodic flows. Indiana Univ. Math. J., 57(5):2137–2152, 2008.
  • [12] Alexei Novikov and Lenya Ryzhik. Boundary layers and kpp fronts in a cellular flow. Archive for rational mechanics and analysis, 184(1):23–48, 2007.
  • [13] Jean-Luc Thiffeault. Using multiscale norms to quantify mixing and transport. Nonlinearity, 25(2):R1, 2012.
  • [14] Juan Luis Vazquez. The Porous Medium Equation: Mathematical Theory. Oxford University Press, 2006.
  • [15] Dongyi Wei. Diffusion and mixing in fluid flow via the resolvent estimate. Science China Mathematics, 64(3):507–518, 2021.
  • [16] Michele Coti Zelati, Matias G Delgadino, and Tarek M Elgindi. On the relation between enhanced dissipation timescales and mixing rates. Communications on Pure and Applied Mathematics, 73(6):1205–1244, 2020.
  • [17] Andrej Zlatoš. Reaction-diffusion front speed enhancement by flows. Ann. Inst. H. Poincaré Anal. Non Linéaire, 28(5):711–726, 2011.