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

\stackMath

Nonuniqueness of Leray–Hopf solutions to the forced fractional Navier–Stokes Equations in three dimensions, up to the J. L. Lions exponent

Calvin Khor, Changxing Miao, and Xiaoyan Su
Abstract

In this paper, we show that for α(1/2,5/4)\alpha\in(1/2,5/4), there exists a force ff and two distinct Leray–Hopf flows u1,u2u_{1},u_{2} solving the forced fractional Navier–Stokes equation starting from rest. This shows that the J.L. Lions exponent is sharp in the class of Leray–Hopf solutions for the forced fractional Navier–Stokes equation.
Keywords: fractional Navier–Stokes Equation, nonuniqueness, linear instability, Leray–Hopf solution
MSC 2020 Classification: Primary 35F50; Secondary 35A02, 35Q35

1 Introduction

Consider Leray–Hopf solutions of the fractional Navier–Stokes equation starting from rest, with forcing ff:

tu+uu+p+(Δ)αu=f,u|t=0=0\displaystyle\partial_{t}u+u\cdot\nabla u+\nabla p+(-\Delta)^{\alpha}u=f,\quad u|_{t=0}=0 (1.1)

Recall that a solution is Leray–Hopf if it belongs to the natural energy class L(0,T;L2(3))L2(0,T;Hα(3))L^{\infty}(0,T;L^{2}(\mathbb{R}^{3}))\cap L^{2}(0,T;H^{\alpha}(\mathbb{R}^{3})), (In what follows, we abbreviate Lq(0,T;X(3))L^{q}(0,T;X(\mathbb{R}^{3})) as LTqXL^{q}_{T}X) attains the initial data strongly in L2L^{2}, and satisfies the energy inequality for a.e. t(0,T]t\in(0,T] and all s[0,t]s\in[0,t],

u(t)L22+stu(t)Hα2dtu(s)L22.\|u(t)\|^{2}_{L^{2}}+\int_{s}^{t}\|u(t^{\prime})\|_{H^{\alpha}}^{2}\mathop{}\!\mathrm{d}t^{\prime}\leq\|u(s)\|^{2}_{L^{2}}.

We prove nonuniqueness of Leray–Hopf solutions, following the recent work of Albritton–Brué–Colombo [ABC22] which proved the case α=1\alpha=1. They worked in the ‘similarity variables’ (here adapted to our fractional parabolic scaling)

ξ=x/t1/2α,τ=logt,u(x,t)=1t11/2αU(ξ,τ),p(x,t)=1t21/αP(ξ,τ),f(x,t)=1t21/2αF(ξ,τ),\displaystyle\begin{gathered}\xi=x/t^{1/2\alpha},\quad\tau=\log t,\quad u(x,t)=\frac{1}{t^{1-1/2\alpha}}U(\xi,\tau),\\ p(x,t)=\frac{1}{t^{2-1/\alpha}}P(\xi,\tau),\quad f(x,t)=\frac{1}{t^{2-1/2\alpha}}F(\xi,\tau),\end{gathered} (1.4)

which transforms (1.1) to

τU+(12α112αξξ)U+UξU+(Δ)αU+P=F.\displaystyle\partial_{\tau}U+\left(\frac{1}{2\alpha}-1-\frac{1}{2\alpha}\xi\cdot\nabla_{\xi}\right)U+U\cdot\nabla_{\xi}U+(-\Delta)^{\alpha}U+\nabla P=F. (1.5)

These variables are named as such because the powers of tt in (1.4) are chosen so that a stationary solution to (1.5) corresponds to a self-similar solution of (1.1). We can also write the equation for the vorticity ω(x,t)=1tΩ(ξ,τ)\omega(x,t)=\frac{1}{t}\Omega(\xi,\tau) in similarity variables,

τΩ+(112αξξ)Ω+UξΩΩξU+(Δ)αΩ=G.\displaystyle\partial_{\tau}\Omega+\Big{(}{-1}-\frac{1}{2\alpha}\xi\cdot\nabla_{\xi}\Big{)}\Omega+U\cdot\nabla_{\xi}\Omega-\Omega\cdot\nabla_{\xi}U+(-\Delta)^{\alpha}\Omega=G. (1.6)

Here, GG is the corresponding forcing, and UU is recovered from the vorticity Ω\Omega by inverting the curl via the Biot–Savart operator, i.e. U=BSΩU=\operatorname{BS}\Omega where

BSΩ(ξ)=3ξ×Ω(ξξ)4π|ξ|3dξ.\operatorname{BS}\Omega(\xi)=-\int_{\mathbb{R}^{3}}\frac{\xi^{\prime}\times\Omega(\xi-\xi^{\prime})}{4\pi|\xi^{\prime}|^{3}}\mathop{}\!\mathrm{d}\xi^{\prime}.

Most of their effort is devoted to the following construction of a compactly supported stationary unstable (forced) Euler vortex Ω¯=ξ×U¯\overline{\Omega}=\nabla_{\xi}\times\overline{U}. Recall that vector fields can be written in cylindrical coordinates V=Vrer+Vθeθ+Vzez.V=V^{r}e_{r}+V^{\theta}e_{\theta}+V^{z}e_{z}. Let Laps2L^{2}_{{\textup{aps}}} denote the space of square integrable, divergence-free, axisymmetric pure-swirl vector fields, i.e.

Laps2={VL2:divV=0,V(r,θ,z)=Vθ(r,z)eθ}.L^{2}_{{\textup{aps}}}=\{V\in L^{2}:\operatorname{div}V=0,\ V(r,\theta,z)=V^{\theta}(r,z)e_{\theta}\}.
Lemma 1.1 (Albritton–Brué–Colombo [ABC22]).

There exists a vector field Ω¯Cc(3;3)\overline{\Omega}\in C^{\infty}_{c}(\mathbb{R}^{3};\mathbb{R}^{3}) such that the operator

𝑳ΩU¯ΩΩU¯+UΩ¯Ω¯U\displaystyle-\boldsymbol{L}\Omega\coloneq\overline{U}\cdot\nabla\Omega-\Omega\cdot\nabla\overline{U}+U\cdot\nabla\overline{\Omega}-\overline{\Omega}\cdot\nabla U (1.7)

with domain

D(𝑳)={ΩLaps2:U¯ΩL2}\displaystyle D(\boldsymbol{L})=\{\Omega\in L^{2}_{{\textup{aps}}}:\overline{U}\cdot\nabla\Omega\in L^{2}\} (1.8)

has an unstable eigenvalue λ\lambda, i.e. an eigenvalue with Reλ>0\operatorname{Re}\lambda>0.

Their construction is not trivial because the other known unstable flows in 3D like shear flows or 2.5D Euler flows have no decay. They instead built on the recent instability result of Vishik [Vis18, Vis18a], for the forced 2D Euler equations. Albritton–Brué–Colombo observed that the 2D Euler equations arises as a formal limit as rr\to\infty of the 3D axisymmetric-without-swirl Euler equations. This allows a perturbative argument to transfer the instability from 2D to 3D.

The rest of Albritton–Brué–Colombo’s paper turns the linear instability in Euler into nonuniqueness for Navier–Stokes. In this note, we extend this nonuniqueness result to the equation (1.1), where α(1/2,5/4)\alpha\in(1/2,5/4). Well-posedness for (1.1) when α5/4\alpha\geq 5/4 is a well-known result due to J.L. Lions [Lio69]. Our result shows that this is sharp in a stronger way than the non-uniqueness of distributional solutions in [LT20], albeit with a non-zero forcing term.

Following Albritton–Brué–Colombo, we proceed as follows. First, we show that for β\beta sufficiently large, the linearised operator 𝑳α,β:D(𝑳α,β)Laps2Laps2\boldsymbol{L}_{\alpha,\beta}:D(\boldsymbol{L}_{\alpha,\beta})\subset L^{2}_{{\textup{aps}}}\to L^{2}_{{\textup{aps}}} around βΩ¯\beta\overline{\Omega}, defined by

D(𝑳α,β){ΩLaps2:ΩH2α,ξΩL2},𝑳α,βΩ(112αξξ+(Δ)α)Ωβ𝑳Ω,\displaystyle\begin{aligned} D(\boldsymbol{L}_{\alpha,\beta})\coloneq\{\Omega\in L^{2}_{{\textup{aps}}}:\Omega\in H^{2\alpha},\ \xi\cdot\nabla\Omega\in L^{2}\},\\ -\boldsymbol{L}_{\alpha,\beta}\Omega\coloneq(-1-\frac{1}{2\alpha}\xi\cdot\nabla_{\xi}+(-\Delta)^{\alpha})\Omega-\beta\boldsymbol{L}\Omega,\end{aligned} (1.9)

has an unstable eigenvalue. This corresponds to an unstable eigenvalue for the operator 𝑻α,β\boldsymbol{T}_{\!\alpha,\beta}, which is 𝑳α,β\boldsymbol{L}_{\alpha,\beta} written in the velocity formulation (see (2.9) below). Let η\eta be the corresponding eigenvector of 𝑻α,β\boldsymbol{T}_{\!\alpha,\beta}, and define UL=Re(eλtη)U^{\textsf{L}}=\operatorname{Re}(e^{\lambda t}\eta), which solves the linear equation tUL=𝑳α,βUL\partial_{t}U^{\textsf{L}}=\boldsymbol{L}_{\alpha,\beta}U^{\textsf{L}}. To solve (1.5), we use the ansatz

U=βU¯+UL+UPU=\beta\overline{U}+U^{\textsf{L}}+U^{\textsf{P}}

and find UPU^{\textsf{P}} by a fixed point argument in Subsection 3.2. Finally, undoing the similarity variable transformation proves the following:

Theorem 1.2.

Let α(1/2,5/4)\alpha\in(1/2,5/4) and U¯=BSΩ¯\overline{U}=\operatorname{BS}\overline{\Omega} be the smooth unstable velocity from Lemma 1.1. Set τ=logt\tau=\log t, ξ=xt1/2α\xi=xt^{-1/2\alpha}. Then for τT\tau\leq T, u¯(x,t)βt11/2αU¯(ξ,τ)\overline{u}(x,t)\coloneq\frac{\beta}{t^{1-1/2\alpha}}\overline{U}(\xi,\tau) and u(x,t)1t11/2αU(ξ)u(x,t)\coloneq\frac{1}{t^{1-1/2\alpha}}U(\xi), where UU is constructed in Subsection 3.2, are two distinct Leray–Hopf solutions to (1.1) with force f(x,t)=1t21/2αF(ξ,τ)f(x,t)=\frac{1}{t^{2-1/2\alpha}}F(\xi,\tau) on a time interval [0,eT][0,e^{T}], where

Fβ(12α112αξξ)U¯+β2U¯ξU¯+β(Δ)αU¯.F\coloneq\beta\left(\frac{1}{2\alpha}-1-\frac{1}{2\alpha}\xi\cdot\nabla_{\xi}\right)\overline{U}+\beta^{2}\overline{U}\cdot\nabla_{\xi}\overline{U}+\beta(-\Delta)^{\alpha}\overline{U}.

In addition, u¯\overline{u} and uu belong to the borderline space LTL32α1,L^{\infty}_{T}L^{\frac{3}{2\alpha-1},\infty} and for any k,j0k,j\geq 0, p[2,]p\in[2,\infty] and t<eTt<e^{T},

tk/2u¯(t)W̊kα,p+tk/2u(t)W̊kα,p+tj+1+k/2tjf(t)W̊kα,pktp+32αp1,t^{k/2}\|\overline{u}(t)\|_{\mathring{W}^{k\alpha,p}}+t^{k/2}\|u(t)\|_{\mathring{W}^{k\alpha,p}}+t^{j+1+k/2}\|\partial_{t}^{j}f(t)\|_{\mathring{W}^{k\alpha,p}}\lesssim_{k}t^{\frac{p+3}{2\alpha p}-1}, (1.10)

where we have written vW̊s,p(Δ)s/2vLp\|v\|_{\mathring{W}^{s,p}}\coloneq\|(-\Delta)^{s/2}v\|_{L^{p}}.

Remark 1.

The only place that α<5/4\alpha<5/4 is used is in the last step to attain the initial data via (1.10). The proof also extends to any subcritical diffusion just below the 5/45/4 exponent, for instance a logarithmically modified version of (Δ)5/4(-\Delta)^{5/4}. The condition α>1/2\alpha>1/2 ensures that the diffusion is stronger than the effects from the material derivative, which we use in Lemma 2.5, Lemma 3.2, and Proposition 3.6 below.

Remark 2.

The methods used in Vishik and Albritton–Brué–Colombo, and hence here, strongly rely on the equation having a forcing term that we can choose. In addition to relying on Vishik’s unstable vortex, which only solves a forced Euler equation, this gives us the flexibility to linearise around vector fields that do not solve the unforced equations.

Remark 3.

Recently, Albritton and Colombo have shown [AC22] non-uniqueness for the hypodissipative (i.e. α<1\alpha<1 in our notation) forced 2D Navier–Stokes equation, as announced in [ABC22]. In addition, Albritton, Brué and Colombo have recently shown [ABC22a] that the result of [ABC22] extends to 𝕋3\mathbb{T}^{3} and bounded domains of 3\mathbb{R}^{3} when the no-slip boundary condition is enforced. One can also see the recent preprint [LR22] which shows well-posedness for sufficiently small forces in various senses, and initial data in the critical space BMO1\textup{BMO}^{-1}.

2 Linear instability

To replace the heat semigroup estimates used in [ABC22], we need to recall the following basic estimate for the fractional heat equation. For given functions u0=u0(x)u_{0}=u_{0}(x) and f=f(x,t)f=f(x,t), we write u=et(Δ)αu0u=\textup{e}^{-t(-\Delta)^{\alpha}}u_{0} for the solution to

tu+(Δ)αu=0,u|t=0=u0,\partial_{t}u+(-\Delta)^{\alpha}u=0,\qquad u|_{t=0}=u_{0},

and write u=𝒢0fu=\mathcal{G}_{0}f where 𝒢0f=0te(ts)(Δ)α(f(,s))ds\mathcal{G}_{0}f=\int_{0}^{t}\textup{e}^{-(t-s)(-\Delta)^{\alpha}}(f(\cdot,s))\mathop{}\!\mathrm{d}s is the solution to

tu+(Δ)αu=f,u|t=0=0.\partial_{t}u+(-\Delta)^{\alpha}u=f,\qquad u|_{t=0}=0.

Then, the solution of the general initial value problem

tu+(Δ)αu=f,u|t=0=u0\partial_{t}u+(-\Delta)^{\alpha}u=f,\qquad u|_{t=0}=u_{0}

is the sum u=et(Δ)αu0+𝒢0f.u=\textup{e}^{-t(-\Delta)^{\alpha}}u_{0}+\mathcal{G}_{0}f.

Lemma 2.1.

[MYZ08, Lemma 3.1] For 1rp1\leq r\leq p\leq\infty and β0\beta\geq 0,

(Δ)β/2et(Δ)αu0Lpt32α(β+1r1p)u0Lr.\|(-\Delta)^{\beta/2}\textup{e}^{-t(-\Delta)^{\alpha}}u_{0}\|_{L^{p}}\lesssim t^{-\frac{3}{2\alpha}(\beta+\frac{1}{r}-\frac{1}{p})}\|u_{0}\|_{L^{r}}.
Lemma 2.2.

𝑳α,β\boldsymbol{L}_{\alpha,\beta} with the domain D(𝐋α,β)D(\boldsymbol{L}_{\alpha,\beta}) as in (1.9) is a closed, densely defined operator.

Proof.

If ΩnD(𝑳α,β)\Omega_{n}\in D(\boldsymbol{L}_{\alpha,\beta}) converges to Ω\Omega and 𝑳α,βΩL2\boldsymbol{L}_{\alpha,\beta}\Omega\in L^{2} converges to WW, then W=𝑳α,βΩW=\boldsymbol{L}_{\alpha,\beta}\Omega in the sense of distributions. Since 𝑳\boldsymbol{L} is closed, in order to show that 𝑳α,β\boldsymbol{L}_{\alpha,\beta} is closed, we only need to show that each ZL2Z\in L^{2} corresponds to a unique solution ΩH2α\Omega\in H^{2\alpha} to the equation (𝑳α,β+1+β𝑳)Ω=Z(-\boldsymbol{L}_{\alpha,\beta}+1+\beta\boldsymbol{L})\Omega=Z, i.e.

12αξΩ+(Δ)αΩ=Z.-\frac{1}{2\alpha}\xi\cdot\nabla\Omega+(-\Delta)^{\alpha}\Omega=Z.

We deal with the unbounded term here and in the sequel by using the fact that transforming back to physical variables by ξ=x/t1/2α\xi=x/t^{1/2\alpha} converts ξξ\xi\cdot\nabla_{\xi} into a time derivative. Specifically, we consider the functions

h(x,t)=Ω(x/t1/2α),g(x,t)=1tZ(x/t1/2α).\displaystyle h(x,t)=\Omega(x/t^{1/2\alpha}),\quad g(x,t)=\frac{1}{t}Z(x/t^{1/2\alpha}). (2.1)

It is easy to check that hh solves the fractional heat equation

th+(Δ)αh=g,\partial_{t}h+(-\Delta)^{\alpha}h=g,

and attains zero initial data in L2L^{2} since α>0\alpha>0:

hL2=t3/4αΩL2t00.\|h\|_{L^{2}}=t^{3/4\alpha}\|\Omega\|_{L^{2}}\xrightarrow[t\to 0]{}0.

Since gL2=t1+3/4αZL2\|g\|_{L^{2}}=t^{-1+3/4\alpha}\|Z\|_{L^{2}} is integrable, we obtain via the theory of the fractional heat equation that Ω=h(,1)H2α\Omega=h(\cdot,1)\in H^{2\alpha}. Indeed, firstly, h(,1/2)h(\cdot,1/2) is L2L^{2} by Lemma 2.1. Then for times t1/2t\geq 1/2, by writing

h(x,t)=et(Δ)αh(x,1/2)+𝒢0(g(,t+1/2))(x,t1/2),h(x,t)=\textup{e}^{-t(-\Delta)^{\alpha}}h(x,1/2)+\mathcal{G}_{0}(g(\cdot,t+1/2))(x,t-1/2),

since g(,t+1/2)g(\cdot,t+1/2) is not singular at t=0t=0, Lemma 2.1 shows that (Δ)αhC([1/2,1];L2)(-\Delta)^{\alpha}h\in C([1/2,1];L^{2}), so we obtain the claimed result. Hence, 𝑳α,β\boldsymbol{L}_{\alpha,\beta} is closed. ∎

We write 𝑳α,β=1β𝑫α+𝑴+𝑲+SS\boldsymbol{L}^{\prime}_{\alpha,\beta}=\frac{1}{\beta}\boldsymbol{D}_{\alpha}+\boldsymbol{M}+\boldsymbol{K}+\SS as a dissipative term, main term, compact term, and small term respectively, where:

𝑫αΩ\displaystyle-\boldsymbol{D}_{\alpha}\Omega 34αΩ12αξΩ+(Δ)αΩ,\displaystyle\coloneq-\frac{3}{4\alpha}\Omega-\frac{1}{2\alpha}\xi\cdot\nabla\Omega+(-\Delta)^{\alpha}\Omega,
𝑴Ω\displaystyle-\boldsymbol{M}\Omega U¯Ω,\displaystyle\coloneq\overline{U}\cdot\nabla\Omega,
𝑲Ω\displaystyle-\boldsymbol{K}\Omega UΩ¯,\displaystyle\coloneq U\cdot\nabla\overline{\Omega},
SSΩ\displaystyle-\SS\Omega Ω¯U+ΩU¯.\displaystyle\coloneq\overline{\Omega}\cdot\nabla U+\Omega\cdot\nabla\overline{U}.

Then 𝑳α,β=β𝑳α,β+134α\boldsymbol{L}_{\alpha,\beta}=\beta\boldsymbol{L}^{\prime}_{\alpha,\beta}+1-\frac{3}{4\alpha}. The constant 134α1-\frac{3}{4\alpha} is added because ξΩ\xi\cdot\nabla\Omega is not skew-adjoint.

Lemma 2.3 (Resolvent estimates via inviscid limit).

For all β>0\beta>0 and all λ\lambda with Reλ>μSSLaps2L2\operatorname{Re}\lambda>\mu\coloneq\|\SS\|_{L^{2}_{\textup{aps}}\to L^{2}},

R(λ,β1𝑫α+𝑴+SS)Laps2Laps21Reλμ,\displaystyle\|R(\lambda,\beta^{-1}\boldsymbol{D}_{\alpha}+\boldsymbol{M}+\SS)\|_{L^{2}_{{\textup{aps}}}\to L^{2}_{{\textup{aps}}}}\leq\frac{1}{\operatorname{Re}\lambda-\mu}, (2.2)

and the resolvent of β1𝐃α+𝐌+SS\beta^{-1}\boldsymbol{D}_{\alpha}+\boldsymbol{M}+\SS converges in the strong topology, i.e. for all Ω0Laps2\Omega_{0}\in L^{2}_{{\textup{aps}}},

R(λ,β1𝑫α+𝑴+SS)Ω0βR(λ,𝑴+SS)Ω0.R(\lambda,\beta^{-1}\boldsymbol{D}_{\alpha}+\boldsymbol{M}+\SS)\Omega_{0}\xrightarrow[\beta\to\infty]{}R(\lambda,\boldsymbol{M}+\SS)\Omega_{0}.
Proof.

We want to use that the Laplace transform of the semigroup is the resolvent (see (2.6) below), so let Ωβ\Omega^{\beta} be the solution to the initial value problem (τβ1𝑫α𝑴SS)Ω=0(\partial_{\tau}-\beta^{-1}\boldsymbol{D}_{\alpha}-\boldsymbol{M}-\SS)\Omega=0, i.e.

{τΩβ1β(34α+12αξ(Δ)α)Ωβ+U¯Ωβ=SSΩβ,Ωβ|τ=0=Ω0,\left\{\begin{aligned} \partial_{\tau}\Omega^{\beta}-\frac{1}{\beta}\Big{(}\frac{3}{4\alpha}+\frac{1}{2\alpha}\xi\cdot\nabla-(-\Delta)^{\alpha}\Big{)}\Omega^{\beta}+\overline{U}\cdot\nabla\Omega^{\beta}&=\SS\Omega^{\beta},\\ \Omega^{\beta}|_{\tau=0}&=\Omega_{0},\end{aligned}\right.

and let Ω\Omega solve the β\beta\to\infty limit equation

{τΩ+U¯Ω=SSΩ,Ω|τ=0=Ω0,\left\{\begin{aligned} \partial_{\tau}\Omega+\overline{U}\cdot\nabla\Omega&=\SS\Omega,\\ \Omega|_{\tau=0}&=\Omega_{0},\end{aligned}\right.

For smooth, compactly supported data, Ω\Omega exists and is smooth. Ωβ\Omega^{\beta} can be shown to exist as a strong solution in C0H2αC^{0}H^{2\alpha} by again undoing the similarity variable transform which removes the term ξ\xi\cdot\nabla to get a more standard fractional heat equation with drift, similarly to (2.1). Specifically, one sets hβ(x,t)=Ωβ(xt1/2α,βlogt)h^{\beta}(x,t)=\Omega^{\beta}(xt^{-1/2\alpha},\beta\log t) and fβ(x,t)=βt1SSΩβ(xt1/2α,βlogt)f^{\beta}(x,t)=\beta t^{-1}\SS\Omega^{\beta}(xt^{-1/2\alpha},\beta\log t) to get

thβ(34α(Δ)α)hβ+βu¯hβ=fβ.\partial_{t}h^{\beta}-\left(\frac{3}{4\alpha}-(-\Delta)^{\alpha}\right)h^{\beta}+\beta\overline{u}\cdot\nabla h^{\beta}=f^{\beta}.

Notably, u¯\overline{u} is smooth in both space and time, as the initial data is prescribed at t=1t=1, so well-posedness is easily proven first for fβ=0f^{\beta}=0, then for fβ0f^{\beta}\neq 0 with Duhamel’s formula.

From the energy inequality for Ωβ\Omega^{\beta}, and from the Lagrangian formulation for Ω\Omega, we have the estimates

ΩL22+ΩβL22eτSSLaps2L2Ω0L22=eτμΩ0L22.\displaystyle\|\Omega\|_{L^{2}}^{2}+\|\Omega^{\beta}\|_{L^{2}}^{2}\lesssim\textup{e}^{\tau\smash{\|\SS\|_{L^{2}_{{\textup{aps}}}\to L^{2}}}}\|\Omega_{0}\|^{2}_{L^{2}}=e^{\tau\mu}\|\Omega_{0}\|^{2}_{L^{2}}. (2.3)

For the limit, we study the equation for the difference Ω~ΩβΩ\tilde{\Omega}\coloneq\Omega^{\beta}-\Omega,

(τβ1𝑫α𝑴SS)Ω~=β1𝑫αΩ,(\partial_{\tau}-\beta^{-1}\boldsymbol{D}_{\alpha}-\boldsymbol{M}-\SS)\tilde{\Omega}=\beta^{-1}\boldsymbol{D}_{\alpha}\Omega,

initially for Ω0Cc\Omega_{0}\in C^{\infty}_{c}. Testing against Ω~\tilde{\Omega} in L2L^{2}, and using Young’s inequality (aba2/2+b2/2ab\leq a^{2}/2+b^{2}/2) gives

ddτΩ~L22+1β(Δ)α/2Ω~L22μΩ~L22+1β𝑫αΩ,Ω~Hα,HαμΩ~L22+12β𝑫αΩHα2+12β(Ω~L22+(Δ)α/2Ω~L22).\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}\tau}\|\tilde{\Omega}\|_{L^{2}}^{2}+\frac{1}{\beta}\|(-\Delta)^{\alpha/2}\tilde{\Omega}\|_{L^{2}}^{2}\leq\mu\|\tilde{\Omega}\|_{L^{2}}^{2}+\frac{1}{\beta}\langle\boldsymbol{D}_{\alpha}\Omega,\tilde{\Omega}\rangle_{H^{-\alpha},H^{\alpha}}\\ \leq\mu\|\tilde{\Omega}\|_{L^{2}}^{2}+\frac{1}{2\beta}\|\boldsymbol{D}_{\alpha}\Omega\|_{H^{-\alpha}}^{2}+\frac{1}{2\beta}\Big{(}\|\tilde{\Omega}\|_{L^{2}}^{2}+\|(-\Delta)^{\alpha/2}\tilde{\Omega}\|_{L^{2}}^{2}\Big{)}. (2.4)

Absorbing 12β(Δ)α/2Ω~L22\frac{1}{2\beta}\|(-\Delta)^{\alpha/2}\tilde{\Omega}\|_{L^{2}}^{2} into the left-hand side, integrating in 0τT0\leq\tau\leq T and using that Ω~|τ=0=0\tilde{\Omega}|_{\tau=0}=0, we obtain for each fixed TT that

Ω~L(0,T;L2)2Tβ1suptT𝑫αΩ(,t)Hα2β0.\displaystyle\|\tilde{\Omega}\|_{L^{\infty}(0,T;L^{2})}^{2}\lesssim_{T}\beta^{-1}\sup_{t\leq T}\|\boldsymbol{D}_{\alpha}\Omega(\cdot,t)\|_{H^{-\alpha}}^{2}\xrightarrow[\beta\to\infty]{}0. (2.5)

Recall that for a densely defined operator 𝑨\boldsymbol{A} with sufficient growth bounds on the semigroup et𝑨e^{t\boldsymbol{A}}, the resolvent R(λ,𝑨)R(\lambda,\boldsymbol{A}) can be written as the Laplace transform

R(λ,𝑨)Ω0=0et(λ𝑨)Ω0dt.\displaystyle R(\lambda,\boldsymbol{A})\Omega_{0}=\int_{0}^{\infty}\textup{e}^{-t(\lambda-\boldsymbol{A})}\Omega_{0}\mathop{}\!\mathrm{d}t. (2.6)

In our case, this formula and the growth bound (2.3) shows the boundedness (2.2). It follows that for all λ\lambda with Reλ>μ\operatorname{Re}\lambda>\mu,

R(λ,β1𝑫α+𝑴+SS)Ω0R(λ,𝑴+SS)Ω0L2\displaystyle\|R(\lambda,\beta^{-1}\boldsymbol{D}_{\alpha}+\boldsymbol{M}+\SS)\Omega_{0}-R(\lambda,\boldsymbol{M}+\SS)\Omega_{0}\|_{L^{2}}
0eReλtet(β1𝑫α+𝑴+SS)Ω0et(𝑴+SS)Ω0dt\displaystyle\leq\int_{0}^{\infty}\textup{e}^{-\operatorname{Re}\lambda t}\|\textup{e}^{t(\beta^{-1}\boldsymbol{D}_{\alpha}+\boldsymbol{M}+\SS)}\Omega_{0}-\textup{e}^{t(\boldsymbol{M}+\SS)}\Omega_{0}\|\mathop{}\!\mathrm{d}t
Ω~L(0,T;L2)0TeReλtdt+TeReλt(Ω(t)L2+Ωβ(t)L2)dt\displaystyle\leq\|\tilde{\Omega}\|_{L^{\infty}(0,T;L^{2})}\int_{0}^{T}\textup{e}^{-\operatorname{Re}\lambda t}\mathop{}\!\mathrm{d}t+\int_{T}^{\infty}\textup{e}^{-\operatorname{Re}\lambda t}(\|\Omega(t)\|_{L^{2}}+\|\Omega^{\beta}(t)\|_{L^{2}})\mathop{}\!\mathrm{d}t
1ReλΩ~L(0,T;L2)eReλT+1ReλμΩ0L2e(Reλμ)T.\displaystyle\leq\frac{1}{\operatorname{Re}\lambda}\|\tilde{\Omega}\|_{L^{\infty}(0,T;L^{2})}e^{-\operatorname{Re}\lambda T}+\frac{1}{\operatorname{Re}\lambda-\mu}\|\Omega_{0}\|_{L^{2}}e^{-(\operatorname{Re}\lambda-\mu)T}.

The strong convergence now follows from (2.5) for all compactly supported data, by first taking T1T\gg 1 and then β\beta\to\infty. We can then cover the case Ω0L2\Omega_{0}\in L^{2} by density. ∎

Lemma 2.4.

Let λ\lambda_{\infty} be an unstable eigenvalue of 𝐋\boldsymbol{L}. Then, for each ϵ(0,Reλμ)\epsilon\in(0,\operatorname{Re}\lambda_{\infty}-\mu), there is β0\beta_{0} such that for all β>β0\beta>\beta_{0}, 𝐋α,β\boldsymbol{L}_{\alpha,\beta} has an unstable eigenvalue λα,β=βλα,β+(134α)\lambda_{\alpha,\beta}=\beta\lambda^{\prime}_{\alpha,\beta}+(1-\frac{3}{4\alpha}), with |λλα,β|<ϵ|\lambda_{\infty}-\lambda^{\prime}_{\alpha,\beta}|<\epsilon.

Proof.

To prove this, we let γρ(𝑳){Reλ>μ}\gamma\subset\rho(\boldsymbol{L})\cap\{\operatorname{Re}\lambda>\mu\} be a sufficiently small circle around the unstable eigenvalue λ\lambda_{\infty} which encloses no other part of σ(𝑳)\sigma(\boldsymbol{L}). We need to show that we can define the spectral projection operator

Prα,β12πiγR(λ,𝑳α,β)dλ.\displaystyle\operatorname{Pr}_{\alpha,\beta}\coloneq\frac{1}{2\pi i}\int_{\gamma}R(\lambda,\boldsymbol{L}^{\prime}_{\alpha,\beta})\mathop{}\!\mathrm{d}\lambda. (2.7)

and show it converges in norm to the corresponding spectral projection for 𝑳=𝑴+SS+𝑲\boldsymbol{L}=\boldsymbol{M}+\SS+\boldsymbol{K},

Pr12πiγR(λ,𝑳)dλ.\displaystyle\operatorname{Pr}\coloneq\frac{1}{2\pi i}\int_{\gamma}R(\lambda,\boldsymbol{L})\mathop{}\!\mathrm{d}\lambda. (2.8)

From Lemma 1.1, Pr\operatorname{Pr} is non-trivial, so Prα,β\operatorname{Pr}_{\alpha,\beta} is nontrivial for sufficiently large β\beta. This means that γ\gamma encloses an eigenvalue of 𝑳α,β\boldsymbol{L}_{\alpha,\beta} with finite multiplicity, which proves the Lemma.

So we just need to justify (2.7). Consider the identity

λ𝑳α,β=(λβ1𝑫α𝑴SS)(IR(λ,β1𝑫α+𝑴+SS)𝑲),\lambda-\boldsymbol{L}^{\prime}_{\alpha,\beta}=(\lambda-\beta^{-1}\boldsymbol{D}_{\alpha}-\boldsymbol{M}-\SS)\big{(}I-R(\lambda,\beta^{-1}\boldsymbol{D}_{\alpha}+\boldsymbol{M}+\SS)\boldsymbol{K}\big{)},

which makes sense for Reλ>μ\operatorname{Re}\lambda>\mu by (2.2). As R(λ,β1𝑫α+𝑴+SS)R(λ,𝑴+SS)R(\lambda,\beta^{-1}\boldsymbol{D}_{\alpha}+\boldsymbol{M}+\SS)\to R(\lambda,\boldsymbol{M}+\SS) in the strong topology and 𝑲\boldsymbol{K} is compact, we have the norm convergence

R(λ,β1𝑫α+𝑴+SS)𝑲βR(λ,𝑴+SS)𝑲,R(\lambda,\beta^{-1}\boldsymbol{D}_{\alpha}+\boldsymbol{M}+\SS)\boldsymbol{K}\xrightarrow[\beta\to\infty]{}R(\lambda,\boldsymbol{M}+\SS)\boldsymbol{K},

locally uniformly in λ\lambda. In addition, for λρ(𝑳){Reλ>μ}ρ(𝑴+SS)\lambda\in\rho(\boldsymbol{L})\cap\{\operatorname{Re}\lambda>\mu\}\subset\rho(\boldsymbol{M}+\SS), the identity

λ𝑳=λ(𝑴+SS+𝑲)=(λ𝑴+SS)(IR(λ,𝑴+SS)𝑲)\lambda-\boldsymbol{L}=\lambda-(\boldsymbol{M}+\SS+\boldsymbol{K})=(\lambda-\boldsymbol{M}+\SS)(I-R(\lambda,\boldsymbol{M}+\SS)\boldsymbol{K})

shows that IR(λ,𝑴+SS)𝑲I-R(\lambda,\boldsymbol{M}+\SS)\boldsymbol{K} is invertible. As the set of invertible operators is open, it follows that IR(λ,β1𝑫α+𝑴+SS)𝑲I-R(\lambda,\beta^{-1}\boldsymbol{D}_{\alpha}+\boldsymbol{M}+\SS)\boldsymbol{K} is invertible for β1\beta\gg 1, locally uniformly in λ\lambda. Hence, R(λ,𝑳α,β)R(\lambda,\boldsymbol{L}^{\prime}_{\alpha,\beta}) for β1\beta\gg 1 makes sense on the compact set γ\gamma, which verifies (2.7). This finishes the proof. ∎

Lemma 2.5.

For β1\beta\gg 1, the linearised Navier–Stokes velocity operator in similarity variables,

𝑻α,β:D(𝑻α,β)L2L2,D(𝑻α,β){UH2α:ξUL2},𝑻α,βU(12α112αξξ)U+β(U¯ξU+UξU¯)+(Δ)αU,\displaystyle\begin{aligned} \boldsymbol{T}_{\!\alpha,\beta}&:D(\boldsymbol{T}_{\!\alpha,\beta})\subset L^{2}\to L^{2},\\ D(\boldsymbol{T}_{\!\alpha,\beta})&\coloneq\{U\in H^{2\alpha}:\xi\cdot\nabla U\in L^{2}\},\\ -\boldsymbol{T}_{\!\alpha,\beta}U&\coloneq\Big{(}\frac{1}{2\alpha}-1-\frac{1}{2\alpha}\xi\cdot\nabla_{\xi}\Big{)}U+\beta\mathbb{P}(\overline{U}\cdot\nabla_{\xi}U+U\cdot\nabla_{\xi}\overline{U})\\ &\quad+(-\Delta)^{\alpha}U,\end{aligned} (2.9)

has the same unstable eigenvalue λα,β\lambda_{\alpha,\beta} as 𝐋α,β\boldsymbol{L}_{\alpha,\beta}. Here, \mathbb{P} denotes the usual Leray projector to the space of divergence-free fields.

Proof.

Algebraically, this is true because the operators in velocity and vorticity formulation are linked by conjugation with the curl operator,

×𝑻α,βV=𝑳α,β×V.\nabla\times\boldsymbol{T}_{\!\alpha,\beta}V=\boldsymbol{L}_{\alpha,\beta}\nabla\times V.

Hence, 𝑻α,β\boldsymbol{T}_{\!\alpha,\beta} has the same unstable eigenvalue as 𝑳α,β\boldsymbol{L}_{\alpha,\beta}, and if ζ\zeta is the eigenfunction for 𝑳α,β\boldsymbol{L}_{\alpha,\beta}, then η=BSζ\eta=\operatorname{BS}\zeta is formally the eigenfunction for 𝑻α,β\boldsymbol{T}_{\!\alpha,\beta}. The only detail to check is that ηD(𝑻α,β)\eta\in D(\boldsymbol{T}_{\!\alpha,\beta}), and the conditions defining D(𝑻α,β)D(\boldsymbol{T}_{\!\alpha,\beta}) are all satisfied save for potentially ηL2\eta\in L^{2}. As BS\operatorname{BS} is bounded from L2L^{2} to L6L^{6}, and from L1L^{1} to L3/2L^{3/2}, it suffices by interpolation to show that ζL1\zeta\in L^{1}.

As an eigenfunction of 𝑳α,β\boldsymbol{L}_{\alpha,\beta}, ζ\zeta satisfies the equation

λα,βζ(1+12αξ(Δ)α)ζ=𝑳ζ.\lambda_{\alpha,\beta}\zeta-\Big{(}1+\frac{1}{2\alpha}\xi\cdot\nabla-(-\Delta)^{\alpha}\Big{)}\zeta=\boldsymbol{L}\zeta.

We again undo the similarity variables by setting ξ=x/t1/2α\xi=x/t^{1/2\alpha} and

h(x,t)tλα,β1ζ(xt1/2α),g(x,t)tλα,β2(𝑳ζ)(xt1/2α).h(x,t)\coloneq t^{\lambda_{\alpha,\beta}-1}\zeta\Big{(}\frac{x}{t^{1/2\alpha}}\Big{)},\quad g(x,t)\coloneq t^{\lambda_{\alpha,\beta}-2}(\boldsymbol{L}\zeta)\Big{(}\frac{x}{t^{1/2\alpha}}\Big{)}.

The corresponding equation is

th+(Δ)αh=g,h|t=0=0,\partial_{t}h+(-\Delta)^{\alpha}h=g,\quad h|_{t=0}=0,

where the initial data is attained in L2L^{2} once Reλα,β>134α\operatorname{Re}\lambda_{\alpha,\beta}>1-\frac{3}{4\alpha}, using β1\beta\gg 1. Observe that 𝑳ζL2\boldsymbol{L}\zeta\in L^{2} is compactly supported. It follows from the integral formulation of the Duhamel-type operator 𝒢0\mathcal{G}_{0} that gC0([0,T];L1)g\in C^{0}([0,T];L^{1}) for each TT (again using β1\beta\gg 1). Lemma 2.1 shows that hh is also C0L1C^{0}L^{1}, so in particular ζ=h(,1)L1\zeta=h(\cdot,1)\in L^{1}, as needed. ∎

3 Nonlinear construction

3.1 The semigroup generated by 𝑻α,β\boldsymbol{T}_{\!\alpha,\beta}

Here, we adapt the arguments of [JS15, §2] to prove some results for the spectrum and the semigroup generated by 𝑻α,β\boldsymbol{T}_{\!\alpha,\beta}. Since 𝑻α,β𝑻α,0\boldsymbol{T}_{\!\alpha,\beta}-\boldsymbol{T}_{\alpha,0} will turn out to be 𝑻α,0\boldsymbol{T}_{\alpha,0}-compact, we first study the spectrum of 𝑻α,0\boldsymbol{T}_{\alpha,0}.

Lemma 3.1.

σ(𝑻α,0){Reλ154α}\sigma(\boldsymbol{T}_{\alpha,0})\subset\{\operatorname{Re}\lambda\leq 1-\frac{5}{4\alpha}\}.

Proof.

Let Reλ>154α.\operatorname{Re}\lambda>1-\frac{5}{4\alpha}. To show that λρ(𝑻α,0)\lambda\in\rho(\boldsymbol{T}_{\alpha,0}), we need to show that (λ𝑻α,0)U=F(\lambda-\boldsymbol{T}_{\alpha,0})U=F can be solved for UU with UH2αFL2\|U\|_{H^{2\alpha}}\lesssim\|F\|_{L^{2}}. This will be proven by once again undoing the similarity transform, as in the proof of Lemma 2.2. Suppose we are given UU. The definition (2.9) leads us to define

h(x,t)tλ1+12αU(xt1/2α),g(x,t)=tλ2+12αF(xt1/2α).h(x,t)\coloneq t^{\lambda-1+\frac{1}{2\alpha}}U\left(\frac{x}{t^{1/2\alpha}}\right),\quad g(x,t)=t^{\lambda-2+\frac{1}{2\alpha}}F\left(\frac{x}{t^{1/2\alpha}}\right).

They solve th+(Δ)αh=g\partial_{t}h+(-\Delta)^{\alpha}h=g attaining the data h|t=0=0h|_{t=0}=0 strongly in L2L^{2}, since Reλ>154α\operatorname{Re}\lambda>1-\frac{5}{4\alpha}. Since gC0((0,T];L2)g\in C^{0}((0,T];L^{2}), it follows that hC0((0,T];H2α)h\in C^{0}((0,T];H^{2\alpha}), in particular U=h(,1)H2αU=h(\cdot,1)\in H^{2\alpha} with (Δ)2αUL2FL2.\|(-\Delta)^{2\alpha}U\|_{L^{2}}\lesssim\|F\|_{L^{2}}. Read backwards, the above explains how to construct UU given the force FF, using the well-posedness of the fractional heat equation. This shows that {Reλ>154α}ρ(𝑻α,0)\{\operatorname{Re}\lambda>1-\frac{5}{4\alpha}\}\subset\rho(\boldsymbol{T}_{\alpha,0}), hence the result. ∎

Lemma 3.2.

𝑲α,β\boldsymbol{K}_{\alpha,\beta} defined by 𝐊α,βU𝐓α,βU𝐓α,0U=β(U¯U+UU¯)\boldsymbol{K}_{\alpha,\beta}U\coloneq\boldsymbol{T}_{\!\alpha,\beta}U-\boldsymbol{T}_{\alpha,0}U=\beta\mathbb{P}(\overline{U}\cdot\nabla U+U\cdot\nabla\overline{U}) is 𝐓α,0\boldsymbol{T}_{\alpha,0}-compact.

Proof.

By the estimates in the proof of Lemma 3.1, the sequence UnU_{n} belongs to D(𝑻α,β)D(\boldsymbol{T}_{\!\alpha,\beta}) with 𝑻α,0Un\boldsymbol{T}_{\alpha,0}U_{n} bounded in norm, and is in particular bounded in H2αH^{2\alpha}. Hence, the relative compactness follows by the compactness of the support of U¯\overline{U}, which allows the use of the compact embedding of Sobolev spaces on bounded domains. ∎

Lemma 3.3.

The semigroup eτ𝐓α,β\textup{e}^{\tau\boldsymbol{T}_{\!\alpha,\beta}} is well-defined and strongly continuous. It is the solution operator for the initial value problem

τU𝑻α,βU=0,U|τ=0=U0\partial_{\tau}U-\boldsymbol{T}_{\!\alpha,\beta}U=0,\quad U|_{\tau=0}=U_{0}
Proof.

We undo the similarity variables to find the more standard problem

tu+β(u¯u+uu¯)+(Δ)αu=0,u|t=1=U0.\partial_{t}u+\beta\mathbb{P}(\overline{u}\cdot\nabla u+u\cdot\nabla\overline{u})+(-\Delta)^{\alpha}u=0,\quad u|_{t=1}=U_{0}.
u(t)L2+(t1)k/2αku(t)L2U¯,Tu(1)L2,\|u(t)\|_{L^{2}}+(t-1)^{k/2\alpha}\|\nabla^{k}u(t)\|_{L^{2}}\lesssim_{\overline{U},T}\|u(1)\|_{L^{2}},

which translate to the following estimates for UU,

U(τ)L2+τk/2αkU(τ)L2U¯,TU0L2.\displaystyle\|U(\tau)\|_{L^{2}}+\tau^{k/2\alpha}\|\nabla^{k}U(\tau)\|_{L^{2}}\lesssim_{\overline{U},T}\|U_{0}\|_{L^{2}}. (3.1)

This well-defines the solution operator StU0=U(,t)S_{t}U_{0}=U(\cdot,t) which is generated by some closed operator 𝑨\boldsymbol{A} densely defined on some domain D(𝑨)D(\boldsymbol{A}), and we need to show that 𝑨=𝑻α,β\boldsymbol{A}=\boldsymbol{T}_{\!\alpha,\beta}.

D(𝑨)D(\boldsymbol{A}) contains the subspace of divergence-free Schwartz fields, and 𝑨=𝑻α,β\boldsymbol{A}=\boldsymbol{T}_{\!\alpha,\beta} for such fields. So, we are done if we show that D(𝑻α,β)D(𝑨).D(\boldsymbol{T}_{\!\alpha,\beta})\subset D(\boldsymbol{A}). Let VD(𝑻α,β)V\in D(\boldsymbol{T}_{\!\alpha,\beta}), λρ(𝑻α,β)\lambda\in\rho(\boldsymbol{T}_{\!\alpha,\beta}) and let VnV_{n} be a sequence of divergence-free Schwartz fields such that VnVV_{n}\to V in H2αH^{2\alpha}. Then (𝑻α,βλ)Vn(𝑻α,βλ)V(\boldsymbol{T}_{\!\alpha,\beta}-\lambda)V_{n}\to(\boldsymbol{T}_{\!\alpha,\beta}-\lambda)V in L2L^{2}. But 𝑨\boldsymbol{A} is closed and agrees with 𝑻α,β\boldsymbol{T}_{\!\alpha,\beta} along the sequence VnV_{n}, so VD(𝑨)V\in D(\boldsymbol{A}). ∎

For a closed, densely defined operator 𝑨\boldsymbol{A} that generates a semigroup eτ𝑨e^{\tau\boldsymbol{A}}, recall from [EN00] the following definitions:

ω0(𝑨)\displaystyle\omega_{0}(\boldsymbol{A}) infτ>01τlogeτ𝑨,\displaystyle\coloneq\inf_{\tau>0}\frac{1}{\tau}\log\big{\|}\textup{e}^{\tau\boldsymbol{A}}\big{\|},
ωess(𝑨)\displaystyle\omega_{\textup{ess}}(\boldsymbol{A}) infτ>01τloginf{eτ𝑨K:K is compact}.\displaystyle\coloneq\inf_{\tau>0}\frac{1}{\tau}\log\inf\big{\{}\big{\|}\textup{e}^{\tau\boldsymbol{A}}-K\big{\|}:K\text{ is compact}\big{\}}.
Lemma 3.4.

ωess(𝑻α,β)154α.\omega_{\textup{ess}}(\boldsymbol{T}_{\!\alpha,\beta})\leq 1-\frac{5}{4\alpha}.

Proof.

By the relative compactness proved in Lemma 3.2, we have

ωess(𝑻α,β)=ωess(𝑻α,0)ω0(𝑻α,0),\omega_{\textup{ess}}(\boldsymbol{T}_{\!\alpha,\beta})=\omega_{\textup{ess}}(\boldsymbol{T}_{\alpha,0})\leq\omega_{0}(\boldsymbol{T}_{\alpha,0}),

and ω0(𝑻α,0)logsup{Reλ:λσ(eτ𝑻α,0)}\omega_{0}(\boldsymbol{T}_{\alpha,0})\leq\log\sup\{\operatorname{Re}\lambda:\lambda\in\sigma(\textup{e}^{\tau\boldsymbol{T}_{\alpha,0}})\}, which is in turn upper bounded by 154α1-\frac{5}{4\alpha} by Lemma 3.1. ∎

From [EN00, Cor 2.11, p. 258], it follows that for all w>ωess(𝑻α,β)w>\omega_{\textup{ess}}(\boldsymbol{T}_{\!\alpha,\beta}), σ(𝑻α,β){Reλ>w}\sigma(\boldsymbol{T}_{\!\alpha,\beta})\cap\{\operatorname{Re}\lambda>w\} consists of only finitely many eigenvalues with finite multiplicity. In addition, for each δ>0\delta>0 and for each divergence-free U0L2U_{0}\in L^{2},

eτ𝑻α,βU0L2δeτ(a+δ)U0L2.\displaystyle\|\textup{e}^{\tau\boldsymbol{T}_{\!\alpha,\beta}}U_{0}\|_{L^{2}}\lesssim_{\delta}\textup{e}^{\tau(a+\delta)}\|U_{0}\|_{L^{2}}. (3.2)

We write λ=a+bi\lambda=a+b\textup{i}, a>0a>0 for the largest eigenvalue of 𝑻α,β\boldsymbol{T}_{\!\alpha,\beta}, with eigenfunction η\eta. It follows that

eτ𝑻α,βηL2=eτaηL2.\displaystyle\|\textup{e}^{\tau\boldsymbol{T}_{\!\alpha,\beta}}\eta\|_{L^{2}}=e^{\tau a}\|\eta\|_{L^{2}}. (3.3)
Lemma 3.5 (Parabolic regularity).

For all σσ0\sigma^{\prime}\geq\sigma\geq 0, δ>0\delta>0, and τ>0\tau>0,

eτ𝑻α,βU0Hσσ,σ,δeτ(a+δ)U0Hστ(σσ)/2α.\displaystyle\|\textup{e}^{\tau\boldsymbol{T}_{\!\alpha,\beta}}U_{0}\|_{H^{\sigma^{\prime}}}\lesssim_{\sigma,\sigma^{\prime},\delta}\frac{\textup{e}^{\tau(a+\delta)}\|U_{0}\|_{H^{\sigma}}}{\tau^{(\sigma^{\prime}-\sigma)/2\alpha}}. (3.4)
Proof.

For times τ1\tau\leq 1 say, and σ=0\sigma=0, these estimates are implied by the estimates in (3.1). The estimates for σ>0\sigma>0 are similarly proven by differentiating the equation for u(x,t)=1t11/2αU(xt1/2α,logt)u(x,t)=\frac{1}{t^{1-1/2\alpha}}U(xt^{-1/2\alpha},\log t).

For large times, we first use the small time estimate to drop to L2L^{2}:

eτ𝑻α,βU0Hσa,σ,δe(τ1)𝑻α,βU0L2.\|\textup{e}^{\tau\boldsymbol{T}_{\!\alpha,\beta}}U_{0}\|_{H^{\sigma^{\prime}}}\lesssim_{a,\sigma,\delta}\|\textup{e}^{(\tau-1)\boldsymbol{T}_{\!\alpha,\beta}}U_{0}\|_{L^{2}}.

It follows from (3.2) that eτ𝑻α,βU0Hσa,σ,δeτ(a+δ)U0L2\|\textup{e}^{\tau\boldsymbol{T}_{\!\alpha,\beta}}U_{0}\|_{H^{\sigma^{\prime}}}\lesssim_{a,\sigma,\delta}e^{\tau(a+\delta)}\|U_{0}\|_{L^{2}}, which proves the lemma. ∎

3.2 Nonuniqueness

For completeness, we repeat the argument laid out in [ABC22, §4] to construct the second solution of (1.1).

We look for a solution UU to (1.5) that vanishes as τ\tau\to-\infty with the ansatz U=βU¯+UL+UPU=\beta\overline{U}+U^{\textsf{L}}+U^{\textsf{P}}, where ULRe(eλτη)U^{\textsf{L}}\coloneq\operatorname{Re}(e^{\lambda\tau}\eta) solves the linear equation (τ𝑻α,β)UL=0(\partial_{\tau}-\boldsymbol{T}_{\!\alpha,\beta})U^{\textsf{L}}=0. Substituting this ansatz into (1.5), we find that the perturbation UPU^{\textsf{P}} satisfies the integral equation UP=𝒯(UP)U^{\textsf{P}}=\mathcal{T}(U^{\textsf{P}}), where

𝒯(UP)(τ)\displaystyle\mathcal{T}(U^{\textsf{P}})(\tau)\coloneq
τe(τs)𝑻α,β(ULUL+UPUL+ULUP+UPUP)ds.\displaystyle-\int\limits_{-\infty}^{\tau}\textup{e}^{(\tau-s)\boldsymbol{T}_{\!\alpha,\beta}}\mathbb{P}\Big{(}U^{\textsf{L}}\cdot\nabla U^{\textsf{L}}+U^{\textsf{P}}\cdot\nabla U^{\textsf{L}}+U^{\textsf{L}}\cdot\nabla U^{\textsf{P}}+U^{\textsf{P}}\cdot\nabla U^{\textsf{P}}\Big{)}\mathop{}\!\mathrm{d}s.

We construct UPU^{\textsf{P}} as a fixed point of 𝒯\mathcal{T} via the contraction mapping theorem. So, we set N>5/2N>5/2, ϵ0(0,a)\epsilon_{0}\in(0,a) and introduce the space

XX(N,T)C0((,T];HN(3;3))X\coloneq X(N,T)\coloneq C^{0}((-\infty,T];H^{N}(\mathbb{R}^{3};\mathbb{R}^{3}))

with the weighted norm (ϵ0>0\epsilon_{0}>0 to be chosen momentarily)

UXsupτTe(a+ϵ0)τU(τ)HN.\|U\|_{X}\coloneq\sup_{\tau\leq T}\textup{e}^{-(a+\epsilon_{0})\tau}\|U(\tau)\|_{H^{N}}.
Proposition 3.6.

Let BXB_{X} be the closed unit ball of XX. Then for TT sufficiently large and negative, and N>5/2N>5/2, 𝒯\mathcal{T} maps BXBXB_{X}\to B_{X} and is a contraction.

Proof.

Fix δ<a\delta<a and ϵ0<a\epsilon_{0}<a. 𝒯\mathcal{T} splits into three terms 𝒯(U)=(U,U)+U+𝒢\mathcal{T}(U)=\mathcal{B}(U,U)+\mathcal{L}U+\mathcal{G}, where

(U,V)\displaystyle-\mathcal{B}(U,V) =τe(τs)𝑻α,β(UV)ds,\displaystyle=\int_{-\infty}^{\tau}\textup{e}^{(\tau-s)\boldsymbol{T}_{\!\alpha,\beta}}\mathbb{P}(U\cdot\nabla V)\mathop{}\!\mathrm{d}s, (3.5)
U\displaystyle-\mathcal{L}U =τe(τs)𝑻α,β(ULU+UUL)ds,\displaystyle=\int_{-\infty}^{\tau}\textup{e}^{(\tau-s)\boldsymbol{T}_{\!\alpha,\beta}}\mathbb{P}(U^{\textsf{L}}\cdot\nabla U+U\cdot\nabla U^{\textsf{L}})\mathop{}\!\mathrm{d}s, (3.6)
𝒢\displaystyle-\mathcal{G} =τe(τs)𝑻α,β(ULUL)ds,\displaystyle=\int_{-\infty}^{\tau}\textup{e}^{(\tau-s)\boldsymbol{T}_{\!\alpha,\beta}}\mathbb{P}(U^{\textsf{L}}\cdot\nabla U^{\textsf{L}})\mathop{}\!\mathrm{d}s, (3.7)

Recall that HN1H^{N-1} for N>5/2N>5/2 is a Banach algebra. So, we have the estimate UVHN1UHNVHN\|U\cdot\nabla V\|_{H^{N-1}}\leq\|U\|_{H^{N}}\|V\|_{H^{N}}. Combining this estimate with Lemma 3.5 gives the following crude estimate

(U,U)(τ)HN+(α1/2)N,δτe(τs)(a+δ)(τs)(2α+1)/4αUHN2ds\displaystyle\|\mathcal{B}(U,U)(\tau)\|_{H^{N+(\alpha-1/2)}}\lesssim_{N,\delta}\int_{-\infty}^{\tau}\frac{\textup{e}^{(\tau-s)(a+\delta)}}{(\tau-s)^{(2\alpha+1)/4\alpha}}\|U\|_{H^{N}}^{2}\mathop{}\!\mathrm{d}s
e2τ(a+ϵ0)UX2τe(τs)(δa2ϵ0)(τs)(2α+1)/4αdsϵ0,δe2τ(a+ϵ0)UX2,\displaystyle\leq\textup{e}^{2\tau(a+\epsilon_{0})}\|U\|^{2}_{X}\int_{-\infty}^{\tau}\frac{\textup{e}^{(\tau-s)(\delta-a-2\epsilon_{0})}}{(\tau-s)^{(2\alpha+1)/4\alpha}}\mathop{}\!\mathrm{d}s\lesssim_{\epsilon_{0},\delta}\textup{e}^{2\tau(a+\epsilon_{0})}\|U\|_{X}^{2}, (3.8)

as long as δ<a+2ϵ0\delta<a+2\epsilon_{0}. Hence (U,U)XeT(a+ϵ0)UX2\|\mathcal{B}(U,U)\|_{X}\lesssim e^{T(a+\epsilon_{0})}\|U\|_{X}^{2}.

By replacing the estimate of UU in Lemma 3.5 by the estimate (3.3) of ULU^{\textsf{L}}, we similarly obtain (using δ<a\delta<a) the estimates for all N>5/2N>5/2,

U(τ)HNN,ϵ0,δeτ(2a+ϵ0)UX,𝒢(τ)HNN,ϵ0,δe2τa,\displaystyle\|\mathcal{L}U(\tau)\|_{H^{N}}\lesssim_{N,\epsilon_{0},\delta}\textup{e}^{\tau(2a+\epsilon_{0})}\|U\|_{X},\quad\|\mathcal{G}(\tau)\|_{H^{N}}\lesssim_{N,\epsilon_{0},\delta}\textup{e}^{2\tau a}, (3.9)

which lead to the estimates

UXN,ϵ0,δeTaUX,𝒢XN,ϵ0,δeT(aϵ0).\displaystyle\|\mathcal{L}U\|_{X}\lesssim_{N,\epsilon_{0},\delta}\textup{e}^{Ta}\|U\|_{X},\qquad\|\mathcal{G}\|_{X}\lesssim_{N,\epsilon_{0},\delta}\textup{e}^{T(a-\epsilon_{0})}.

It follows that by choosing TT sufficiently large and negative, we can make \|\mathcal{B}\|, \|\mathcal{L}\| and 𝒢X\|\mathcal{G}\|_{X} as small as we wish. It follows that for TT sufficiently large and negative, 𝒯(U)X1\|\mathcal{T}(U)\|_{X}\leq 1 if UX1\|U\|_{X}\leq 1. In addition, since

𝒯(U)𝒯(V)X(2+)UVX,\|\mathcal{T}(U)-\mathcal{T}(V)\|_{X}\leq(2\|\mathcal{B}\|+\|\mathcal{L}\|)\|U-V\|_{X},

Choosing T1T\ll-1 also makes 𝒯|BX\mathcal{T}|_{B_{X}} contractive. This finishes the proof. ∎

Let UPX(N,T)U^{\textsf{P}}\in X(N,T) be the unique fixed point of 𝒯\mathcal{T} guaranteed by Proposition 3.6. The estimates (3.8) and (3.9) let us bootstrap to CC^{\infty} regularity in space, and shows that UPU^{\textsf{P}} decays as e2τae^{2\tau a} as τ\tau\to-\infty. By construction, U=βU¯+UL+UPU=\beta\overline{U}+U^{\textsf{L}}+U^{\textsf{P}} solves (1.5). Since |UL|eτa|U^{\textsf{L}}|\gtrsim e^{\tau a} as τ\tau\to-\infty, UPULU^{\textsf{P}}\neq-U^{\textsf{L}}, so UβU¯U\neq\beta\overline{U}. Finally, undoing the similarity variable transform for both UU and βU¯\beta\overline{U} gives a pair of distinct solutions of (1.1).

4 Acknowledgements

We thank the anonymous referee and the associated editor for their invaluable comments which helped to improve the paper. This work was supported by the National Key Research and Development Program of China, No.2020YFA0712900, No 2022YFA1006700 and NSFC Grant 12071043.

References

  • [ABC22] Dallas Albritton, Elia Brué and Maria Colombo “Non-uniqueness of Leray solutions of the forced Navier-Stokes equations” In Ann. Math. (2) 196.1, 2022, pp. 415–455 DOI: 10.4007/annals.2022.196.1.3
  • [ABC22a] Dallas Albritton, Elia Bruè and Maria Colombo “Gluing non-unique Navier-Stokes solutions” arXiv, 2022 DOI: 10.48550/ARXIV.2209.03530
  • [AC22] Dallas Albritton and Maria Colombo “Non-uniqueness of Leray solutions to the hypodissipative Navier-Stokes equations in two dimensions” arXiv, 2022 DOI: 10.48550/ARXIV.2209.08713
  • [EN00] Klaus-Jochen Engel and Rainer Nagel “One-parameter semigroups for linear evolution equations” 194, Grad. Texts Math. Berlin: Springer, 2000 DOI: 10.1007/b97696
  • [JS15] Hao Jia and Vladimir Sverak “Are the incompressible 3d Navier-Stokes equations locally ill-posed in the natural energy space?” In J. Funct. Anal. 268.12, 2015, pp. 3734–3766 DOI: 10.1016/j.jfa.2015.04.006
  • [Lio69] J. L. Lions “Quelques méthodes de résolution des problèmes aux limites non linéaires.” In Etudes mathematiques XX, 1969
  • [LR22] P. G. Lemarié-Rieusset “Forces for the Navier-Stokes equations and the Koch and Tataru theorem” arXiv, 2022 DOI: 10.48550/ARXIV.2209.12196
  • [LT20] Tianwen Luo and Edriss S. Titi “Non-uniqueness of weak solutions to hyperviscous Navier-Stokes equations: on sharpness of J.-L. Lions exponent” Id/No 92 In Calc. Var. Partial Differ. Equ. 59.3, 2020, pp. 15 DOI: 10.1007/s00526-020-01742-4
  • [MYZ08] Changxing Miao, Baoquan Yuan and Bo Zhang “Well-posedness of the Cauchy problem for the fractional power dissipative equations” In Nonlinear Anal., Theory Methods Appl., Ser. A, Theory Methods 68.3, 2008, pp. 461–484 DOI: 10.1016/j.na.2006.11.011
  • [Vis18] Misha Vishik “Instability and non-uniqueness in the Cauchy problem for the Euler equations of an ideal incompressible fluid. Part I” arXiv, 2018 DOI: 10.48550/ARXIV.1805.09426
  • [Vis18a] Misha Vishik “Instability and non-uniqueness in the Cauchy problem for the Euler equations of an ideal incompressible fluid. Part II” arXiv, 2018 DOI: 10.48550/ARXIV.1805.09440