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

Finite-time blowup for a Navier–Stokes model equation for the self-amplification of strain

Evan Miller University of British Columbia, Department of Mathematics emiller@msri.org
Abstract

In this paper, we consider a model equation for the Navier–Stokes strain equation. This model equation has the same identity for enstrophy growth and a number of the same regularity criteria as the full Navier–Stokes strain equation, and is also an evolution equation on the same constraint space. We prove finite-time blowup for this model equation, which shows that the identity for enstrophy growth and the strain constraint space are not sufficient on their own to guarantee global regularity for Navier–Stokes. The mechanism for the finite-time blowup of this model equation is the self-amplification of strain, which is consistent with recent research suggesting that strain self-amplification, not vortex stretching, is the main mechanism behind the turbulent energy cascade. Because the strain self-amplification model equation is obtained by dropping certain terms from the full Navier–Stokes strain equation, we will also prove a conditional blowup result for the full Navier–Stokes equation involving a perturbative condition on the terms neglected in the model equation.

1 Introduction

The incompressible Navier–Stokes equation is one of the fundamental equations of fluid mechanics. Although it is over 150 years old, much about its solutions, including the global existence of smooth solutions, remains unknown. The Navier–Stokes equation is given by

tuΔu+(u)u+p\displaystyle\partial_{t}u-\Delta u+(u\cdot\nabla)u+\nabla p =0\displaystyle=0 (1.1)
u\displaystyle\nabla\cdot u =0,\displaystyle=0, (1.2)

where u3u\in\mathbb{R}^{3} is the velocity and pp is the pressure. The first equation is a statement of Newton’s second law, F=maF=ma, where tu+(u)u\partial_{t}u+(u\cdot\nabla)u gives the acceleration in the Lagrangian frame, Δu\Delta u describes the viscous forces due to the internal friction of the fluid, and p-\nabla p describes the force due to the pressure. The second equation, the divergence free constraint, comes from the conservation of mass. We will note that pp is not an independently evolving function, but is determined entirely by uu by convolution with the Poisson kernel,

p=(Δ)1i,j=13ujxiuixj.p=(-\Delta)^{-1}\sum_{i,j=1}^{3}\frac{\partial u_{j}}{\partial x_{i}}\frac{\partial u_{i}}{\partial x_{j}}. (1.3)

It is possible to state the incompressible Navier–Stokes equation without giving any reference to pressure at all by making use of the Helmholtz projection onto the space of divergence free vector fields, yielding the equation

tuΔu+Pdf(uu)=0.\partial_{t}u-\Delta u+P_{df}\nabla\cdot\left(u\otimes u\right)=0. (1.4)

Note that we have used the fact that (uu)=(u)u,\nabla\cdot(u\otimes u)=(u\cdot\nabla)u, because u=0\nabla\cdot u=0, and the fact that the Helmholtz decomposition implies that Pdf(p)=0P_{df}(\nabla p)=0.

The first major advances towards a rigorous mathematical understanding of the Navier–Stokes equation came in the seminal paper by Leray [Leray]. For all initial data u0Ldf2,u^{0}\in L^{2}_{df}, Leray proved the global-in-time existence of weak solutions, in the sense of integrating against smooth test functions, satisfying the energy inequality, which states that for all t>0t>0,

12u(t)L22+0tu(τ)H˙12dτ12u0L22.\frac{1}{2}\|u(t)\|_{L^{2}}^{2}+\int_{0}^{t}\|u(\tau)\|_{\dot{H}^{1}}^{2}\mathop{}\!\mathrm{d}\tau\leq\frac{1}{2}\left\|u^{0}\right\|_{L^{2}}^{2}. (1.5)

Unfortunately, while such solutions are well suited to study in that the sense that global-in-time existence is guaranteed for all finite-energy initial data, they are not known to be either smooth or unique, leaving major problems for the well-posedness theory.

The lack of a uniqueness and regularity theory for Leray weak solutions led Fujita and Kato to develop the notion of mild solutions, which satisfy (1.4) in the sense of Duhamel’s formula. Unlike Leray’s weak solutions, mild solutions must be both smooth and unique. Kato and Fujita proved the local-in-time existence, uniqueness, and smoothness of mild solutions for initial data in H˙1\dot{H}^{1}, with the time of existence bounded below uniformly in the H˙1\dot{H}^{1} norm [KatoFujita].

Theorem 1.1.

There exists an absolute constant C>0,C>0, such that for all initial data u0H˙df1u^{0}\in\dot{H}^{1}_{df}, there exists TmaxCu0H˙14T_{max}\geq\frac{C}{\left\|u^{0}\right\|_{\dot{H}^{1}}^{4}} and a unique mild solution to the Navier–Stokes equation uC([0,Tmax);H˙df1)u\in C\left([0,T_{max});\dot{H}^{1}_{df}\right). Furthermore, we have the higher regularity uC((0,Tmax)×3).u\in C^{\infty}\left((0,T_{max})\times\mathbb{R}^{3}\right). If in addition we have u0Hdf1,u^{0}\in H^{1}_{df}, then the energy inequality holds with equality, that is for all 0<t<Tmax0<t<T_{max},

12u(t)L22+0tu(τ)H˙12dτ=12u0L22.\frac{1}{2}\|u(t)\|_{L^{2}}^{2}+\int_{0}^{t}\|u(\tau)\|_{\dot{H}^{1}}^{2}\mathop{}\!\mathrm{d}\tau=\frac{1}{2}\left\|u^{0}\right\|_{L^{2}}^{2}. (1.6)

We will note that because mild solutions are smooth and unique, the initial value problem for mild solutions of the Navier–Stokes equation is locally well-posed in H˙1\dot{H}^{1}—and also in a number of larger spaces; however, it is not known to be globally well-posed. Whether the Navier–Stokes equation has global smooth solutions or admits smooth solutions that blowup in finite-time is one of the biggest open problems in PDEs and one of the “Millennium Problems” put forward by the Clay Mathematics Institute [Clay].

The main difficulty is that the only bounds that are available on the growth of solutions are the bounds in LtLx2L^{\infty}_{t}L^{2}_{x} and Lt2H˙x1L^{2}_{t}\dot{H}^{1}_{x} due to the energy equality, and these bounds are not enough to guarantee the global existence of smooth solutions because the energy equality is super-critical with respect to the invariant rescaling of the Navier–Stokes equation. The solution set of the Navier–Stokes equation is preserved under the rescaling,

uλ(x,t)=λu(λx,λ2t),u^{\lambda}(x,t)=\lambda u(\lambda x,\lambda^{2}t), (1.7)

for all λ>0.\lambda>0. This means that is not enough to control the LtLx2L^{\infty}_{t}L^{2}_{x} or Lt2H˙x1L^{2}_{t}\dot{H}^{1}_{x} norms of u,u, which are supercritical in terms of scaling; in order to guarantee global regularity, we need to control a scale critical norm. Ladyzhenskaya [Ladyzhenskaya], Prodi [Prodi], and Serrin [Serrin] independently proved a family of scale critical regularity criteria, which state that if Tmax<+,T_{max}<+\infty, and 2p+3q=1,\frac{2}{p}+\frac{3}{q}=1, with 3<q+3<q\leq+\infty then

0Tmaxu(t)Lqpdt=+.\int_{0}^{T_{max}}\|u(t)\|_{L^{q}}^{p}\mathop{}\!\mathrm{d}t=+\infty. (1.8)

Escauriaza, Seregin and Sv̌erák [L3] extended this result to the endpoint case q=3q=3. They proved that if Tmax<+,T_{max}<+\infty, then

lim suptTmaxu(t)L3=+.\limsup_{t\to T_{max}}\|u(t)\|_{L^{3}}=+\infty. (1.9)

Recently, Tao further extended this regularity criterion giving a quantitative lower bound on the rate of the blowup of the L3L^{3} norm [TaoL3]. This result is very slightly supercritical—in fact triple logarithmically—with respect to scaling, and is the first supercritical regularity criterion for the Navier–Stokes equation.

Two crucially important objects for the study of the Navier–Stokes equation are the strain, which is the symmetric gradient of the velocity, S=symu,S=\nabla_{sym}u, with Sij=12(iuj+jui),S_{ij}=\frac{1}{2}\left(\partial_{i}u_{j}+\partial_{j}u_{i}\right), and the vorticity, which is a vector that represents the anti-symmetric part of the velocity, and is given by ω=×u.\omega=\nabla\times u. Physically, the strain describes how a parcel of the fluid is deformed, while the vorticity describes how a parcel of the fluid is rotated.

Taking the curl of (1.1), we find the evolution equation for ω\omega is given by

tωΔω+(u)ωSω=0.\partial_{t}\omega-\Delta\omega+(u\cdot\nabla)\omega-S\omega=0. (1.10)

Taking the symmetric gradient of (1.1), we find the evolution equation for SS is given by,

tSΔS+(u)S+S2+14ωω14|ω|2I3+Hess(p)=0.\partial_{t}S-\Delta S+(u\cdot\nabla)S+S^{2}+\frac{1}{4}\omega\otimes\omega-\frac{1}{4}|\omega|^{2}I_{3}+\operatorname*{Hess}(p)=0. (1.11)

We will note that the vorticity equation is invariant under the rescaling,

ωλ(x,t)=λ2ω(λx,λ2t),\omega^{\lambda}(x,t)=\lambda^{2}\omega(\lambda x,\lambda^{2}t), (1.12)

and the strain equation is invariant under the rescaling

Sλ(x,t)=λ2S(λx,λ2t).S^{\lambda}(x,t)=\lambda^{2}S(\lambda x,\lambda^{2}t). (1.13)

The extra factor of λ\lambda comes from the fact that both ω\omega and SS scale like u.\nabla u.

The vorticity has been studied fairly exhaustively for its role in the dynamics of the Navier–Stokes equation. For instance the Beale-Kato-Majda regularity criterion [BKM], which holds for smooth solutions of both the Euler and Navier–Stokes equations, states that if Tmax<+,T_{max}<+\infty, then

0Tmaxω(,t)Ldt=+.\int_{0}^{T_{max}}\|\omega(\cdot,t)\|_{L^{\infty}}\mathop{}\!\mathrm{d}t=+\infty. (1.14)

Chae and Choe proved a regularity criterion on two components of vorticity that has a geometric significance, guaranteeing that blowup must be fully three dimensional [ChaeVort]. They showed that if a smooth solution of the Navier–Stokes equation blows up in finite-time Tmax<+T_{max}<+\infty, then for all 32<q<+,2p+3q=2\frac{3}{2}<q<+\infty,\frac{2}{p}+\frac{3}{q}=2,

0Tmaxe3×ω(,t)Lqpdt=+.\int_{0}^{T_{max}}\|e_{3}\times\omega(\cdot,t)\|_{L^{q}}^{p}\mathop{}\!\mathrm{d}t=+\infty. (1.15)

The fixed direction condition in this regularity criterion was recently loosened by the author in [MillerVort]. In another key result involving vorticity, Constantin and Fefferman proved that the direction of the vorticity must vary rapidly in regions where the vorticity is large if there is finite-time blowup [VortDirection]. There are many other results involving vorticity, far too many to list here.

The strain equation has been investigated much less thoroughly, but can provide some insights that do not follow as clearly from the vorticity equation. We will refer to the evolution equation for SS in (1.11) as the Navier–Stokes strain equation. This equation is an evolution equation on the constraint space Lst2,L^{2}_{st}, the space of strain matrices, which replaces the divergence free constraint for the Navier–Stokes and vorticity equations. We define Lst2L^{2}_{st} as follows.

Definition 1.2.

Define Lst2L2(3;𝕊3×3)L^{2}_{st}\subset L^{2}\left(\mathbb{R}^{3};\mathbb{S}^{3\times 3}\right) by

Lst2={symu:uH˙1,u=0}.L^{2}_{st}=\left\{\nabla_{sym}u:u\in\dot{H}^{1},\nabla\cdot u=0\right\}. (1.16)

The role of this constraint space in the evolution equation (1.11) was examined by the author in [MillerStrain]. One geometric restriction on the matrices SLst2S\in L^{2}_{st} is that they must be trace free, because

tr(S)=u=0.\operatorname*{tr}(S)=\nabla\cdot u=0. (1.17)

Furthermore, in that paper, the author proved that Hessians and scalar multiples of the identity matrix must be in the orthogonal compliment of Lst2.L^{2}_{st}.

Proposition 1.3.

For all fH˙2(3),f\in\dot{H}^{2}\left(\mathbb{R}^{3}\right), and for all gL2(3),g\in L^{2}\left(\mathbb{R}^{3}\right), we have Hess(f),gI3(Lst2)\operatorname*{Hess}(f),gI_{3}\in\left(L^{2}_{st}\right)^{\perp}. That is for all SLst2,S\in L^{2}_{st},

Hess(f),S\displaystyle\left<\operatorname*{Hess}(f),S\right> =0,\displaystyle=0, (1.18)
gI3,S\displaystyle\left<gI_{3},S\right> =0.\displaystyle=0. (1.19)

For sufficiently smooth solutions to the Navier–Stokes strain equation 14|ω|2,Hess(p)L2,\frac{1}{4}|\omega|^{2},\operatorname*{Hess}(p)\in L^{2}, so we can conclude that the terms 14|ω|2I3\frac{1}{4}|\omega|^{2}I_{3} and Hess(p)\operatorname*{Hess}(p) are orthogonal to the constraint space, 14|ω|2I3,Hess(p)(Lst2).\frac{1}{4}|\omega|^{2}I_{3},\operatorname*{Hess}(p)\in\left(L^{2}_{st}\right)^{\perp}. This means that the Navier–Stokes strain equation can be expressed in terms of the projection onto Lst2L^{2}_{st} as

tSΔS+Pst((u)S+S2+14ωω)=0.\partial_{t}S-\Delta S+P_{st}\left((u\cdot\nabla)S+S^{2}+\frac{1}{4}\omega\otimes\omega\right)=0. (1.20)

This is analogous to defining the Navier–Stokes equation without any reference to p\nabla p by using the Helmholtz projection onto the space of divergence free vector fields in (1.4). We will use (1.20) to define mild solutions to the Navier–Stokes strain equation in section 3.

It is not actually necessary to separately prove the existence of mild solutions to the strain equation, as it is straightforward to reduce this problem to the existence of mild solutions of the Navier–Stokes equation. The author proved the equivalence of these formulations in [MillerStrain].

Proposition 1.4.

A velocity field uC([0,Tmax);H˙df1)u\in C\left([0,T_{max});\dot{H}^{1}_{df}\right) is a mild solution of the Navier–Stokes equation, if and only if SC([0,Tmax);Lst2)S\in C\left([0,T_{max});L^{2}_{st}\right) is a mild solution to the Navier–Stokes strain equation, where S=symu,u=2div(Δ)1SS=\nabla_{sym}u,u=-2\operatorname*{div}(-\Delta)^{-1}S

The strain evolution equation is extremely useful, because it allows us to prove a simplified identity for enstrophy growth, which can equivalently be defined in terms of the square of L2L^{2} norm of S,ω,S,\omega, or u\nabla u based on an isometry proven by the author in [MillerStrain].

Proposition 1.5.

For all 32<α<32,-\frac{3}{2}<\alpha<\frac{3}{2}, and for all SH˙stαS\in\dot{H}^{\alpha}_{st}

SH˙α2=12ωH˙α2=12uH˙α2\|S\|_{\dot{H}^{\alpha}}^{2}=\frac{1}{2}\|\omega\|_{\dot{H}^{\alpha}}^{2}=\frac{1}{2}\|\nabla u\|_{\dot{H}^{\alpha}}^{2} (1.21)
Remark 1.6.

We should note here that the factor of 12\frac{1}{2} in Proposition 1.5 is entirely an artifact of how the vorticity is defined. The vorticity is a vector representation of the anti-symmetric part of u,\nabla u, with

A=12(0ω3ω2ω30ω1ω2ω10),A=\frac{1}{2}\left(\begin{matrix}0&\omega_{3}&-\omega_{2}\\ -\omega_{3}&0&\omega_{1}\\ \omega_{2}&-\omega_{1}&0\\ \end{matrix}\right), (1.22)

where AA is the anti-symmetric part of u\nabla u given by Aij=12(iujjui).A_{ij}=\frac{1}{2}\left(\partial_{i}u_{j}-\partial_{j}u_{i}\right). From this identity we can see that

SH˙α2=AH˙α2,\|S\|_{\dot{H}^{\alpha}}^{2}=\|A\|_{\dot{H}^{\alpha}}^{2}, (1.23)

so the isometry in Proposition 1.5 tells us that all the Hilbert norms of the symmetric and anti-symmetric parts of the gradient of a divergence free vector field are equal.

Definition 1.7.

Based on the isometry in Proposition 1.5, we will define the enstrophy of a solution to the Navier–Stokes equation, which can be equivalently expressed by

E(t)\displaystyle E(t) =S(t)L22\displaystyle=\|S(t)\|_{L^{2}}^{2} (1.24)
=12ω(t)L22\displaystyle=\frac{1}{2}\|\omega(t)\|_{L^{2}}^{2} (1.25)
=12u(t)L22,\displaystyle=\frac{1}{2}\|\nabla u(t)\|_{L^{2}}^{2}, (1.26)

and the energy of a solution of the Navier–Stokes equation, which can be equivalently expressed by

K(t)\displaystyle K(t) =S(t)H˙12\displaystyle=\|S(t)\|_{\dot{H}^{-1}}^{2} (1.27)
=12ω(t)H˙12\displaystyle=\frac{1}{2}\|\omega(t)\|_{\dot{H}^{-1}}^{2} (1.28)
=12u(t)L22,\displaystyle=\frac{1}{2}\|u(t)\|_{L^{2}}^{2}, (1.29)
Remark 1.8.

The energy equality for smooth solutions of the Navier–Stokes equation can be stated in terms of energy and enstrophy as

K(t)+20tE(τ)dτ=K0K(t)+2\int_{0}^{t}E(\tau)\mathop{}\!\mathrm{d}\tau=K_{0} (1.30)

Enstrophy is a very important quantity because Theorem 1.1 states that a smooth solution of the Navier–Stokes equation must exist locally in time for initial data in u0H˙1.u^{0}\in\dot{H}^{1}. This implies that enstrophy controls regularity, because as long as enstrophy remains bounded on some time interval, a smooth solution can be continued to some later time.

The standard estimate for enstrophy growth is given in terms of nonlocal interaction of the vorticity and the strain:

ddt12ω(t)L22=ωH˙12+S,ωω.\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\frac{1}{2}\|\omega(t)\|_{L^{2}}^{2}=-\|\omega\|_{\dot{H}^{1}}^{2}+\left<S,\omega\otimes\omega\right>. (1.31)

This is a nonlocal identity because SS can be determined in terms of ω\omega by a nonlocal, zeroth order pseudo-differential operator, with S=sym×(Δ)1ω.S=\nabla_{sym}\nabla\times(-\Delta)^{-1}\omega. Using the isometry in Proposition 1.5, and the evolution equations for both the strain and the vorticity, this identity can be drastically simplified, with the nonlocal term replaced by a term involving only the determinant of S.S.

Proposition 1.9.

Suppose uC([0,Tmax);H˙df1)u\in C\left([0,T_{max});\dot{H}^{1}_{df}\right) is a mild solution of the Navier–Stokes equation. Note that this is equivalent to assuming that SC([0,Tmax);Lst2)S\in C\left([0,T_{max});L^{2}_{st}\right) is a mild solution to the Navier–Stokes strain equation. Then for all 0<t<Tmax,0<t<T_{max},

ddtS(t)L22\displaystyle\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\|S(t)\|_{L^{2}}^{2} =2SH˙1243tr(S3)\displaystyle=-2\|S\|_{\dot{H}^{1}}^{2}-\frac{4}{3}\int\operatorname*{tr}\left(S^{3}\right) (1.32)
=2SH˙124det(S).\displaystyle=-2\|S\|_{\dot{H}^{1}}^{2}-4\int\det(S). (1.33)

This identity was first proven by Neustupa and Penel in [NeustupaPenel1, NeustupaPenel2]. The analogous result without the dissipation term 2SH˙12-2\|S\|_{\dot{H}^{1}}^{2} was later proven independently by Chae in the context of smooth solutions of the Euler equation in [ChaeStrain] using similar methods to Neustupa and Penel. This identity was also proven using the evolution equation for the strain, a different approach to that of Neustupa and Penel, by the author in [MillerStrain]. The identity in Proposition 1.9 directly implies a family of scale-invariant regularity criteria in terms of the positive part of the middle eigenvalue of S.S.

Theorem 1.10.

Suppose uC([0,Tmax);H˙df1)u\in C\left([0,T_{max});\dot{H}^{1}_{df}\right) is a mild solution of the Navier–Stokes equation, or equivalently that SC([0,Tmax);Lst2)S\in C\left([0,T_{max});L^{2}_{st}\right) is a mild solution to the Navier–Stokes strain equation. Let λ1(x,t)λ2(x,t)λ3(x,t)\lambda_{1}(x,t)\leq\lambda_{2}(x,t)\leq\lambda_{3}(x,t) be the eigenvalues of S(x,t),S(x,t), and let λ2+(x,t)=max{0,λ2(x,t)}.\lambda^{+}_{2}(x,t)=\max\left\{0,\lambda_{2}(x,t)\right\}. Then for all 3q+2p=2,32<q+,\frac{3}{q}+\frac{2}{p}=2,\frac{3}{2}<q\leq+\infty, there exists Cq>0C_{q}>0 depending only on qq such that for all 0<t<Tmax0<t<T_{max}

S(t)L22S0L22exp(Cq0tλ2+(τ)Lqpdτ).\|S(t)\|_{L^{2}}^{2}\leq\left\|S^{0}\right\|_{L^{2}}^{2}\exp\left(C_{q}\int_{0}^{t}\left\|\lambda_{2}^{+}(\tau)\right\|_{L^{q}}^{p}\mathop{}\!\mathrm{d}\tau\right). (1.34)

In particular, if Tmax<+,T_{max}<+\infty, then

0Tmaxλ2+(t)Lqpdt=+.\int_{0}^{T_{max}}\left\|\lambda_{2}^{+}(t)\right\|_{L^{q}}^{p}\mathop{}\!\mathrm{d}t=+\infty. (1.35)

This regularity criterion was first proven by Neustupa and Penel in [NeustupaPenel1, NeustupaPenel2, NeustupaPenel3]. It was also proven independently by the author in [MillerStrain]. Note that because tr(S)=0,\operatorname*{tr}(S)=0, this regularity criterion significantly restricts the geometry of any finite-time blowup for the Navier–Stokes equation: any blowup must be driven by unbounded planar stretching and axial compression, with the strain having two positive eigenvalues, and one very negative eigenvalue.

There are many other conditional regularity results, which guarantee the regularity of solutions as long as some scale critical quantity remains finite, including regularity criteria involving the derivative in just one direction 3u\partial_{3}u [Kukavica], and involving just one velocity direction u3u_{3} [CheminOneDirection, CheminOneDirection2]. For a more thorough, but by no means exhaustive, treatment of regularity criteria for the Navier–Stokes equation see Chapter 11 in [NS21].

In this paper, we will take the opposite approach. We will prove finite-time blowup for solutions of the Navier–Stokes equation with a fairly broad set of initial data, assuming that a certain scale invariant quantity related to the structure of the nonlinearity remains small. We will do this first by considering a model equation for the Navier–Stokes strain equation and proving finite-time blowup for solutions of this this model equation, and then by viewing the actual Navier–Stokes strain equation as a perturbation of the model equation.

In order to do this, we will drop the advection and the vorticity terms from the evolution equation (1.20) entirely, along with a piece of the S2S^{2} term so that the enstrophy growth identity in Proposition 1.9 still holds. We will show that

S,ωω=4det(S)=43S2,S,\left<S,\omega\otimes\omega\right>=-4\int\det(S)=-\frac{4}{3}\left<S^{2},S\right>, (1.36)

and

(u)S,S=0,\left<(u\cdot\nabla)S,S\right>=0, (1.37)

and therefore

Pst((u)S+13S2+14ωω),S=0.\left<P_{st}\left((u\cdot\nabla)S+\frac{1}{3}S^{2}+\frac{1}{4}\omega\otimes\omega\right),S\right>=0. (1.38)

Using this identity, we can rewrite the full Navier–Stokes strain equation as

tSΔS+23Pst(S2)+Pst((u)S+13S2+14ωω)=0,\partial_{t}S-\Delta S+\frac{2}{3}P_{st}\left(S^{2}\right)+P_{st}\left((u\cdot\nabla)S+\frac{1}{3}S^{2}+\frac{1}{4}\omega\otimes\omega\right)=0, (1.39)

Dropping the term Pst((u)S+13S2+14ωω)P_{st}\left((u\cdot\nabla)S+\frac{1}{3}S^{2}+\frac{1}{4}\omega\otimes\omega\right) from the evolution equation, our strain model equation will be given by

tSΔS+23Pst(S2)=0.\partial_{t}S-\Delta S+\frac{2}{3}P_{st}\left(S^{2}\right)=0. (1.40)

We will refer to (1.40) as the strain self-amplification model equation, because it isolates the interaction of the strain with itself, discarding the nonlocal interaction with the vorticity and the effects of advection. In the model equation, we are dropping a combination of terms that are orthogonal to SS in L2,L^{2}, while keeping the two terms that contribute to the evolution in time of the L2L^{2} norm to first order. We will also show that for solutions of the strain self-amplification model equation

ddtS(t)L22=2SH˙124det(S),\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\|S(t)\|_{L^{2}}^{2}=-2\|S\|_{\dot{H}^{1}}^{2}-4\int\det(S), (1.41)

so the strain self-amplification model equation does in fact have the same identity for enstrophy growth as the Navier–Stokes equation, and consequently has a regularity criterion for λ2+\lambda_{2}^{+} in the critical Lebesgue spaces LtpLxqL^{p}_{t}L^{q}_{x} entirely analogous to the regularity criterion for the Navier–Stokes equation in Theorem 1.10.

Solutions of this model equation blowup in finite-time for a fairly wide range of initial conditions.

Theorem 1.11.

Suppose SC([0,Tmax);Hst1)S\in C\left([0,T_{max});H^{1}_{st}\right) is a mild solution of the strain self-amplification model equation, such that

3S0H˙124det(S0)>0.-3\left\|S^{0}\right\|_{\dot{H}^{1}}^{2}-4\int\det\left(S^{0}\right)>0. (1.42)

Then for all 0<t<Tmax0<t<T_{max},

E(t)>E0(1r0t)2,E(t)>\frac{E_{0}}{(1-r_{0}t)^{2}}, (1.43)

where

r0=3S0H˙124det(S0)2S0L22.r_{0}=\frac{-3\left\|S^{0}\right\|_{\dot{H}^{1}}^{2}-4\int\det\left(S^{0}\right)}{2\left\|S^{0}\right\|_{L^{2}}^{2}}. (1.44)

Note in particular that this implies

Tmax2S0L223S0L224det(S0).T_{max}\leq\frac{2\left\|S^{0}\right\|_{L^{2}}^{2}}{-3\left\|S^{0}\right\|_{L^{2}}^{2}-4\int\det\left(S^{0}\right)}. (1.45)

Furthermore, for all 2p+3q=2,32<q+,\frac{2}{p}+\frac{3}{q}=2,\frac{3}{2}<q\leq+\infty,

0Tmaxλ2+(t)Lqpdt=+.\int_{0}^{T_{max}}\|\lambda_{2}^{+}(t)\|_{L^{q}}^{p}\mathop{}\!\mathrm{d}t=+\infty. (1.46)
Remark 1.12.

The key to the proof of Theorem 1.11, which is the main result of this paper, is a 32\frac{3}{2} lower bound on the rate of enstrophy growth for a wide range of initial conditions. In particular, we will show that if

g0:=3S0H˙124det(S0)S0L23>0,g_{0}:=\frac{-3\left\|S^{0}\right\|_{\dot{H}^{1}}^{2}-4\int\det\left(S^{0}\right)}{\left\|S^{0}\right\|_{L^{2}}^{3}}>0, (1.47)

then for all 0<t<Tmax0<t<T_{max},

ddtE(t)>g0E(t)32,\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}E(t)>g_{0}E(t)^{\frac{3}{2}}, (1.48)

which immediately yields the estimate (1.43) in Theorem 1.11.

Remark 1.13.

Theorem 1.11 shows that the regularity criterion in Theorem 1.10, which guarantees the existence of smooth solutions of the Navier–Stokes equation so long as λ2+LtpLxq,\lambda_{2}^{+}\in L_{t}^{p}L_{x}^{q}, is not enough to guarantee the global existence of smooth solutions to the Navier–Stokes equation just by making use of the constraint space. For solutions of the strain self-amplification model equation, which is an evolution equation on the constraint space Lst2,L^{2}_{st}, λ2+\lambda_{2}^{+} becomes unbounded in this whole family of scale critical spaces. The regularity criterion on λ2+\lambda_{2}^{+} implies that blowup for the Navier–Stokes equation must be characterized by unbounded planar stretching and axial compression, corresponding to a strain matrix with two positive eigenvalues and one very negative eigenvalue in turbulent regions. One physical example of such a structure in turbulent fluids is two colliding jets. The blowup result for the strain self-amplification model equation shows that blowup with these features is possible within the relevant constraint space.

Because we chose our strain self-amplification model equation (1.40) by dropping some terms from the full strain equation, we can prove a new conditional blowup result for the full Navier–Stokes equation, by viewing the actual strain equation as a perturbation of the strain self-amplification model equation.

Theorem 1.14.

Suppose uC([0,Tmax);Hdf2)u\in C\left([0,T_{max});H^{2}_{df}\right) is a mild solution of the Navier–Stokes equation such that

f0:=3S0H˙124det(S0)>0,f_{0}:=-3\left\|S^{0}\right\|_{\dot{H}^{1}}^{2}-4\int\det\left(S^{0}\right)>0, (1.49)

and for all 0<t<Tmax0<t<T_{max}

Pst((u)S+13S2+14ωω)(,t)L2(ΔS+Pst(12(u)S+56S2+18ωω))(,t)L22.\frac{\left\|P_{st}\left((u\cdot\nabla)S+\frac{1}{3}S^{2}+\frac{1}{4}\omega\otimes\omega\right)(\cdot,t)\right\|_{L^{2}}}{\left\|\left(-\Delta S+P_{st}\left(\frac{1}{2}(u\cdot\nabla)S+\frac{5}{6}S^{2}+\frac{1}{8}\omega\otimes\omega\right)\right)(\cdot,t)\right\|_{L^{2}}}\leq 2. (1.50)

Then there is finite-time blowup with

Tmax<T:=E0+E02+f0K0f0,T_{max}<T_{*}:=\frac{-E_{0}+\sqrt{E_{0}^{2}+f_{0}K_{0}}}{f_{0}}, (1.51)

where K0K_{0} and E0E_{0} are taken as in Definition 1.7, and f0f_{0} is as defined above.

Remark 1.15.

Theorem 1.14 quantifies how close solutions of the Navier–Stokes strain equation have to be to solving the model equation in order to be guaranteed to blowup in finite-time. This result is—to the knowledge of the author—the first of its kind. There are many results stating that if some scale invariant quantity is finite, then solutions of the Navier–Stokes equation must be smooth, such as the aforementioned Ladyzhenskaya-Prodi-Serrin and Beale-Kato-Majda regularity criteria. Theorem 1.14 is the first result to say that, for some set of initial data, if a scale invariant quantity remains small enough for the history of the solution, there must be blowup in finite-time.

Remark 1.16.

The mechanism for blowup proposed in Theorem 1.14 for the Navier–Stokes equation is also consistent with research on the turbulent energy cascade. Very recently, Carbone and Bragg showed both theoretically and numerically that strain self-amplification is a more important factor in the average turbulent energy cascade than vortex stretching [CarboneBragg]. This gave a concrete statement to a line of inquiry on the turbulent energy cascade begun by Tsinober in [Tsinober]. The turbulent energy cascade is directly tied to the Navier–Stokes regularity problem, as finite-time blowup requires a transfer of energy to arbitrarily small scales, so this suggests that the self-amplification of strain is the most likely potential mechanism for the finite-time blowup of solutions of the Navier–Stokes equation. The conditional blowup result in this paper gives a quantitative estimate on the structure of the nonlinearity that will lead to finite-time blowup for the Navier–Stokes equation via the self-amplification of strain if it is maintained by the dynamics.

Remark 1.17.

Turbulent solutions of the Navier–Stokes equation are, almost by definition, difficult to impossible to write down in closed form. This poses a significant barrier to proving the existence of smooth solutions of the Navier–Stokes equation that blowup in finite-time: if finite-time blowup solutions do in fact exist, it will still almost certainly not be possible to give a negative answer to the Navier–Stokes regularity problem by providing a counterexample in closed form. Any progress on the Navier–Stokes regularity problem in the direction of proving the existence of finite-time blowup will likely require an interplay of analysis and numerics. Theorem 1.14 provides a quantitative criterion that could guide further numerical work searching for possible blowup solutions.

We cannot show that there are any solutions of Navier–Stokes equation which satisfy the perturbative condition in Theorem 1.14 up until TT^{*}. If we could, then this would solve the Navier–Stokes regularity problem by implying the existence of finite-time blowup. We can, however, use scaling arguments to prove that this condition is satisfied for short times for some solutions of the Navier–Stokes equation.

Theorem 1.18.

There exists a mild solution of the Navier–Stokes equation uC([0,Tmax);Hdf3)u\in C\left([0,T_{max});H^{3}_{df}\right) and ϵ>0\epsilon>0 such that

3S0H˙124det(S0)>0-3\left\|S^{0}\right\|_{\dot{H}^{1}}^{2}-4\int\det\left(S^{0}\right)>0 (1.52)

and for all 0t<ϵ0\leq t<\epsilon

Pst((u)S+13S2+14ωω)(,t)L2(ΔS+Pst(12(u)S+56S2+18ωω))(,t)L22.\frac{\left\|P_{st}\left((u\cdot\nabla)S+\frac{1}{3}S^{2}+\frac{1}{4}\omega\otimes\omega\right)(\cdot,t)\right\|_{L^{2}}}{\left\|\left(-\Delta S+P_{st}\left(\frac{1}{2}(u\cdot\nabla)S+\frac{5}{6}S^{2}+\frac{1}{8}\omega\otimes\omega\right)\right)(\cdot,t)\right\|_{L^{2}}}\leq 2. (1.53)
Remark 1.19.

In this paper we have taken the viscosity to be ν=1\nu=1. For the Navier–Stokes regularity problem, we can fix the viscosity to be ν=1\nu=1 without loss of generality, because it is equivalent up to rescaling to the Navier–Stokes regularity problem for arbitrary ν>0.\nu>0. It is useful, however, to see how the blowup results that we will prove scale with respect to the viscosity parameter ν>0.\nu>0. If we take the viscosity to be ν>0,\nu>0, then the Navier–Stokes equation is now given by

tuνΔu+Pdf(uu)=0,\partial_{t}u-\nu\Delta u+P_{df}\nabla\cdot(u\otimes u)=0, (1.54)

and the strain self-amplification model equation is given by

tSνΔS+23Pst(S2)=0.\partial_{t}S-\nu\Delta S+\frac{2}{3}P_{st}\left(S^{2}\right)=0. (1.55)

In this case, the condition

3S0H˙124det(S0)>0,-3\left\|S^{0}\right\|_{\dot{H}^{1}}^{2}-4\int\det\left(S^{0}\right)>0, (1.56)

in Theorems 1.11 and 1.14 is replaced with the condition

3νS0H˙124det(S0)>0.-3\nu\left\|S^{0}\right\|_{\dot{H}^{1}}^{2}-4\int\det\left(S^{0}\right)>0. (1.57)

Likewise the condition

Pst((u)S+13S2+14ωω)(,t)L2(ΔS+Pst(12(u)S+56S2+18ωω))(,t)L22.\frac{\left\|P_{st}\left((u\cdot\nabla)S+\frac{1}{3}S^{2}+\frac{1}{4}\omega\otimes\omega\right)(\cdot,t)\right\|_{L^{2}}}{\left\|\left(-\Delta S+P_{st}\left(\frac{1}{2}(u\cdot\nabla)S+\frac{5}{6}S^{2}+\frac{1}{8}\omega\otimes\omega\right)\right)(\cdot,t)\right\|_{L^{2}}}\leq 2. (1.58)

in Theorems 1.14 and 1.18 is replaced by

Pst((u)S+13S2+14ωω)(,t)L2(νΔS+Pst(12(u)S+56S2+18ωω))(,t)L22.\frac{\left\|P_{st}\left((u\cdot\nabla)S+\frac{1}{3}S^{2}+\frac{1}{4}\omega\otimes\omega\right)(\cdot,t)\right\|_{L^{2}}}{\left\|\left(-\nu\Delta S+P_{st}\left(\frac{1}{2}(u\cdot\nabla)S+\frac{5}{6}S^{2}+\frac{1}{8}\omega\otimes\omega\right)\right)(\cdot,t)\right\|_{L^{2}}}\leq 2. (1.59)
Remark 1.20.

We should note in particular this means that if S0Hst1S^{0}\in H^{1}_{st} and det(S0)>0,-\int\det\left(S^{0}\right)>0, then for all

0<ν<ν0:=4det(S0)3S0H˙12,0<\nu<\nu^{0}:=\frac{-4\int\det\left(S^{0}\right)}{3\left\|S^{0}\right\|_{\dot{H}^{1}}^{2}}, (1.60)

the strain self-amplification model equation with viscosity ν\nu blows up in finite-time. This implies that blowup for the strain self-amplification model equation is generic at sufficiently large Reynolds number, subject only to the geometric sign constraint on initial data, that det(S0)>0-\int\det\left(S^{0}\right)>0. This suggests that the self-amplification of strain is likely the driving factor behind possible blowup for the full Navier–Stokes equation, and any depletion of nonlinearity preventing finite-time blowup must come from the effects of advection and the nonlocal interaction of strain and vorticity.

This also means that finite-time blowup may occur for the strain self-amplification model equation even in simplified geometric settings where blowup is ruled out for the full Navier–Stokes equation. We will show that there is finite-time blowup for the strain self-amplification model equation even when restricted to axisymmetric, swirl-free solutions. This contrasts strongly with the Navier–Stokes equation where there is global regularity for arbitrarily large initial data in the axisymmetric, swirl-free case. There are also axisymmetric, swirl-free solutions of the Navier–Stokes equation that satisfy the perturbative condition for short times, as in Theorem 1.18. Such solutions cannot, of course, satisfy the conditions of Theorem 1.14, because they cannot blowup in finite-time, and hence the perturbative condition can only be satisfied for short times in such cases.

Remark 1.21.

Because such a wide range of initial data lead to finite-time blowup for the strain self-amplification model equation, the set of initial data for which there is finite-time blowup for this model equation is too broad a set to consider as possible candidates for finite-time blowup for the Navier–Stokes equation. While initial data that blowup in finite-time are ubiquitous at high Reynolds number, subject only to a sign constraint on the integral of the determinant of the strain, this does not necessarily mean that blowup itself is generic. There could be certain structures or scaling laws that emerge as the blowup time is approached for any blowup solution; further study is needed.

One possible avenue for further work would be to allow the dynamics of the strain self-amplification model equation to select candidates for blowup for the full Navier–Stokes equation. Consider a solution of the strain self-amplification model equation that is not axisymmetric, swirl-free, and that blows up in finite-time Tmax<+T_{max}<+\infty. If we take S(,Tmaxϵ)S(\cdot,T_{max}-\epsilon) for some 0<ϵTmax0<\epsilon\ll T_{max} as our initial data for the full Navier–Stokes equation, then this would be a very natural candidate for blowup if blowup does in fact occur for the full Navier–Stokes equation. To consider such an approach, more detailed study of the qualitative features of blowup solutions of the strain self-amplification model equation is needed. At present, essentially all we know about such solutions is a lower bound on the growth of enstrophy and that λ2+\lambda_{2}^{+} blows up in the scale-critical LtpLxqL_{t}^{p}L_{x}^{q} spaces.

In section 2, we will discuss the relationship between our results and previous results for simplified model equations for Navier–Stokes. In section 3, we will define a number of the spaces used in our analysis and give precise definitions of mild solutions. In section 4, we will develop the local well-posedness theory for the strain self-amplification model equation, including proving global well-posedness for small initial data, and scale critical regularity criteria in terms of λ2+\lambda_{2}^{+} and in terms of two vorticity components. In section 5, we will prove Theorem 1.11, demonstrating the existence of finite-time blowup for solutions of the strain self-amplification model equation, and will prove a number of properties about the set of initial data satisfying the hypothesis of this theorem. Finally in section 6, we will prove Theorem 1.14, the conditional blowup result for the full Navier–Stokes equation when a perturbative condition is satisfied by the history of the solution, and further show that this perturbative condition is satisfied for short times for some solutions of the Navier–Stokes equation.

2 Relationship to previous literature

There are a number of previous results that prove blowup for simplified model equations for Navier–Stokes with the hope of elucidating possibilities of extending this to the full Navier–Stokes equation. Montgomery-Smith introduced a scalar toy model equation, replacing the first order pseudo-differential operator PdfP_{df}\nabla\cdot by (Δ)12,-(-\Delta)^{\frac{1}{2}}, and replacing the quadratic term uuu\otimes u by u2,u^{2}, giving the scalar equation

tuΔu(Δ)12(u2)=0,\partial_{t}u-\Delta u-(-\Delta)^{\frac{1}{2}}\left(u^{2}\right)=0, (2.1)

and proved the existence of finite-time blowup solutions for this equation [MontgomerySmith]. This blowup result was extended by Gallagher and Paicu to a model equation on the space of divergence free vector fields by adjusting the Fourier symbol of the first order pseudo differential operator [GallagherToy]. However, while Gallagher and Paicu’s model equation is an evolution equation on natural constraint space, the space of divergence free vector fields, neither of these model equations respects the energy equality, and so both are still quite far from the actual fluid equations. They are nonetheless important in that they establish that it is not possible to prove global regularity for the Navier–Stokes equation using heat semi-group methods alone.

Tao improved on these earlier blowup results by introducing a Fourier space averaged Navier–Stokes model equation [TaoModel]. His model equation is given by

tuΔu+B~(u,u)=0,\partial_{t}u-\Delta u+\tilde{B}(u,u)=0, (2.2)

where B~(u,u)\tilde{B}(u,u) is a Fourier space averaged version of Pdf(uu)P_{df}\nabla\cdot(u\otimes u). This equation is an improvement over the previous results because B~\tilde{B} is constructed so that

B~(u,u),u=0,\left<\tilde{B}(u,u),u\right>=0, (2.3)

so Tao’s model equation (2.2) respects the energy equality, with for all 0<t<Tmax0<t<T_{max},

12u(t)L22+0tu(τ)H˙12dτ=12u0L22,\frac{1}{2}\|u(t)\|_{L^{2}}^{2}+\int_{0}^{t}\|u(\tau)\|_{\dot{H}^{1}}^{2}\mathop{}\!\mathrm{d}\tau=\frac{1}{2}\left\|u^{0}\right\|_{L^{2}}^{2}, (2.4)

while also exhibiting finite-time blowup. The operator B~\tilde{B} also has some of the same harmonic analysis bounds as those found for full Navier–Stokes equation, in particular

B~(u,u)L2CuL4uL4.\|\tilde{B}(u,u)\|_{L^{2}}\leq C\|u\|_{L^{4}}\|\nabla u\|_{L^{4}}. (2.5)

The fact that there are finite-time blowup solutions to Tao’s model equation shows that if there is global regularity for solutions of the Navier–Stokes equation with arbitrary smooth initial data, the proof will require more than the energy equality and the standard harmonic analysis techniques. New a priori bounds are needed. We will note in particular that the bound in (2.5) implies that Tao’s model equation respects the Ladyzhenskaya-Prodi-Serrin regularity criterion, that is if Tmax<+T_{max}<+\infty for a solution uu of (2.2), then for all 2p+3q=1,3<q+,\frac{2}{p}+\frac{3}{q}=1,3<q\leq+\infty,

0TmaxuLqp=+.\int_{0}^{T_{max}}\|u\|_{L^{q}}^{p}=+\infty. (2.6)

While the Tao model equation respects the energy equality and some of the structure of the velocity equation, it does not respect the structure of the vorticity or strain equations. In particular, Tao’s model does not respect—or at least has not been shown to respect—the identity for enstrophy growth in Proposition 1.9, the regularity criterion on λ2+\lambda_{2}^{+} in Theorem 1.10, or the regularity criterion on two components of the vorticity in (1.15). The finite-time blowup result for the strain self-amplification model equation is an advance on Tao’s model equation if the Navier–Stokes regularity problem is considered from the point of view of enstrophy growth. The model equation considered here, unlike Tao’s model equation, does not respect the energy equality; however, from a mathematical point of view, the energy equality is less fundamental to the Navier–Stokes regularity problem than the identity for enstrophy growth, because energy does not control regularity. Blowup for the Navier–Stokes equation in finite-time is equivalent to the blowup of enstrophy in finite-time, so mathematically it is very significant that we are able to show blowup for an evolution equation on Lst2L^{2}_{st} that respects the identity for enstrophy growth in Proposition 1.9. In summary, Tao’s model equation reflects more of the structure of the velocity formulation of the Navier–Stokes regularity problem, while the strain self-amplification model equation reflects more of the structure of both the strain and vorticity formulations of the regularity problem.

The strain self-amplification model equation is the first model equation of possible Navier–Stokes blowup that respects regularity criteria for the Navier–Stokes equation based not just on size, but on geometric structure as well. It is straightforward to show that the strain self-amplification, Montgomery-Smith, Gallagher-Paicu, and Tao model equations all respect the Ladyzhenskaya-Prodi-Serrin regularity criterion on the size of uu, but the strain self-amplification model equation also respects the regularity criterion on λ2+\lambda_{2}^{+} in Theorem 1.10. This implies as a corollary that the strain self-amplification model equation must respect the regularity criterion on two vorticity components proven by Chae and Choe as well. This suggests it captures significantly more of the geometry of potential Navier–Stokes blowup than any of the previous model equations, at least as far as deformation and vorticity are concerned.

Theorem 1.14 shows that the local part of the nonlinearity of the strain evolution equation tends to lead to finite-time blowup for a wide range of initial conditions, so there must be finite-time blowup for the Navier–Stokes equation similar to the blowup for the model equation for the self-amplification of strain unless the vorticity and advection terms act to deplete this nonlinearity and prevent blowup. This is consistent with a number of previous works for model equations related to the Navier–Stokes and Euler equations that suggest that advection plays a regularizing role. For instance, there are a number of previous works on the Constantin-Lax-Majda [ConstantinLaxMajda] and De Gregorio [DeGregorio] 1D models for the vorticity equation, which showed that advection may have a regularizing effect [DiegoCordoba, ElgindiAdvection, Sverak1DModels]. Theorem 1.11, which states that finite-time blowup occurs for a wide range of initial data for the strain self-amplification model equation, extends the analysis of the regularizing role of advection from 1D models that do not respect the structure of the constraint space, to a 3D model that does respect the structure of the constraint space.

There is also previous research on model equations for the axisymmetric Navier–Stokes and Euler equations which preserve more of the structures of three dimensional fluid mechanics than the Constantin-Lax-Majda or De Gregorio models. These model equations also show that advection plays a regularizing role [HouZhen, HouLiuWang]. Furthermore, there has been research on the possible role of advection in the depletion of nonlinearity related to its interaction with the pressure in the growth of subcritical LqL^{q} norms of uu [TranYu]. Theorem 1.14 is entirely novel, however, because it is the first perturbative, finite-time blowup result related to the possible role of nonlinear depletion by advection. The previous results were either heuristic or numerical; in contrast, Theorem 1.14 provides a quantitative condition guaranteeing blowup as long as the terms which could potentially deplete the nonlinear self-amplification of strain are small enough relative to strain self-amplification.

Finally, we should mention that very recently, Elgindi [ElgindiFiniteTimeEuler] and Elgindi, Ghoul, and Masmoudi [ElgindiGhoulMasmoudi] proved finite-time blowup for a class of C1,α(3)C^{1,\alpha}\left(\mathbb{R}^{3}\right) solutions of the Euler equation that conserve energy. While the question of blowup for smooth solutions to the Euler equation remains open, this represents an enormous step forward in providing an example of classical solutions to the Euler equation that blowup in finite-time. The blowup solutions of the Euler equation constructed in [ElgindiFiniteTimeEuler, ElgindiGhoulMasmoudi] are axisymetric and swirl-free, and are closely related to an example of finite-time blowup that we will construct for the strain self-amplification model equation. We will discuss this further in section 5.

3 Definitions

We begin by defining the homogeneous and inhomogeneous Hilbert spaces.

Definition 3.1.

For all s,Hs(3)s\in\mathbb{R},H^{s}\left(\mathbb{R}^{3}\right) will be the Hilbert space with norm

fHs2=3(1+4π2|ξ|2)s|f^(ξ)|2𝑑ξ=(1+4π2|ξ|2)s2f^L22,\|f\|_{H^{s}}^{2}=\int_{\mathbb{R}^{3}}\left(1+4\pi^{2}|\xi|^{2}\right)^{s}|\hat{f}(\xi)|^{2}d\xi=\left\|\left(1+4\pi^{2}|\xi|^{2}\right)^{\frac{s}{2}}\hat{f}\right\|_{L^{2}}^{2}, (3.1)

and for all 32<s<32,-\frac{3}{2}<s<\frac{3}{2}, H˙s(3)\dot{H}^{s}\left(\mathbb{R}^{3}\right) will be the homogeneous Hilbert space with norm

fH˙s2=3(2π|ξ|)2s|f^(ξ)|2𝑑ξ=(2π|ξ|)sf^L22.\|f\|_{\dot{H}^{s}}^{2}=\int_{\mathbb{R}^{3}}(2\pi|\xi|)^{2s}|\hat{f}(\xi)|^{2}d\xi=\left\|(2\pi|\xi|)^{s}\hat{f}\right\|_{L^{2}}^{2}. (3.2)

Note that when referring to Hs(3),H˙s(3),orLp(3),H^{s}\left(\mathbb{R}^{3}\right),\dot{H}^{s}\left(\mathbb{R}^{3}\right),\text{or}\>L^{p}\left(\mathbb{R}^{3}\right), the 3\mathbb{R}^{3} will often be omitted for brevity’s sake. All Hilbert and Lebesgue norms are taken over 3\mathbb{R}^{3} unless otherwise specified. Furthermore, 𝕊3×3\mathbb{S}^{3\times 3} will refer to the space of three by three symmetric matrices,

𝕊3×3={(adedbfefc):a,b,c,d,e,f}.\mathbb{S}^{3\times 3}=\left\{\left(\begin{array}[]{ccc}a&d&e\\ d&b&f\\ e&f&c\end{array}\right):a,b,c,d,e,f\in\mathbb{R}\right\}. (3.3)

We now define the subspaces of divergence free vector fields and strain matrices in Hilbert spaces.

Definition 3.2.

For all ss\in\mathbb{R} define HdfsHs(3;3)H^{s}_{df}\subset H^{s}\left(\mathbb{R}^{3};\mathbb{R}^{3}\right) by

Hdfs={uHs(3;3):ξu^(ξ)=0,almost everywhere ξ3}.H^{s}_{df}=\left\{u\in H^{s}\left(\mathbb{R}^{3};\mathbb{R}^{3}\right):\xi\cdot\hat{u}(\xi)=0,\text{almost everywhere }\xi\in\mathbb{R}^{3}\right\}. (3.4)

For all 32<s<32,-\frac{3}{2}<s<\frac{3}{2}, define H˙dfsH˙s(3;3)\dot{H}^{s}_{df}\subset\dot{H}^{s}\left(\mathbb{R}^{3};\mathbb{R}^{3}\right) by

H˙dfs={uH˙s(3;3):ξu^(ξ)=0,almost everywhere ξ3}.\dot{H}^{s}_{df}=\left\{u\in\dot{H}^{s}\left(\mathbb{R}^{3};\mathbb{R}^{3}\right):\xi\cdot\hat{u}(\xi)=0,\text{almost everywhere }\xi\in\mathbb{R}^{3}\right\}. (3.5)
Definition 3.3.

For all ss\in\mathbb{R} define HstsHs(3;𝕊3×3)H^{s}_{st}\subset H^{s}\left(\mathbb{R}^{3};\mathbb{S}^{3\times 3}\right) by

Hsts={sym(Δ)12u:uHdfs}.H^{s}_{st}=\left\{\nabla_{sym}(-\Delta)^{-\frac{1}{2}}u:u\in H^{s}_{df}\right\}. (3.6)

For all 32<s<32,-\frac{3}{2}<s<\frac{3}{2}, define H˙stsH˙s(3;𝕊3×3)\dot{H}^{s}_{st}\subset\dot{H}^{s}\left(\mathbb{R}^{3};\mathbb{S}^{3\times 3}\right) by

H˙sts={sym(Δ)12u:uH˙dfs}.\dot{H}^{s}_{st}=\left\{\nabla_{sym}(-\Delta)^{-\frac{1}{2}}u:u\in\dot{H}^{s}_{df}\right\}. (3.7)
Definition 3.4.

For all 1<q<+1<q<+\infty, define LstqL^{q}_{st} by

Lstq={SLq(3;𝕊3×3):tr(S)=0,S=sym(Δ)1(2div(S))}.L^{q}_{st}=\left\{S\in L^{q}\left(\mathbb{R}^{3};\mathbb{S}^{3\times 3}\right):\operatorname*{tr}(S)=0,S=\nabla_{sym}(-\Delta)^{-1}(-2\operatorname*{div}(S))\right\}. (3.8)
Remark 3.5.

We will note that we have already defined Lst2L^{2}_{st} in the introduction, so we now have two definitions of Lst2L^{2}_{st}. These definitions are equivalent, as was proven by the author in [MillerStrain]. The key reason for this is that, just as the vorticity can be inverted to obtain the velocity, with

u=×(Δ)1ω,u=\nabla\times(-\Delta)^{-1}\omega, (3.9)

so too can the strain be inverted to obtain the velocity, with

u=2div(Δ)1S.u=-2\operatorname*{div}(-\Delta)^{-1}S. (3.10)

This means that for all SLst2,S\in L^{2}_{st},

S=symuS=\nabla_{sym}u (3.11)

if and only if

u=2div(Δ)1S.u=-2\operatorname*{div}(-\Delta)^{-1}S. (3.12)

This implies the condition in Definition 3.4 in the case q=2q=2 is equivalent to the the condition in Definition 1.2. See [MillerStrain] for more details. We will also note that Definition 3.4 is well defined because the operator 2symdiv(Δ)1-2\nabla_{sym}\operatorname*{div}(-\Delta)^{-1} is a bounded linear operator mapping LqLqL^{q}\longrightarrow L^{q} for all 1<q<+1<q<+\infty. This follows from the boundedness of the Riesz transform R=(Δ)12R=\nabla(-\Delta)^{-\frac{1}{2}} because

2symdiv(Δ)1S=2RsymRS.-2\nabla_{sym}\operatorname*{div}(-\Delta)^{-1}S=-2R_{sym}R\cdot S. (3.13)

We will also define axisymmetric, swirl-free vector fields and strain matrices.

Definition 3.6.

Begin by letting

r\displaystyle r =x12+x22\displaystyle=\sqrt{x_{1}^{2}+x_{2}^{2}} (3.14)
z\displaystyle z =x3\displaystyle=x_{3} (3.15)
er\displaystyle e_{r} =1r(x1,x2,0)\displaystyle=\frac{1}{r}(x_{1},x_{2},0) (3.16)
ez\displaystyle e_{z} =(0,0,1).\displaystyle=(0,0,1). (3.17)

We will say that uH˙df1u\in\dot{H}^{1}_{df} is an axisymmetric, swirl-free vector field if

u(x)=ur(r,z)er+uz(r,z)ez.u(x)=u_{r}(r,z)e_{r}+u_{z}(r,z)e_{z}. (3.18)

Note that the divergence free condition can be expressed in this case by

u=rur+1rur+zuz=0.\nabla\cdot u=\partial_{r}u_{r}+\frac{1}{r}u_{r}+\partial_{z}u_{z}=0. (3.19)

We will say that SLst2S\in L^{2}_{st} is an axisymmetric, swirl-free strain matrix, if

S=symu,S=\nabla_{sym}u, (3.20)

where uH˙df1u\in\dot{H}^{1}_{df} is an axisymmetric, swirl-free vector field.

We conclude this section by providing the precise definitions for mild solutions of the Navier–Stokes equation, the Navier–Stokes strain equation, and the strain self-amplification model equation.

Definition 3.7.

A velocity field uC([0,Tmax);H˙df1)u\in C\left([0,T_{max});\dot{H}^{1}_{df}\right) is a mild solution to the Navier–Stokes equation if it satisfies (1.4) in the sense of Duhamel’s formula, that is, if for all 0<t<Tmax0<t<T_{max}

u(t)=etΔu00teτΔPdf(uu)(tτ)dτ.u(t)=e^{t\Delta}u^{0}-\int_{0}^{t}e^{\tau\Delta}P_{df}\nabla\cdot(u\otimes u)(t-\tau)\mathop{}\!\mathrm{d}\tau. (3.21)

Note that etΔe^{t\Delta} is defined in terms of convolution with the heat kernel

G(x,t)=1(4πt)32exp(|x|24t)G(x,t)=\frac{1}{(4\pi t)^{\frac{3}{2}}}\exp\left(-\frac{|x|^{2}}{4t}\right) (3.22)

so that

etΔf=G(,t)f.e^{t\Delta}f=G(\cdot,t)*f. (3.23)
Remark 3.8.

TmaxT_{max} is the maximal time of existence for a mild solution. If there is a mild solution globally in time for some initial data u0H˙df1u^{0}\in\dot{H}^{1}_{df}, then Tmax=+,T_{max}=+\infty, and if there is not a mild solution globally in time, then Tmax<+T_{max}<+\infty is the blowup time when the solution becomes singular.

Definition 3.9.

A strain matrix SC([0,Tmax);Lst2)S\in C\left([0,T_{max});L^{2}_{st}\right) is a mild solution to the Navier–Stokes strain equation if it satisfies (1.20) in the sense of Duhamel’s formula, that is, if for all 0<t<Tmax0<t<T_{max}

S(t)=etΔS00teτΔPst((u)S+S2+14ωω)(tτ)dτ,S(t)=e^{t\Delta}S^{0}-\int_{0}^{t}e^{\tau\Delta}P_{st}\left((u\cdot\nabla)S+S^{2}+\frac{1}{4}\omega\otimes\omega\right)(t-\tau)\mathop{}\!\mathrm{d}\tau, (3.24)

with u=(Δ)1(2div(S))u=(-\Delta)^{-1}\left(-2\operatorname*{div}(S)\right) and ω=×u.\omega=\nabla\times u.

Definition 3.10.

SC([0,Tmax);Lst2)S\in C\left([0,T_{max});L^{2}_{st}\right) is a mild solution to the strain self-amplification model equation (1.40) if SS satisfies this equation in the sense of Duhamel’s formula, that is for all 0<t<Tmax0<t<T_{max}

S(t)=etΔS0230teτΔPst(S2)(tτ)dτS(t)=e^{t\Delta}S^{0}-\frac{2}{3}\int_{0}^{t}e^{\tau\Delta}P_{st}\left(S^{2}\right)(t-\tau)\mathop{}\!\mathrm{d}\tau (3.25)

4 Some properties of strain self-amplification model equation

We begin this section by considering the local-in-time existence of mild solutions to the strain self-amplification model equation.

Theorem 4.1.

Let C=(332gL2)4,C=\left(\frac{3}{32\|g\|_{L^{2}}}\right)^{4}, where g(x)=1(2π)32exp(|x|24).g(x)=\frac{1}{(2\pi)^{\frac{3}{2}}}\exp\left(-\frac{|x|^{2}}{4}\right). For all S0Lst2S^{0}\in L^{2}_{st}, there exists a unique mild solution to the strain self-amplification model equation, SC([0,Tmax);Lst2),S\in C\left([0,T_{max});L^{2}_{st}\right), with TmaxCS0L24T_{max}\geq\frac{C}{\left\|S^{0}\right\|_{L^{2}}^{4}}. Furthermore, SC((0,Tmax);H),S\in C\left((0,T_{max});H^{\infty}\right), and is therefore smooth for all positive times up until possible blowup.

Proof.

The proof of Theorem 4.1 is essentially the same as the proof of local existence of mild solutions for the Navier–Stokes equation introduced by Kato and Fujita. It will be based on a Banach fixed point argument.

We begin by fixing

T<CS0L24.T<\frac{C}{\left\|S^{0}\right\|_{L^{2}}^{4}}. (4.1)

Note that this implies that

323gL2S0L2T14<1.\frac{32}{3}\|g\|_{L^{2}}\left\|S^{0}\right\|_{L^{2}}T^{\frac{1}{4}}<1. (4.2)

Define the map W:C([0,T];Lst2)C([0,T];Lst2),W:C\left([0,T];L^{2}_{st}\right)\to C\left([0,T];L^{2}_{st}\right), by

W[M](t)=etΔS0+0teτΔPst(M2(tτ))dτ.W[M](t)=e^{t\Delta}S^{0}+\int_{0}^{t}e^{\tau\Delta}P_{st}\left(M^{2}(t-\tau)\right)\mathop{}\!\mathrm{d}\tau. (4.3)

Note that SS being a mild solution of the heat equation is equivalent to SS being a fixed point of this map with W[S]=S.W[S]=S.

We will first show that if MC([0,T];L2)2S0L2,\|M\|_{C\left([0,T];L^{2}\right)}\leq 2\left\|S^{0}\right\|_{L^{2}}, then W[M]C([0,T];L2)2S0L2.\|W[M]\|_{C\left([0,T];L^{2}\right)}\leq 2\left\|S^{0}\right\|_{L^{2}}. Recall that

etΔf=G(,t)f,e^{t\Delta}f=G(\cdot,t)*f, (4.4)

where

G(x,t)=t32g(t12x).G(x,t)=t^{-\frac{3}{2}}g(t^{-\frac{1}{2}}x). (4.5)

Therefore we can compute that

G(,t)L2=gL2t34.\|G(\cdot,t)\|_{L^{2}}=\|g\|_{L^{2}}t^{-\frac{3}{4}}. (4.6)

Applying Young’s inequality for convolutions we find that for all 0tT0\leq t\leq T

W[M](t)L2\displaystyle\|W[M](t)\|_{L^{2}} S0L2+230tPst(eτΔ)M2)L2dτ\displaystyle\leq\left\|S^{0}\right\|_{L^{2}}+\frac{2}{3}\int_{0}^{t}\left\|P_{st}\left(e^{\tau\Delta})M^{2}\right)\right\|_{L^{2}}\mathop{}\!\mathrm{d}\tau (4.7)
S0L2+230tG(,t)L2M2(tτ)L1dτ\displaystyle\leq\left\|S^{0}\right\|_{L^{2}}+\frac{2}{3}\int_{0}^{t}\|G(\cdot,t)\|_{L^{2}}\|M^{2}(t-\tau)\|_{L^{1}}\mathop{}\!\mathrm{d}\tau (4.8)
S0L2+23M2C([0,T];L1)0tgL2τ34dτ\displaystyle\leq\left\|S^{0}\right\|_{L^{2}}+\frac{2}{3}\left\|M^{2}\right\|_{C\left([0,T];L^{1}\right)}\int_{0}^{t}\|g\|_{L^{2}}\tau^{-\frac{3}{4}}\mathop{}\!\mathrm{d}\tau (4.9)
S0L2+83MC([0,T];L2)2gL2t14\displaystyle\leq\left\|S^{0}\right\|_{L^{2}}+\frac{8}{3}\left\|M\right\|^{2}_{C\left([0,T];L^{2}\right)}\|g\|_{L^{2}}t^{\frac{1}{4}} (4.10)
S0L2+83MC([0,T];L2)2gL2T14.\displaystyle\leq\left\|S^{0}\right\|_{L^{2}}+\frac{8}{3}\left\|M\right\|^{2}_{C\left([0,T];L^{2}\right)}\|g\|_{L^{2}}T^{\frac{1}{4}}. (4.11)

Using the fact that MC([0,T];L2)2S0L2,\left\|M\right\|_{C\left([0,T];L^{2}\right)}\leq 2\left\|S^{0}\right\|_{L^{2}}, and recalling that 323gL2S0L2T14<1,\frac{32}{3}\|g\|_{L^{2}}\left\|S^{0}\right\|_{L^{2}}T^{\frac{1}{4}}<1, we can see that

83MC([0,T];L2)2gL2T14\displaystyle\frac{8}{3}\left\|M\right\|^{2}_{C\left([0,T];L^{2}\right)}\|g\|_{L^{2}}T^{\frac{1}{4}} 323S0L22gL2T14\displaystyle\leq\frac{32}{3}\left\|S^{0}\right\|_{L^{2}}^{2}\|g\|_{L^{2}}T^{\frac{1}{4}} (4.12)
S0L2.\displaystyle\leq\left\|S^{0}\right\|_{L^{2}}. (4.13)

This implies that

W[M]C([0,T];L2)2S0L2.\|W[M]\|_{C\left([0,T];L^{2}\right)}\leq 2\left\|S^{0}\right\|_{L^{2}}. (4.14)

Therefore WW is an automorphism on the closed ball

B={MC([0,T];Lst2):MC([0,T];L2)2S0L2}.B=\left\{M\in C\left([0,T];L^{2}_{st}\right):\|M\|_{C\left([0,T];L^{2}\right)}\leq 2\left\|S^{0}\right\|_{L^{2}}\right\}. (4.15)

We will now show that WW is a contraction mapping on B.B. Fix M,QB.M,Q\in B. Using Young’s convolution inequality as above we can compute that for all 0tT0\leq t\leq T

W[M](t)W[Q](t)L2\displaystyle\|W[M](t)-W[Q](t)\|_{L^{2}} =13Pst0tetΔ((M+Q)(MQ)+(MQ)(M+Q))L2\displaystyle=\frac{1}{3}\left\|P_{st}\int_{0}^{t}e^{t\Delta}\left((M+Q)(M-Q)+(M-Q)(M+Q)\right)\right\|_{L^{2}} (4.16)
83gL2t14M+QC([0,T];L2)MQC([0,T];L2)\displaystyle\leq\frac{8}{3}\|g\|_{L^{2}}t^{\frac{1}{4}}\|M+Q\|_{C\left([0,T];L^{2}\right)}\|M-Q\|_{C\left([0,T];L^{2}\right)} (4.17)
83gL2t14(MC([0,T];L2)+QC([0,T];L2))MQC([0,T];L2)\displaystyle\leq\frac{8}{3}\|g\|_{L^{2}}t^{\frac{1}{4}}\left(\|M\|_{C\left([0,T];L^{2}\right)}+\|Q\|_{C\left([0,T];L^{2}\right)}\right)\|M-Q\|_{C\left([0,T];L^{2}\right)} (4.18)
323gL2S0L2T14MQC([0,T];L2).\displaystyle\leq\frac{32}{3}\|g\|_{L^{2}}\left\|S^{0}\right\|_{L^{2}}T^{\frac{1}{4}}\|M-Q\|_{C\left([0,T];L^{2}\right)}. (4.19)

Letting

r\displaystyle r =323gL2S0L2T14\displaystyle=\frac{32}{3}\|g\|_{L^{2}}\left\|S^{0}\right\|_{L^{2}}T^{\frac{1}{4}} (4.20)
<1,\displaystyle<1, (4.21)

we can find that

W[M]W[Q]C([0,T];L2)rMQC([0,T];L2).\|W[M]-W[Q]\|_{C\left([0,T];L^{2}\right)}\leq r\|M-Q\|_{C\left([0,T];L^{2}\right)}. (4.22)

Note that BB is a complete metric space so by the Banach fixed point theorem, we can conclude that there exists a unique SBC([0,T];Lst2),S\in B\subset C\left([0,T];L^{2}_{st}\right), such that

W[S]=S.W[S]=S. (4.23)

This implies that there is a unique, mild solution with initial data in S0Lst2S^{0}\in L^{2}_{st} locally in time. Note that the higher regularity SC((0,T];H)S\in C\left((0,T];H^{\infty}\right) is a result of the smoothing due to the heat kernel, but we will not go through the details of that here. This higher regularity follows from a a bootstrapping argument that is essentially the same as the argument in the case of the Navier–Stokes equation given in [KatoFujita].

We will now prove a useful proposition giving an identity for the determinant of 3×33\times 3, symmetric, trace free matrices.

Proposition 4.2.

Suppose M𝕊3×3M\in\mathbb{S}^{3\times 3} is a 3×33\times 3 symmetric matrix such that tr(M)=0.\operatorname*{tr}(M)=0. Then

tr(M3)=3det(M).\operatorname*{tr}(M^{3})=3\det(M). (4.24)
Proof.

Every symmetric matrix is diagonalizable over ,\mathbb{R}, so let λ1λ2λ3\lambda_{1}\leq\lambda_{2}\leq\lambda_{3} be the eigenvalues of M.M. Using the trace free condition we can see that

tr(M)\displaystyle\operatorname*{tr}(M) =λ1+λ2+λ3\displaystyle=\lambda_{1}+\lambda_{2}+\lambda_{3} (4.25)
=0.\displaystyle=0. (4.26)

Therefore we can compute that

tr(M3)\displaystyle\operatorname*{tr}\left(M^{3}\right) =λ13+λ23+λ33\displaystyle=\lambda_{1}^{3}+\lambda_{2}^{3}+\lambda_{3}^{3} (4.27)
=(λ1λ2)3+λ13+λ23\displaystyle=(-\lambda_{1}-\lambda_{2})^{3}+\lambda_{1}^{3}+\lambda_{2}^{3} (4.28)
=3λ12λ23λ1λ22\displaystyle=-3\lambda_{1}^{2}\lambda_{2}-3\lambda_{1}\lambda_{2}^{2} (4.29)
=3(λ1λ2)λ1λ2\displaystyle=3(-\lambda_{1}-\lambda_{2})\lambda_{1}\lambda_{2} (4.30)
=3λ1λ2λ3\displaystyle=3\lambda_{1}\lambda_{2}\lambda_{3} (4.31)
=3det(M).\displaystyle=3\det(M). (4.32)

This completes the proof. ∎

Using this proposition, we will show that the strain self-amplification model equation has the same identity for enstrophy growth as the Navier–Stokes strain equation.

Proposition 4.3.

Suppose SC([0,Tmax);Lst2)S\in C\left([0,T_{max});L^{2}_{st}\right) is a mild solution to the strain self-amplification model equation. Then for all 0<t<Tmax,0<t<T_{max},

ddtS(,t)L22\displaystyle\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\|S(\cdot,t)\|_{L^{2}}^{2} =2SH˙1243tr(S3)\displaystyle=-2\|S\|_{\dot{H}^{1}}^{2}-\frac{4}{3}\int\operatorname*{tr}\left(S^{3}\right) (4.33)
=2SH˙124det(S).\displaystyle=-2\|S\|_{\dot{H}^{1}}^{2}-4\int\det(S). (4.34)
Proof.

Taking the derivative in time of the L2L^{2} norm, we plug into the strain self-amplification model equation (1.40), finding that

ddtS(,t)L22\displaystyle\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\|S(\cdot,t)\|_{L^{2}}^{2} =2tS,S\displaystyle=2\left<\partial_{t}S,S\right> (4.35)
=2ΔS+23Pst(S2),S\displaystyle=-2\left<-\Delta S+\frac{2}{3}P_{st}\left(S^{2}\right),S\right> (4.36)
=2SH˙1243Pst(S2),S\displaystyle=-2\|S\|_{\dot{H}^{1}}^{2}-\frac{4}{3}\left<P_{st}\left(S^{2}\right),S\right> (4.37)
=2SH˙1243S2,S\displaystyle=-2\|S\|_{\dot{H}^{1}}^{2}-\frac{4}{3}\left<S^{2},S\right> (4.38)
=2SH˙1243tr(S3)\displaystyle=-2\|S\|_{\dot{H}^{1}}^{2}-\frac{4}{3}\int\operatorname*{tr}\left(S^{3}\right) (4.39)
=2SH˙124det(S),\displaystyle=-2\|S\|_{\dot{H}^{1}}^{2}-4\int\det(S), (4.40)

where we have used the fact that SLst2S\in L^{2}_{st} to drop the projection PstP_{st}, and the fact that SS is symmetric to compute the inner product, and finally applied Proposition 4.2. This completes the proof. ∎

In fact, the vortex stretching and the integral of the determinant of the strain can be related in a general way as follows. This will be useful in showing the term we dropped in the model equation does not contribute to enstrophy growth.

Proposition 4.4.

For all SLst3,S\in L^{3}_{st},

4det(S)=S;ωω,-4\int\det(S)=\left<S;\omega\otimes\omega\right>, (4.41)

where u=2div(Δ)1Su=-2\operatorname*{div}(-\Delta)^{-1}S and ω=×u.\omega=\nabla\times u. In particular this implies that

S,Pst(13S2+14ωω)=0.\left<S,P_{st}\left(\frac{1}{3}S^{2}+\frac{1}{4}\omega\otimes\omega\right)\right>=0. (4.42)
Proof.

The first step of the proof will be to show that

tr((u)3)=0.\int\operatorname*{tr}\left(\left(\nabla u\right)^{3}\right)=0. (4.43)

We will begin by recalling that by definition

S=sym(Δ)1(2div(S)),S=\nabla_{sym}(-\Delta)^{-1}(-2\operatorname*{div}(S)), (4.44)

and so we can see that S=symuS=\nabla_{sym}u. We may conclude that

u=2div(Δ)1S,\nabla u=-2\nabla\operatorname*{div}(-\Delta)^{-1}S, (4.45)

Using the boundedness of the Riesz from L3L3L^{3}\longrightarrow L^{3}, this implies that uL3,\nabla u\in L^{3}, so the integral in (4.43) is absolutely convergent.

Using the divergence free condition we note that

i=13uixi=u=0.\sum_{i=1}^{3}\frac{\partial u_{i}}{\partial x_{i}}=\nabla\cdot u=0. (4.46)

Therefore for all uCc(3;3),u=0,u\in C^{\infty}_{c}\left(\mathbb{R}^{3};\mathbb{R}^{3}\right),\nabla\cdot u=0, we can integrate by parts—without worrying about boundary terms because of compact support—finding that

tr((u)3)\displaystyle\int\operatorname*{tr}\left((\nabla u)^{3}\right) =i,j,k=13ujxiukxjuixk\displaystyle=\sum_{i,j,k=1}^{3}\int\frac{\partial u_{j}}{\partial x_{i}}\frac{\partial u_{k}}{\partial x_{j}}\frac{\partial u_{i}}{\partial x_{k}} (4.47)
=i,j,k=13uj2ukxixjuixk\displaystyle=-\sum_{i,j,k=1}^{3}\int u_{j}\frac{\partial^{2}u_{k}}{\partial x_{i}\partial x_{j}}\frac{\partial u_{i}}{\partial x_{k}} (4.48)
=i,j,k=13ujukxi2uixjxk\displaystyle=\sum_{i,j,k=1}^{3}\int u_{j}\frac{\partial u_{k}}{\partial x_{i}}\frac{\partial^{2}u_{i}}{\partial x_{j}\partial x_{k}} (4.49)
=i,j,k=13ujxkukxiuixj\displaystyle=-\sum_{i,j,k=1}^{3}\int\frac{\partial u_{j}}{\partial x_{k}}\frac{\partial u_{k}}{\partial x_{i}}\frac{\partial u_{i}}{\partial x_{j}} (4.50)
=tr((u)3)\displaystyle=-\int\operatorname*{tr}\left((\nabla u)^{3}\right) (4.51)
=0.\displaystyle=0. (4.52)

Because Cc(3;3)C^{\infty}_{c}\left(\mathbb{R}^{3};\mathbb{R}^{3}\right) is dense in L3(3;3),L^{3}\left(\mathbb{R}^{3};\mathbb{R}^{3}\right), this is sufficient to guarantee that for all uL3,u=0,\nabla u\in L^{3},\nabla\cdot u=0,

tr((u)3)=0.\int\operatorname*{tr}\left(\left(\nabla u\right)^{3}\right)=0. (4.53)

We know that u=S+A.\nabla u=S+A. Using the fact that SS is symmetric and AA is anti-symmetric, and that all anti-symmetric matrices are trace free, we compute that

tr((u)3)=tr(S3)+3tr(SA2).\operatorname*{tr}\left((\nabla u)^{3}\right)=\operatorname*{tr}\left(S^{3}\right)+3\operatorname*{tr}\left(SA^{2}\right). (4.54)

Recall from the introduction that

A=12(0ω3ω2ω30ω1ω2ω10),A=\frac{1}{2}\left(\begin{matrix}0&\omega_{3}&-\omega_{2}\\ -\omega_{3}&0&\omega_{1}\\ \omega_{2}&-\omega_{1}&0\\ \end{matrix}\right), (4.55)

and we can compute that

A2=14ωω14|ω|2I3.A^{2}=\frac{1}{4}\omega\otimes\omega-\frac{1}{4}|\omega|^{2}I_{3}. (4.56)

Therefore we find that

3tr(SA2)\displaystyle 3\operatorname*{tr}\left(SA^{2}\right) =34(S:ωω)+34|ω|2tr(S)\displaystyle=\frac{3}{4}\left(S:\omega\otimes\omega\right)+\frac{3}{4}|\omega|^{2}\operatorname*{tr}(S) (4.57)
=34(S:ωω).\displaystyle=\frac{3}{4}\left(S:\omega\otimes\omega\right). (4.58)

Applying Proposition 4.2, we find that

tr(S3)=3det(S).\operatorname*{tr}\left(S^{3}\right)=3\det(S). (4.59)

Therefore we find that

tr((u)3)=3det(S)+34(S:ωω).\operatorname*{tr}\left((\nabla u)^{3}\right)=3\det(S)+\frac{3}{4}\left(S:\omega\otimes\omega\right). (4.60)

Integrating this equality over 3\mathbb{R}^{3} we find

S;ωω+4det(S)\displaystyle\left<S;\omega\otimes\omega\right>+4\int\det(S) =43(3det(S)+34(S:ωω))\displaystyle=\frac{4}{3}\int\left(3\det(S)+\frac{3}{4}\left(S:\omega\otimes\omega\right)\right) (4.61)
=43tr((u)3)\displaystyle=\frac{4}{3}\int\operatorname*{tr}\left(\left(\nabla u\right)^{3}\right) (4.62)
=0.\displaystyle=0. (4.63)

Finally we will compute that

Pst(13S2+14ωω);S\displaystyle\left<P_{st}\left(\frac{1}{3}S^{2}+\frac{1}{4}\omega\otimes\omega\right);S\right> =13S2+14ωω;S\displaystyle=\left<\frac{1}{3}S^{2}+\frac{1}{4}\omega\otimes\omega;S\right> (4.64)
=14S;ωω+13tr(S3)\displaystyle=\frac{1}{4}\left<S;\omega\otimes\omega\right>+\frac{1}{3}\int\operatorname*{tr}\left(S^{3}\right) (4.65)
=14S;ωω+det(S)\displaystyle=\frac{1}{4}\left<S;\omega\otimes\omega\right>+\int\det(S) (4.66)
=0,\displaystyle=0, (4.67)

and this completes the proof. The author would like to thank the anonymous referee from an earlier version of [MillerStrain] for this observation. ∎

Using this result, we will observe that the term we have dropped from the Navier–Stokes strain equation to obtain our strain self-amplification model equation is orthogonal to SS with respect to the L2L^{2} inner product.

Corollary 4.5.

Suppose SHst1,S\in H^{1}_{st}, with S=symuS=\nabla_{sym}u and ω=×u,\omega=\nabla\times u, then

Pst((u)S+13S2+14ωω);S=0.\left<P_{st}\left((u\cdot\nabla)S+\frac{1}{3}S^{2}+\frac{1}{4}\omega\otimes\omega\right);S\right>=0. (4.68)
Proof.

We begin by observing that H1H˙12L3H^{1}\hookrightarrow\dot{H}^{\frac{1}{2}}\hookrightarrow L^{3}, and so clearly SLst3S\in L^{3}_{st}. Applying Proposition 4.4 we can see that

Pst(13S2+14ωω);S=0.\left<P_{st}\left(\frac{1}{3}S^{2}+\frac{1}{4}\omega\otimes\omega\right);S\right>=0. (4.69)

Next we use the divergence free condition, u=0,\nabla\cdot u=0, and the fact that we have sufficient regularity to integrate by parts to compute that

Pst((u)S);S\displaystyle\left<P_{st}\left((u\cdot\nabla)S\right);S\right> =(u)S;S\displaystyle=\left<(u\cdot\nabla)S;S\right> (4.70)
=S;(u)S\displaystyle=-\left<S;(u\cdot\nabla)S\right> (4.71)
=0.\displaystyle=0. (4.72)

This completes the proof. ∎

Note that this means that the term Pst((u)S+13S2+14ωω),P_{st}\left((u\cdot\nabla)S+\frac{1}{3}S^{2}+\frac{1}{4}\omega\otimes\omega\right), does not contribute to enstrophy growth, so when we write the Navier–Stokes strain equation as

tSΔS+23Pst(S2)+Pst((u)S+13S2+14ωω)=0,\partial_{t}S-\Delta S+\frac{2}{3}P_{st}\left(S^{2}\right)+P_{st}\left((u\cdot\nabla)S+\frac{1}{3}S^{2}+\frac{1}{4}\omega\otimes\omega\right)=0, (4.73)

only the terms ΔS-\Delta S and 23Pst(S2)\frac{2}{3}P_{st}\left(S^{2}\right) contribute to enstrophy growth. This is the justification for studying the dynamics of enstrophy growth using a model equation that drops the term Pst((u)S+13S2+14ωω),P_{st}\left((u\cdot\nabla)S+\frac{1}{3}S^{2}+\frac{1}{4}\omega\otimes\omega\right), retaining only the terms that actually contribute to the growth of enstrophy.

The strain self-amplification model equation, like the Navier–Stokes strain equation, is invariant under the rescaling

Sλ(s,t)=λ2S(λx,λ2t).S^{\lambda}(s,t)=\lambda^{2}S(\lambda x,\lambda^{2}t). (4.74)

We will now show the existence of global smooth solutions of the strain self-amplification model equation with small initial data in the critical Hilbert space H˙12\dot{H}^{-\frac{1}{2}}.

Theorem 4.6.

Suppose S0Lst2H˙st12S^{0}\in L^{2}_{st}\cap\dot{H}^{-\frac{1}{2}}_{st} and

S0H˙12<3342π.\left\|S^{0}\right\|_{\dot{H}^{-\frac{1}{2}}}<\frac{3\sqrt{3}}{4\sqrt{2}}\pi. (4.75)

Then there exists a unique, global smooth solution to the strain self-amplification model equation SC([0,+);Lst2),S\in C\left([0,+\infty);L^{2}_{st}\right), that is Tmax=+.T_{max}=+\infty.

Proof.

We begin by observing there must be a smooth solution SC((0,Tmax);Lst2),S\in C\left((0,T_{max});L^{2}_{st}\right), for some Tmax>0.T_{max}>0. We will consider the growth of the H˙12\dot{H}^{-\frac{1}{2}} on this time interval. We will use the fractional Sobolev inequality proven by Lieb [Lieb, LiebLoss]. For all fL32(3)f\in L^{\frac{3}{2}}\left(\mathbb{R}^{3}\right)

fH˙121216π13fL32,\|f\|_{\dot{H}^{-\frac{1}{2}}}\leq\frac{1}{2^{\frac{1}{6}}\pi^{\frac{1}{3}}}\|f\|_{L^{\frac{3}{2}}}, (4.76)

and for all gL3(3),g\in L^{3}\left(\mathbb{R}^{3}\right),

gL31216π13gH˙12.\|g\|_{L^{3}}\leq\frac{1}{2^{\frac{1}{6}}\pi^{\frac{1}{3}}}\|g\|_{\dot{H}^{\frac{1}{2}}}. (4.77)

Applying both fractional Sobolev inequalities we find that

ddtS(t)H˙122\displaystyle\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\|S(t)\|_{\dot{H}^{-\frac{1}{2}}}^{2} =2SH˙12243(Δ)12S,S2\displaystyle=-2\|S\|_{\dot{H}^{\frac{1}{2}}}^{2}-\frac{4}{3}\left<(-\Delta)^{-\frac{1}{2}}S,S^{2}\right> (4.78)
2SH˙122+43(Δ)12SH˙12S2H˙12\displaystyle\leq-2\|S\|_{\dot{H}^{\frac{1}{2}}}^{2}+\frac{4}{3}\left\|(-\Delta)^{-\frac{1}{2}}S\right\|_{\dot{H}^{\frac{1}{2}}}\left\|S^{2}\right\|_{\dot{H}^{-\frac{1}{2}}} (4.79)
2SH˙122+431216π13SH˙12S2L32\displaystyle\leq-2\|S\|_{\dot{H}^{\frac{1}{2}}}^{2}+\frac{4}{3}\frac{1}{2^{\frac{1}{6}}\pi^{\frac{1}{3}}}\left\|S\right\|_{\dot{H}^{-\frac{1}{2}}}\left\|S^{2}\right\|_{L^{\frac{3}{2}}} (4.80)
2SH˙122+431216π13SH˙12SL32\displaystyle\leq-2\|S\|_{\dot{H}^{\frac{1}{2}}}^{2}+\frac{4}{3}\frac{1}{2^{\frac{1}{6}}\pi^{\frac{1}{3}}}\left\|S\right\|_{\dot{H}^{-\frac{1}{2}}}\left\|S\right\|_{L^{3}}^{2} (4.81)
2SH˙122+431212πSH˙12SH˙122\displaystyle\leq-2\|S\|_{\dot{H}^{\frac{1}{2}}}^{2}+\frac{4}{3}\frac{1}{2^{\frac{1}{2}}\pi}\left\|S\right\|_{\dot{H}^{-\frac{1}{2}}}\left\|S\right\|_{\dot{H}^{\frac{1}{2}}}^{2} (4.82)
2SH˙122(1+23πSH˙12).\displaystyle\leq 2\|S\|_{\dot{H}^{\frac{1}{2}}}^{2}\left(-1+\frac{\sqrt{2}}{3\pi}\|S\|_{\dot{H}^{-\frac{1}{2}}}\right). (4.83)

From this bound on the growth of the H˙12\dot{H}^{-\frac{1}{2}} norm it is clear that if

S(t)H˙12<3π2,\|S(t)\|_{\dot{H}^{-\frac{1}{2}}}<\frac{3\pi}{\sqrt{2}}, (4.84)

then

ddtS(t)H˙12<0.\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\|S(t)\|_{\dot{H}^{-\frac{1}{2}}}<0. (4.85)

We know that

S0H˙12\displaystyle\left\|S^{0}\right\|_{\dot{H}^{-\frac{1}{2}}} <3342π\displaystyle<\frac{3\sqrt{3}}{4\sqrt{2}}\pi (4.86)
<3π2,\displaystyle<\frac{3\pi}{\sqrt{2}}, (4.87)

so we can conclude that for all 0t<Tmax0\leq t<T_{max}

S(t)H˙12<3342π.\|S(t)\|_{\dot{H}^{-\frac{1}{2}}}<\frac{3\sqrt{3}}{4\sqrt{2}}\pi. (4.88)

To finish the proof we will need to consider bounds on the enstrophy growth in terms of the H˙12\dot{H}^{-\frac{1}{2}} norm. In addition to the fractional sharp Sobolev inequality, we will also make use of the ordinary sharp Sobolev inequality [Talenti, Sobolev], which states that for all fL6(3),f\in L^{6}\left(\mathbb{R}^{3}\right),

fL613(2π)23fH˙1.\|f\|_{L^{6}}\leq\frac{1}{\sqrt{3}}\left(\frac{2}{\pi}\right)^{\frac{2}{3}}\|f\|_{\dot{H}^{1}}. (4.89)

Applying the Sobolev inequality, the fractional Sobolev inequality, Hölder’s inequality, and the product rule to the identity for enstrophy growth Proposition 4.3, we find

ddtS(t)L22\displaystyle\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\|S(t)\|_{L^{2}}^{2} =2SH˙1243S,S2\displaystyle=-2\|S\|_{\dot{H}^{1}}^{2}-\frac{4}{3}\left<S,S^{2}\right> (4.90)
2SH˙12+43SH˙12S2H˙12\displaystyle\leq-2\|S\|_{\dot{H}^{1}}^{2}+\frac{4}{3}\|S\|_{\dot{H}^{-\frac{1}{2}}}\left\|S^{2}\right\|_{\dot{H}^{\frac{1}{2}}} (4.91)
=2SH˙12+43SH˙12(S2)H˙12\displaystyle=-2\|S\|_{\dot{H}^{1}}^{2}+\frac{4}{3}\|S\|_{\dot{H}^{-\frac{1}{2}}}\left\|\nabla\left(S^{2}\right)\right\|_{\dot{H}^{-\frac{1}{2}}} (4.92)
2SH˙12+431216π13SH˙12(S2)L32\displaystyle\leq-2\|S\|_{\dot{H}^{1}}^{2}+\frac{4}{3}\frac{1}{2^{\frac{1}{6}}\pi^{\frac{1}{3}}}\|S\|_{\dot{H}^{-\frac{1}{2}}}\left\|\nabla\left(S^{2}\right)\right\|_{L^{\frac{3}{2}}} (4.93)
2SH˙12+431216π13SH˙122SL2SL6\displaystyle\leq-2\|S\|_{\dot{H}^{1}}^{2}+\frac{4}{3}\frac{1}{2^{\frac{1}{6}}\pi^{\frac{1}{3}}}\|S\|_{\dot{H}^{-\frac{1}{2}}}2\|\nabla S\|_{L^{2}}\|S\|_{L^{6}} (4.94)
2SH˙12+831216π1313(2π)23SH˙12SH˙12\displaystyle\leq-2\|S\|_{\dot{H}^{1}}^{2}+\frac{8}{3}\frac{1}{2^{\frac{1}{6}}\pi^{\frac{1}{3}}}\frac{1}{\sqrt{3}}\left(\frac{2}{\pi}\right)^{\frac{2}{3}}\|S\|_{\dot{H}^{-\frac{1}{2}}}\|S\|_{\dot{H}^{1}}^{2} (4.95)
=2SH˙12(1+4233πSH˙12).\displaystyle=2\|S\|_{\dot{H}^{1}}^{2}\left(-1+\frac{4\sqrt{2}}{3\sqrt{3}\pi}\|S\|_{\dot{H}^{-\frac{1}{2}}}\right). (4.96)

We have already shown that for all 0t<Tmax,0\leq t<T_{max},

4233πS(t)H˙12<1,\frac{4\sqrt{2}}{3\sqrt{3}\pi}\|S(t)\|_{\dot{H}^{-\frac{1}{2}}}<1, (4.97)

so for all 0t<Tmax,0\leq t<T_{max},

ddtS(t)L220.\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\|S(t)\|_{L^{2}}^{2}\leq 0. (4.98)

This implies that for all 0t<Tmax,0\leq t<T_{max},

S(t)L2S0L2.\|S(t)\|_{L^{2}}\leq\left\|S^{0}\right\|_{L^{2}}. (4.99)

We know from Theorem 4.1 that for all 0t<Tmax,0\leq t<T_{max},

Tmaxt>CS(t)L24.T_{max}-t>\frac{C}{\|S(t)\|_{L^{2}}^{4}}. (4.100)

This means that if Tmax<+,T_{max}<+\infty, then

limtTmaxS(t)L2=+.\lim_{t\to T_{max}}\|S(t)\|_{L^{2}}=+\infty. (4.101)

We know that for all 0t<Tmax,S(t)L2S0L2,0\leq t<T_{max},\|S(t)\|_{L^{2}}\leq\left\|S^{0}\right\|_{L^{2}}, so we can conclude that Tmax=+.T_{max}=+\infty. This completes the proof. ∎

Remark 4.7.

We will note that the assumption SH˙12L2S\in\dot{H}^{-\frac{1}{2}}\cap L^{2} is not actually necessary; it is sufficient to have small initial data in H˙12,\dot{H}^{-\frac{1}{2}}, to guarantee global regularity with no assumption that S0L2.S^{0}\in L^{2}. However, dropping this assumption makes the proof a little more technical, and, more importantly, the whole point of a strain self-amplification model equation is to model enstrophy growth, so if our solution is not in L2L^{2} the model does not mean very much.

Likewise, some of the other results in this section are not optimal: for example it should be straightforward to prove the local existence of mild, smooth solutions with initial data in Bp,2+3p,B^{-2+\frac{3}{p}}_{p,\infty}, for 2p<+,2\leq p<+\infty, without too much difficulty. Because the strain self-amplification model equation is adapted specifically to study L2L^{2} solutions however, getting local existence or small data results down to the largest scale critical spaces is not particularly useful or illuminating.

We will now prove that because the strain self-amplification model equation has the same identity for enstrophy growth as the Navier–Stokes equation, it also has a regularity criterion on the positive part of the middle eigenvalue of the strain matrix that is precisely the same as the analogous result for the Navier–Stokes equation, Theorem 1.10.

Theorem 4.8.

Suppose SC([0,Tmax);Lst2)S\in C\left([0,T_{max});L^{2}_{st}\right) is a mild solution to the strain self-amplification model equation. Let λ1(x,t)λ2(x,t)λ3(x,t)\lambda_{1}(x,t)\leq\lambda_{2}(x,t)\leq\lambda_{3}(x,t) be the eigenvalues of S(x,t),S(x,t), and let λ2+(x,t)=max{0,λ2(x,t)}.\lambda^{+}_{2}(x,t)=\max\left\{0,\lambda_{2}(x,t)\right\}. Then for all 3q+2p=2,32<q+,\frac{3}{q}+\frac{2}{p}=2,\frac{3}{2}<q\leq+\infty, there exists Cq>0C_{q}>0 depending only on qq such that for all 0<t<Tmax0<t<T_{max},

S(t)L22S0L22exp(Cq0tλ2+(τ)Lqpdτ).\|S(t)\|_{L^{2}}^{2}\leq\left\|S^{0}\right\|_{L^{2}}^{2}\exp\left(C_{q}\int_{0}^{t}\left\|\lambda_{2}^{+}(\tau)\right\|_{L^{q}}^{p}\mathop{}\!\mathrm{d}\tau\right). (4.102)

In particular, if Tmax<+,T_{max}<+\infty, then

0Tmaxλ2+(t)Lqpdt=+.\int_{0}^{T_{max}}\left\|\lambda_{2}^{+}(t)\right\|_{L^{q}}^{p}\mathop{}\!\mathrm{d}t=+\infty. (4.103)
Proof.

We know from Theorem 4.1 that if Tmax<+,T_{max}<+\infty, then

limtTmaxS(t)L22=+,\lim_{t\to T_{max}}\|S(t)\|_{L^{2}}^{2}=+\infty, (4.104)

so it suffices to prove the estimate (4.102). Because tr(S)=0,\operatorname*{tr}(S)=0, we know that λ10\lambda_{1}\leq 0 and λ30.\lambda_{3}\geq 0. Therefore we know that

λ1λ30.-\lambda_{1}\lambda_{3}\geq 0. (4.105)

We can therefore compute from the identity for enstrophy growth in Proposition 4.3 that

ddtS(t)L22\displaystyle\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\|S(t)\|_{L^{2}}^{2} =2SH˙124det(S)\displaystyle=-2\|S\|_{\dot{H}^{1}}^{2}-4\int\det(S) (4.106)
=2SH˙12+4(λ1λ3)λ2\displaystyle=-2\|S\|_{\dot{H}^{1}}^{2}+4\int(-\lambda_{1}\lambda_{3})\lambda_{2} (4.107)
2SH˙12+4(λ1λ3)λ2+\displaystyle\leq-2\|S\|_{\dot{H}^{1}}^{2}+4\int(-\lambda_{1}\lambda_{3})\lambda_{2}^{+} (4.108)
2SH˙12+2λ2+|S|2\displaystyle\leq-2\|S\|_{\dot{H}^{1}}^{2}+2\int\lambda_{2}^{+}|S|^{2} (4.109)
Cqλ2+LqpSL22,\displaystyle\leq C_{q}\|\lambda_{2}^{+}\|_{L^{q}}^{p}\|S\|_{L^{2}}^{2}, (4.110)

after applying Hölder’s inequality, the Sobolev inequality, and Young’s inequality. This computation is precisely the same as the one done in the proof of the regularity criterion on λ2+\lambda_{2}^{+} in [MillerStrain], so we refer the reader there for more detail on these steps. Applying Grönwall’s inequality, we find for all 0<t<Tmax,0<t<T_{max},

S(t)L22S0L2exp(Cq0tλ2+(τ)Lqpdτ).\|S(t)\|_{L^{2}}^{2}\leq\left\|S^{0}\right\|_{L^{2}}\exp\left(C_{q}\int_{0}^{t}\|\lambda_{2}^{+}(\tau)\|_{L^{q}}^{p}\mathop{}\!\mathrm{d}\tau\right). (4.111)

This completes the proof. ∎

This regularity criterion means that there must be unbounded planar stretching in the scale critical LtpLxqL^{p}_{t}L^{q}_{x} spaces in order for finite-time blowup to occur. The strength of the strain formulation of the Navier–Stokes regularity problem means that not only does the strain self-amplification model equation respect geometric regularity criteria in terms of the strain; it also respects the regularity criterion on two components of the vorticity proven for the full Navier–Stokes equation by Chae and Choe in [ChaeVort].

Corollary 4.9.

Suppose SC([0,Tmax);Lst2)S\in C\left([0,T_{max});L^{2}_{st}\right) is a mild solution to the strain self-amplification model equation. Let ω=×(Δ)1(2div(S))\omega=\nabla\times(-\Delta)^{-1}(-2\operatorname*{div}(S)) be the vorticity associated with the strain SS. Then for all 3q+2p=2,32<q<+,\frac{3}{q}+\frac{2}{p}=2,\frac{3}{2}<q<+\infty, there exists Cq>0C_{q}>0 depending only on qq such that for all 0<t<Tmax0<t<T_{max},

S(t)L22S0L22exp(Cq0te3×ω(τ)Lqpdτ).\|S(t)\|_{L^{2}}^{2}\leq\left\|S^{0}\right\|_{L^{2}}^{2}\exp\left(C_{q}\int_{0}^{t}\left\|e_{3}\times\omega(\tau)\right\|_{L^{q}}^{p}\mathop{}\!\mathrm{d}\tau\right). (4.112)

In particular, if Tmax<+,T_{max}<+\infty, then

0Tmaxe3×ω(t)Lqpdt=+.\int_{0}^{T_{max}}\left\|e_{3}\times\omega(t)\right\|_{L^{q}}^{p}\mathop{}\!\mathrm{d}t=+\infty. (4.113)
Proof.

We know that tr(S)=0\operatorname*{tr}(S)=0 and λ1λ2λ3\lambda_{1}\leq\lambda_{2}\leq\lambda_{3}, so λ2\lambda_{2} is the smallest eigenvalue of SS in magnitude. This implies that for all unit vectors v3v\in\mathbb{R}^{3}, |λ2||Sv||\lambda_{2}|\leq|Sv|. Consequently we can see that for all x3x\in\mathbb{R}^{3},

|λ2||Se3|.|\lambda_{2}|\leq|Se_{3}|. (4.114)

Next we observe that

2Se3\displaystyle 2Se_{3} =u3+3u\displaystyle=\nabla u_{3}+\partial_{3}u (4.115)
e3×ω\displaystyle e_{3}\times\omega =u33u.\displaystyle=\nabla u_{3}-\partial_{3}u. (4.116)

We can see that u3\nabla u_{3} is a gradient and that 3u=0\nabla\cdot\partial_{3}u=0, and so using the Helmholtz projections unto the spaces of divergence free vector fields and gradients, we can see that

u3\displaystyle\nabla u_{3} =Pgr(e3×ω)\displaystyle=P_{gr}(e_{3}\times\omega) (4.117)
3u\displaystyle\partial_{3}u =Pdf(e3×ω).\displaystyle=-P_{df}(e_{3}\times\omega). (4.118)

The boundedness of the Helmholtz decomposition then implies that for all 1<q<+1<q<+\infty,

λ2Lq\displaystyle\|\lambda_{2}\|_{L^{q}} Se3Lq\displaystyle\leq\|Se_{3}\|_{L^{q}} (4.119)
12u3Lq+123uLq\displaystyle\leq\frac{1}{2}\|\nabla u_{3}\|_{L^{q}}+\frac{1}{2}\|\partial_{3}u\|_{L^{q}} (4.120)
Cqe3×ωLq\displaystyle\leq C_{q}\|e_{3}\times\omega\|_{L^{q}} (4.121)

The result then follows as an immediate corollary of Theorem 4.8. ∎

5 Finite-time blowup for the strain self-amplification model equation

In this section, we will prove the existence of finite-time blowup for the strain self-amplification model equation. We begin by proving a nonlinear differential inequality giving a lower bound on the rate of enstrophy growth that is sufficient to guarantee finite-time blowup for some initial data.

Proposition 5.1.

Suppose SC([0,Tmax);Hst1)S\in C\left([0,T_{max});H^{1}_{st}\right) is a mild solution of the strain self-amplification model equation. Then for all 0t<Tmax0\leq t<T_{max},

ddtE(t)g0E(t)32,\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}E(t)\geq g_{0}E(t)^{\frac{3}{2}}, (5.1)

where

g0=3S0H˙124det(S0)S0L23.g_{0}=\frac{-3\left\|S^{0}\right\|_{\dot{H}^{1}}^{2}-4\int\det\left(S^{0}\right)}{\left\|S^{0}\right\|_{L^{2}}^{3}}. (5.2)
Proof.

We will begin by letting

g(t)\displaystyle g(t) =3S(,t)H˙124det(S(,t))S(,t)L23\displaystyle=\frac{-3\|S(\cdot,t)\|_{\dot{H}^{1}}^{2}-4\int\det(S(\cdot,t))}{\|S(\cdot,t)\|_{L^{2}}^{3}} (5.3)
=3SH˙1243tr(S3)E(t)32.\displaystyle=\frac{-3\|S\|_{\dot{H}^{1}}^{2}-\frac{4}{3}\int\operatorname*{tr}\left(S^{3}\right)}{E(t)^{\frac{3}{2}}}. (5.4)

Differentiating gg, we find that for all 0<t<Tmax0<t<T_{max},

ddtg(t)\displaystyle\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}g(t) =6ΔS+23Pst(S2)L22E(t)3232(3SH˙1243tr(S3))(2SH˙1243tr(S3))E(t)52\displaystyle=\frac{6\|-\Delta S+\frac{2}{3}P_{st}\left(S^{2}\right)\|_{L^{2}}^{2}}{E(t)^{\frac{3}{2}}}-\frac{3}{2}\frac{\left(-3\|S\|_{\dot{H}^{1}}^{2}-\frac{4}{3}\int\operatorname*{tr}\left(S^{3}\right)\right)\left(-2\|S\|_{\dot{H}^{1}}^{2}-\frac{4}{3}\int\operatorname*{tr}\left(S^{3}\right)\right)}{E(t)^{\frac{5}{2}}} (5.5)
=6ΔS+23Pst(S2)L22SL2332(3SH˙1243tr(S3))(2SH˙1243tr(S3))SL25\displaystyle=\frac{6\|-\Delta S+\frac{2}{3}P_{st}\left(S^{2}\right)\|_{L^{2}}^{2}}{\|S\|_{L^{2}}^{3}}-\frac{3}{2}\frac{\left(-3\|S\|_{\dot{H}^{1}}^{2}-\frac{4}{3}\int\operatorname*{tr}\left(S^{3}\right)\right)\left(-2\|S\|_{\dot{H}^{1}}^{2}-\frac{4}{3}\int\operatorname*{tr}\left(S^{3}\right)\right)}{\|S\|_{L^{2}}^{5}} (5.6)
6ΔS+23Pst(S2)L22SL2332(2SH˙1243tr(S3))2SL25\displaystyle\geq\frac{6\|-\Delta S+\frac{2}{3}P_{st}\left(S^{2}\right)\|_{L^{2}}^{2}}{\|S\|_{L^{2}}^{3}}-\frac{3}{2}\frac{\left(-2\|S\|_{\dot{H}^{1}}^{2}-\frac{4}{3}\int\operatorname*{tr}\left(S^{3}\right)\right)^{2}}{\|S\|_{L^{2}}^{5}} (5.7)
=6SL25(SL22ΔS+23Pst(S2)L22(SH˙1223tr(S3))2).\displaystyle=\frac{6}{\|S\|_{L^{2}}^{5}}\left(\|S\|_{L^{2}}^{2}\left\|-\Delta S+\frac{2}{3}P_{st}\left(S^{2}\right)\right\|_{L^{2}}^{2}-\left(-\|S\|_{\dot{H}^{1}}^{2}-\frac{2}{3}\int\operatorname*{tr}\left(S^{3}\right)\right)^{2}\right). (5.8)

Applying Young’s inequality, we find

SH˙1223tr(S3)\displaystyle-\|S\|_{\dot{H}^{1}}^{2}-\frac{2}{3}\int\operatorname*{tr}\left(S^{3}\right) =ΔS+23Pst(S2),S\displaystyle=-\left<-\Delta S+\frac{2}{3}P_{st}\left(S^{2}\right),S\right> (5.9)
ΔS+23Pst(S2)L2SL2,\displaystyle\leq\left\|\Delta S+\frac{2}{3}P_{st}\left(S^{2}\right)\right\|_{L^{2}}\|S\|_{L^{2}}, (5.10)

and so we can conclude that for all 0<t<Tmax0<t<T_{max},

ddtg(t)0.\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}g(t)\geq 0. (5.11)

Therefore, we can conclude that for all 0<t<Tmax0<t<T_{max},

g(t)g0.g(t)\geq g_{0}. (5.12)

Finally, we observe that for all 0<t<Tmax0<t<T_{max},

ddtE(t)\displaystyle\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}E(t) =2SH˙124det(S)\displaystyle=-2\|S\|_{\dot{H}^{1}}^{2}-4\int\det(S) (5.13)
3SH˙124det(S)\displaystyle\geq-3\|S\|_{\dot{H}^{1}}^{2}-4\int\det(S) (5.14)
=g(t)E(t)32\displaystyle=g(t)E(t)^{\frac{3}{2}} (5.15)
g0E(t)32,\displaystyle\geq g_{0}E(t)^{\frac{3}{2}}, (5.16)

and this completes the proof. ∎

Remark 5.2.

Note that as long as SS is not the trivial solution—as long as SH˙12>0\|S\|_{\dot{H}^{1}}^{2}>0—then the inequality in (5.14) is strict and therefore for all 0<t<Tmax0<t<T_{max},

ddtE(t)>g0E(t)32.\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}E(t)>g_{0}E(t)^{\frac{3}{2}}. (5.17)

This differential inequality is sufficient to guarantee finite-time blowup for any solution with initial data such that g0>0g_{0}>0. We will now prove Theorem 1.11, which is restated here for the reader’s convenience.

Theorem 5.3.

Suppose SC([0,Tmax);Hst1)S\in C\left([0,T_{max});H^{1}_{st}\right) is a mild solution of the strain self-amplification model equation, such that

3S0H˙124det(S0)>0.-3\left\|S^{0}\right\|_{\dot{H}^{1}}^{2}-4\int\det\left(S^{0}\right)>0. (5.18)

Then for all 0<t<Tmax0<t<T_{max},

E(t)>E0(1r0t)2,E(t)>\frac{E_{0}}{(1-r_{0}t)^{2}}, (5.19)

where

r0=3S0H˙124det(S0)2S0L22.r_{0}=\frac{-3\left\|S^{0}\right\|_{\dot{H}^{1}}^{2}-4\int\det\left(S^{0}\right)}{2\left\|S^{0}\right\|_{L^{2}}^{2}}. (5.20)

Note in particular that this implies

Tmax2S0L223S0L224det(S0).T_{max}\leq\frac{2\left\|S^{0}\right\|_{L^{2}}^{2}}{-3\left\|S^{0}\right\|_{L^{2}}^{2}-4\int\det\left(S^{0}\right)}. (5.21)

Furthermore, for all 2p+3q=2,32<q+,\frac{2}{p}+\frac{3}{q}=2,\frac{3}{2}<q\leq+\infty,

0Tmaxλ2+(t)Lqpdt=+.\int_{0}^{T_{max}}\|\lambda_{2}^{+}(t)\|_{L^{q}}^{p}\mathop{}\!\mathrm{d}t=+\infty. (5.22)
Proof.

The main argument of the proof will be integrating the differential inequality in Theorem 5.1. Applying Proposition 5.1, we can see that for all 0<t<Tmax0<t<T_{max},

ddtE(t)>g0E(t)32,\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}E(t)>g_{0}E(t)^{\frac{3}{2}}, (5.23)

where g0>0g_{0}>0 by hypothesis. Applying the chain rule and (5.23), for all 0<t<Tmax0<t<T_{max},

ddt(E(t)12)\displaystyle\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\left(-E(t)^{-\frac{1}{2}}\right) =12E(t)32ddtE(t)\displaystyle=\frac{1}{2}E(t)^{-\frac{3}{2}}\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}E(t) (5.24)
>12g0,\displaystyle>\frac{1}{2}g_{0}, (5.25)

Integrating this differential inequality, we find that for all 0<t<Tmax0<t<T_{max},

E012E(t)12>12g0t.E_{0}^{-\frac{1}{2}}-E(t)^{-\frac{1}{2}}>\frac{1}{2}g_{0}t. (5.26)

This implies that

E(t)12<E01212g0t,E(t)^{-\frac{1}{2}}<E_{0}^{-\frac{1}{2}}-\frac{1}{2}g_{0}t, (5.27)

and therefore that

E(t)>1(E01212g0t)2.E(t)>\frac{1}{\left(E_{0}^{-\frac{1}{2}}-\frac{1}{2}g_{0}t\right)^{2}}. (5.28)

Multiplying the numerator and denominator by E0E_{0}, we find that for all 0<t<Tmax0<t<T_{max},

E(t)>E0(112g0E012t)2.E(t)>\frac{E_{0}}{\left(1-\frac{1}{2}g_{0}E_{0}^{\frac{1}{2}}t\right)^{2}}. (5.29)

It is easy to check that

r0=12g0E012,r_{0}=\frac{1}{2}g_{0}E_{0}^{\frac{1}{2}}, (5.30)

so we have now established that for all 0<t<Tmax0<t<T_{max},

E(t)>E0(1r0t)2.E(t)>\frac{E_{0}}{(1-r_{0}t)^{2}}. (5.31)

Furthermore, this clearly implies that

Tmax\displaystyle T_{max} 1r0\displaystyle\leq\frac{1}{r_{0}} (5.32)
=2S0L223S0L224det(S0).\displaystyle=\frac{2\left\|S^{0}\right\|_{L^{2}}^{2}}{-3\left\|S^{0}\right\|_{L^{2}}^{2}-4\int\det\left(S^{0}\right)}. (5.33)

Finally, applying Theorem 4.8, we conclude that for all 2p+3q=2,32<q+,\frac{2}{p}+\frac{3}{q}=2,\frac{3}{2}<q\leq+\infty,

0Tmaxλ2+(t)Lqpdt=+,\int_{0}^{T_{max}}\|\lambda_{2}^{+}(t)\|_{L^{q}}^{p}\mathop{}\!\mathrm{d}t=+\infty, (5.34)

and this completes the proof. ∎

Next we will show that the set of initial data satisfying the hypothesis of Theorem 5.3 is nonempty and bounded below in H˙12\dot{H}^{-\frac{1}{2}}. We will also show that λ2+\lambda_{2}^{+} is bounded below in L32L^{\frac{3}{2}} for all SS in this set. First we will need to perform a few calculations related to the determinant of the strain.

Proposition 5.4.

There exists SHst1,S\in H^{1}_{st}, axisymmetric and swirl-free, such that

3det(S)>0.-\int_{\mathbb{R}^{3}}\det(S)>0. (5.35)

Note that we will say that SS is axisymmetric and swirl-free if S=symu,S=\nabla_{sym}u, where uu is an axisymmetric, swirl-free, divergence free vector field.

Proof.

We begin by taking uHdf2u\in H^{2}_{df} using axisymmetric coordinates, letting

u(x)=(r2rz2)exp(r2z2)er+(2z+2r2z)exp(r2z2)ez.u(x)=(r-2rz^{2})\exp\left(-r^{2}-z^{2}\right)e_{r}+(-2z+2r^{2}z)\exp\left(-r^{2}-z^{2}\right)e_{z}. (5.36)

We will observe that

u(x)=((12x32)(x1x20)+(2x3+2(x12+x22)x3)(001))exp((x12+x22+x32)),u(x)=\left((1-2x_{3}^{2})\left(\begin{array}[]{c}x_{1}\\ x_{2}\\ 0\end{array}\right)+\left(-2x_{3}+2\left(x_{1}^{2}+x_{2}^{2}\right)x_{3}\right)\left(\begin{array}[]{c}0\\ 0\\ 1\end{array}\right)\right)\exp\left(-\left(x_{1}^{2}+x_{2}^{2}+x_{3}^{2}\right)\right), (5.37)

and so not only do we have uH2u\in H^{2}, but we have the stronger result that uu must be in the Schwartz class of smooth functions, which have, along with all their derivatives, faster than polynomial decay at infinity. Taking the divergence of uu we find that

u\displaystyle\nabla\cdot u =(r+1r)ur+zuz\displaystyle=\left(\partial_{r}+\frac{1}{r}\right)u_{r}+\partial_{z}u_{z} (5.38)
=((24z22r2+4r2z2)+(2+2r2+4z24r2z2))exp(r2z2)\displaystyle=\Big{(}\left(2-4z^{2}-2r^{2}+4r^{2}z^{2}\right)+\left(-2+2r^{2}+4z^{2}-4r^{2}z^{2}\right)\Big{)}\exp\left(-r^{2}-z^{2}\right) (5.39)
=0,\displaystyle=0, (5.40)

as required so that uHdf2u\in H^{2}_{df}. Taking the curl of uu we will find that

ω\displaystyle\omega =(zurruz)eθ\displaystyle=\left(\partial_{z}u_{r}-\partial_{r}u_{z}\right)e_{\theta} (5.41)
=((4rz2rz+4rz3)(4rz+4rz4r3z))exp(r2z2)eθ\displaystyle=\Big{(}\left(-4rz-2rz+4rz^{3}\right)-(4rz+4rz-4r^{3}z)\Big{)}\exp\left(-r^{2}-z^{2}\right)e_{\theta} (5.42)
=(14rz+4rz3+4r3z)exp(r2z2)eθ.\displaystyle=(-14rz+4rz^{3}+4r^{3}z)\exp\left(-r^{2}-z^{2}\right)e_{\theta}. (5.43)

Next we will observe that the gradient can be represented in axisymmetric coordinates as

=1reθθ+err+ezz.\nabla=\frac{1}{r}e_{\theta}\partial_{\theta}+e_{r}\partial_{r}+e_{z}\partial_{z}. (5.44)

Using this representation and recalling that

er=(cos(θ)sin(θ)0),e_{r}=\left(\begin{array}[]{c}\cos(\theta)\\ \sin(\theta)\\ 0\end{array}\right), (5.45)

so we can see that

θer=eθ\partial_{\theta}e_{r}=e_{\theta} (5.46)

This means we can compute that

tr(S(eθeθ))\displaystyle\operatorname*{tr}\left(S(e_{\theta}\otimes e_{\theta})\right) =tr(u(eθeθ))\displaystyle=\operatorname*{tr}\left(\nabla u(e_{\theta}\otimes e_{\theta})\right) (5.47)
=urr\displaystyle=\frac{u_{r}}{r} (5.48)
=(12z2)exp(r2z2)\displaystyle=\left(1-2z^{2}\right)\exp\left(-r^{2}-z^{2}\right) (5.49)

Applying Proposition 4.4 we find that

det(S)\displaystyle-\int\det(S) =14S;ωω\displaystyle=\frac{1}{4}\left<S;\omega\otimes\omega\right> (5.50)
=143tr(S(eθeθ))(x)|ω(x)|2dx\displaystyle=\frac{1}{4}\int_{\mathbb{R}^{3}}\operatorname*{tr}\left(S(e_{\theta}\otimes e_{\theta})\right)(x)|\omega(x)|^{2}\mathop{}\!\mathrm{d}x (5.51)
=1402πr(12z2)(14rz+4rz3+4r3z)2exp(3r23z2)dzdr\displaystyle=\frac{1}{4}\int_{0}^{\infty}\int_{-\infty}^{\infty}2\pi r\left(1-2z^{2}\right)\left(-14rz+4rz^{3}+4r^{3}z\right)^{2}\exp\left(-3r^{2}-3z^{2}\right)\mathop{}\!\mathrm{d}z\mathop{}\!\mathrm{d}r (5.52)
=π00r(12z2)(14rz+4rz3+4r3z)2exp(3r23z2)dzdr\displaystyle=\pi\int_{0}^{\infty}\int_{0}^{\infty}r\left(1-2z^{2}\right)\left(-14rz+4rz^{3}+4r^{3}z\right)^{2}\exp\left(-3r^{2}-3z^{2}\right)\mathop{}\!\mathrm{d}z\mathop{}\!\mathrm{d}r (5.53)
=4π00r3z2(12z2)(7+2z2+2r2)2exp(3r23z2)dzdr,\displaystyle=4\pi\int_{0}^{\infty}\int_{0}^{\infty}r^{3}z^{2}\left(1-2z^{2}\right)\left(-7+2z^{2}+2r^{2}\right)^{2}\exp\left(-3r^{2}-3z^{2}\right)\mathop{}\!\mathrm{d}z\mathop{}\!\mathrm{d}r, (5.54)

using the fact that integrand is even in zz. Making the substitution, v=z2,w=r2,v=z^{2},w=r^{2}, we find that

det(S)\displaystyle-\int\det(S) =π00wv(12v)(7+2v+2w)2exp(3w3v)dvdw\displaystyle=\pi\int_{0}^{\infty}\int_{0}^{\infty}w\sqrt{v}\left(1-2v\right)\left(-7+2v+2w\right)^{2}\exp\left(-3w-3v\right)\mathop{}\!\mathrm{d}v\mathop{}\!\mathrm{d}w (5.55)
=8π32813\displaystyle=\frac{8\pi^{\frac{3}{2}}}{81\sqrt{3}} (5.56)

Therefore we can conclude that there exists SHst1,S\in H^{1}_{st}, axisymmetric and swirl-free, such that

det(S)>0,-\int\det(S)>0, (5.57)

and this completes the proof. ∎

Theorem 5.5.

Let the set of initial data satisfying the hypotheses of Theorem 5.3, ΓblowupHst1,\Gamma_{\mathrm{blowup}}\subset H^{1}_{st}, be given by

Γblowup={SHst1:3SH˙124det(S)>0}.\Gamma_{\mathrm{blowup}}=\left\{S\in H^{1}_{st}:-3\|S\|_{\dot{H}^{1}}^{2}-4\int\det(S)>0\right\}. (5.58)

Then Γblowup\Gamma_{\mathrm{blowup}} is nonempty.

Proof.

Take any SHst1S\in H^{1}_{st} such that

det(S)>0.-\int\det(S)>0. (5.59)

We know such an SS must exist from Proposition 5.4. If we multiply such an SHst1S\in H^{1}_{st} by a sufficiently large constant we will end up with an element of Γblowup.\Gamma_{\mathrm{blowup}}. In particular we compute

limm+(3mSH˙124det(mS))\displaystyle\lim_{m\to+\infty}\left(-3\|mS\|_{\dot{H}^{1}}^{2}-4\int\det(mS)\right) =limm+(3m2(SH˙12)+4m3(det(S)))\displaystyle=\lim_{m\to+\infty}\left(-3m^{2}\left(\|S\|_{\dot{H}^{1}}^{2}\right)+4m^{3}\left(-\int\det(S)\right)\right) (5.60)
=+.\displaystyle=+\infty. (5.61)

Therefore we may conclude that for all SHst1S\in H^{1}_{st} such that det(S)>0,-\int\det(S)>0, for sufficiently large m>0,mSΓblowup.m>0,mS\in\Gamma_{\mathrm{blowup}}. This completes the proof. ∎

Remark 5.6.

Note that the near the origin, the velocity corresponding to finite-time blowup for the strain self-amplification model equation from Proposition 5.4 and Theorem 5.5 has a very similar geometric structure to the C1,αC^{1,\alpha} finite-time blowup solution to the Euler equation from [ElgindiFiniteTimeEuler, ElgindiGhoulMasmoudi]. Both involve planar stretching and axial compression near the origin. In particular, approximating the velocity in Proposition 5.4 near the origin by the first order Taylor polynomial,

u(x)rer2zez.u(x)\approx re_{r}-2ze_{z}. (5.62)
Remark 5.7.

The fact the Γblowup\Gamma_{\mathrm{blowup}} is nonempty means that the condition in Theorem 5.3 is satisfied for some initial data, and so we can conclude that there must exist solutions of the strain self-amplification model equation that blowup in finite-time. In addition to knowing the Γblowup\Gamma_{\mathrm{blowup}} is nonempty, we also know that Γblowup\Gamma_{\mathrm{blowup}} is bounded below in H˙12\dot{H}^{-\frac{1}{2}}, because Theorem 4.6 states that there is global regularity for solutions of the strain self-amplification model equation with small initial data in H˙12\dot{H}^{-\frac{1}{2}}, and Theorem 5.3 requires that all of the solutions with initial data in Γblowup\Gamma_{\mathrm{blowup}} must blowup in finite-time. This can also be shown directly by computation using the relevant Sobolev embeddings along with Hölder’s inequality. In addition, we have a lower bound on the amount of planar stretching for SΓblowupS\in\Gamma_{\mathrm{blowup}} in the form of a lower bound on λ2+\lambda_{2}^{+} in the scale critical Lebesgue space.

Proposition 5.8.

For all SΓblowupS\in\Gamma_{\mathrm{blowup}},

λ2+L32>92(π2)43.\left\|\lambda_{2}^{+}\right\|_{L^{\frac{3}{2}}}>\frac{9}{2}\left(\frac{\pi}{2}\right)^{\frac{4}{3}}. (5.63)
Proof.

We will prove the contrapositive. Suppose SHst1S\in H^{1}_{st} with

λ2+L3292(π2)43.\left\|\lambda_{2}^{+}\right\|_{L^{\frac{3}{2}}}\leq\frac{9}{2}\left(\frac{\pi}{2}\right)^{\frac{4}{3}}. (5.64)

We will begin by observing that because tr(S)=0\operatorname*{tr}(S)=0, we have λ10\lambda_{1}\leq 0 and λ30\lambda_{3}\geq 0, because three positive (respectively negative) eigenvalues would violate the trace free condition. This implies that λ1λ30-\lambda_{1}\lambda_{3}\geq 0. Therefore, we can compute that

det(S)\displaystyle-\det(S) =(λ1λ3)λ2\displaystyle=(-\lambda_{1}\lambda_{3})\lambda_{2} (5.65)
(λ1λ3)λ2+\displaystyle\leq(-\lambda_{1}\lambda_{3})\lambda_{2}^{+} (5.66)
12(λ12+λ22)λ2+\displaystyle\leq\frac{1}{2}\left(\lambda_{1}^{2}+\lambda_{2}^{2}\right)\lambda_{2}^{+} (5.67)
12λ2+|S|2.\displaystyle\leq\frac{1}{2}\lambda_{2}^{+}|S|^{2}. (5.68)

Applying this estimate, Hölder’s inequality, and the Sobolev inequality, we find that

3SH˙124det(S)\displaystyle-3\|S\|_{\dot{H}^{1}}^{2}-4\int\det(S) 3SH˙12+2λ2+|S|2\displaystyle\leq-3\|S\|_{\dot{H}^{1}}^{2}+2\int\lambda_{2}^{+}|S|^{2} (5.69)
3SH˙12+2λ2+L32SL62\displaystyle\leq-3\|S\|_{\dot{H}^{1}}^{2}+2\left\|\lambda_{2}^{+}\right\|_{L^{\frac{3}{2}}}\|S\|_{L^{6}}^{2} (5.70)
3SH˙12+23(2π)43λ2+L32SH˙12\displaystyle\leq-3\|S\|_{\dot{H}^{1}}^{2}+\frac{2}{3}\left(\frac{2}{\pi}\right)^{\frac{4}{3}}\left\|\lambda_{2}^{+}\right\|_{L^{\frac{3}{2}}}\|S\|_{\dot{H}^{1}}^{2} (5.71)
0.\displaystyle\leq 0. (5.72)

Therefore we can see that SΓblowupS\notin\Gamma_{\mathrm{blowup}}, and this completes the proof. ∎

6 A perturbative blowup condition for the full Navier–Stokes equation

In this section, we will prove a perturbative condition for blowup, and we will also show that this perturbative condition is satisfied at least for short times. We begin by proving Theorem 1.14, which is restated for the reader’s convenience.

Theorem 6.1.

Suppose uC([0,Tmax);Hdf2)u\in C\left([0,T_{max});H^{2}_{df}\right) is a mild solution of the Navier–Stokes equation such that

f0:=3S0H˙124det(S0)>0,f_{0}:=-3\left\|S^{0}\right\|_{\dot{H}^{1}}^{2}-4\int\det\left(S^{0}\right)>0, (6.1)

and for all 0<t<Tmax0<t<T_{max}

Pst((u)S+13S2+14ωω)(,t)L2(ΔS+Pst(12(u)S+56S2+18ωω))(,t)L22.\frac{\left\|P_{st}\left((u\cdot\nabla)S+\frac{1}{3}S^{2}+\frac{1}{4}\omega\otimes\omega\right)(\cdot,t)\right\|_{L^{2}}}{\left\|\left(-\Delta S+P_{st}\left(\frac{1}{2}(u\cdot\nabla)S+\frac{5}{6}S^{2}+\frac{1}{8}\omega\otimes\omega\right)\right)(\cdot,t)\right\|_{L^{2}}}\leq 2. (6.2)

Then there is finite-time blowup with

Tmax<T:=E0+E02+f0K0f0.T_{max}<T_{*}:=\frac{-E_{0}+\sqrt{E_{0}^{2}+f_{0}K_{0}}}{f_{0}}. (6.3)
Proof.

We will begin by letting

f(t)=3S(,t)H˙124det(S)(,t)f(t)=-3\left\|S(\cdot,t)\right\|_{\dot{H}^{1}}^{2}-4\int\det\left(S\right)(\cdot,t) (6.4)

We know that tr(S)=0\operatorname*{tr}(S)=0, and that therefore det(S)=13tr(S3)\det(S)=\frac{1}{3}\int\operatorname*{tr}\left(S^{3}\right). Therefore we can see that

f(t)=3SH˙1243tr(S3)f(t)=-3\|S\|_{\dot{H}^{1}}^{2}-\frac{4}{3}\int\operatorname*{tr}\left(S^{3}\right) (6.5)

Differentiating ff, we find that for all 0<t<Tmax0<t<T_{max},

ddtf(t)=6ΔS+23Pst(S2),ΔS+Pst((u)S+S2+14ωω)\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}f(t)=6\left<\Delta S+\frac{2}{3}P_{st}\left(S^{2}\right),-\Delta S+P_{st}\left((u\cdot\nabla)S+S^{2}+\frac{1}{4}\omega\otimes\omega\right)\right> (6.6)

Observe that for any M,QL2M,Q\in L^{2}

M,M+Q=M+12QL2214QL22,\left<M,M+Q\right>=\left\|M+\frac{1}{2}Q\right\|_{L^{2}}^{2}-\frac{1}{4}\|Q\|_{L^{2}}^{2}, (6.7)

and so letting M=ΔS+23Pst(S2),Q=Pst((u)S+13S2+14ωω)M=-\Delta S+\frac{2}{3}P_{st}\left(S^{2}\right),Q=P_{st}\left((u\cdot\nabla)S+\frac{1}{3}S^{2}+\frac{1}{4}\omega\otimes\omega\right), we find that for all 0<t<Tmax0<t<T_{max},

ddtf(t)=6ΔS+Pst(12(u)S+56S2+14ωω)L2232Pst((u)S+13S2+14ωω)L22.\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}f(t)=6\left\|-\Delta S+P_{st}\left(\frac{1}{2}(u\cdot\nabla)S+\frac{5}{6}S^{2}+\frac{1}{4}\omega\otimes\omega\right)\right\|_{L^{2}}^{2}\\ -\frac{3}{2}\left\|P_{st}\left((u\cdot\nabla)S+\frac{1}{3}S^{2}+\frac{1}{4}\omega\otimes\omega\right)\right\|_{L^{2}}^{2}. (6.8)

Applying the perturbative condition (6.2), we find that for all 0<t<Tmax0<t<T_{max},

ddtf(t)0,\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}f(t)\geq 0, (6.9)

and therefore, for all 0<t<Tmax0<t<T_{max},

f(t)f0.f(t)\geq f_{0}. (6.10)

Using the identity for enstrophy growth we find that for all 0<t<Tmax0<t<T_{max},

ddtE(t)\displaystyle\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}E(t) =2SH˙124det(S)\displaystyle=-2\|S\|_{\dot{H}^{1}}^{2}-4\int\det(S) (6.11)
>3SH˙124det(S)\displaystyle>-3\|S\|_{\dot{H}^{1}}^{2}-4\int\det(S) (6.12)
=f(t)\displaystyle=f(t) (6.13)
f0.\displaystyle\geq f_{0}. (6.14)

Integrating this differential inequality we find that for all 0<t<Tmax0<t<T_{max},

E(t)>E0+f0t,E(t)>E_{0}+f_{0}t, (6.15)

and integrating this lower bound for enstrophy growth, we find that for all 0<t<Tmax0<t<T_{max},

0tE(τ)dτ>E0t+12f0t2\int_{0}^{t}E(\tau)\mathop{}\!\mathrm{d}\tau>E_{0}t+\frac{1}{2}f_{0}t^{2} (6.16)

Now suppose towards contradiction that TmaxTT_{max}\geq T_{*}. Using the definition

T=E0+E02+f0K0f0,T_{*}=\frac{-E_{0}+\sqrt{E_{0}^{2}+f_{0}K_{0}}}{f_{0}}, (6.17)

we find that

0TE(t)dt\displaystyle\int_{0}^{T_{*}}E(t)\mathop{}\!\mathrm{d}t >E0T+12f0T2\displaystyle>E_{0}T_{*}+\frac{1}{2}f_{0}T_{*}^{2} (6.18)
=12K0.\displaystyle=\frac{1}{2}K_{0}. (6.19)

However this contradicts the bound from the energy equality, which requires that

0TE(t)dt12K0.\int_{0}^{T_{*}}E(t)\mathop{}\!\mathrm{d}t\leq\frac{1}{2}K_{0}. (6.20)

Therefore we may conclude that Tmax<TT_{max}<T_{*}, and this complete the proof. ∎

We cannot show that the perturbative condition (6.2) is satisfied up until TT_{*}—if we could this would resolve the Navier–Stokes regularity problem by proving the existence of finite-time blowup. We can however, show that it is satisfied for short times. The first step will be to show that it holds at the level of initial data.

Proposition 6.2.

There exists SHst2H˙st1S\in H^{2}_{st}\cap\dot{H}^{-1}_{st}, such that

3SH˙124det(S)>0-3\|S\|_{\dot{H}^{1}}^{2}-4\int\det(S)>0 (6.21)

and

Pst((u)S+13S2+14ωω)L2ΔS+Pst(12(u)S+56S2+18ωω)L2<2.\frac{\left\|P_{st}\left((u\cdot\nabla)S+\frac{1}{3}S^{2}+\frac{1}{4}\omega\otimes\omega\right)\right\|_{L^{2}}}{\left\|-\Delta S+P_{st}\left(\frac{1}{2}(u\cdot\nabla)S+\frac{5}{6}S^{2}+\frac{1}{8}\omega\otimes\omega\right)\right\|_{L^{2}}}<2. (6.22)
Proof.

Begin by taking MHst2H˙st1M\in H^{2}_{st}\cap\dot{H}^{-1}_{st} such that

3MH˙124det(M)>0,-3\|M\|_{\dot{H}^{1}}^{2}-4\int\det(M)>0, (6.23)

and QHst2H˙st1Q\in H^{2}_{st}\cap\dot{H}^{-1}_{st} not identically zero. For all λ>0\lambda>0, let

Qλ(x)\displaystyle Q^{\lambda}(x) =Q(λx)\displaystyle=Q(\lambda x) (6.24)
Sλ\displaystyle S^{\lambda} =M+Qλ.\displaystyle=M+Q^{\lambda}. (6.25)

It is a simple computation to observe that

QλH˙1\displaystyle\left\|Q^{\lambda}\right\|_{\dot{H}^{1}} =λ12QH˙1\displaystyle=\lambda^{-\frac{1}{2}}\left\|Q\right\|_{\dot{H}^{1}} (6.26)
QλL3\displaystyle\left\|Q^{\lambda}\right\|_{L^{3}} =λ1QL3\displaystyle=\lambda^{-1}\left\|Q\right\|_{L^{3}} (6.27)

Therefore we can see that

limλ+Qλ=0\lim_{\lambda\to+\infty}Q^{\lambda}=0 (6.28)

in both H˙1\dot{H}^{1} and L3L^{3}. This implies that

limλ+3SλH˙124det(Sλ)\displaystyle\lim_{\lambda\to+\infty}-3\left\|S^{\lambda}\right\|_{\dot{H}^{1}}^{2}-4\int\det\left(S^{\lambda}\right) =3MH˙124det(M)\displaystyle=-3\left\|M\right\|_{\dot{H}^{1}}^{2}-4\int\det\left(M\right) (6.29)
>0,\displaystyle>0, (6.30)

so SλS^{\lambda} satisfies (6.21) for sufficiently large λ>0\lambda>0.

Now take

v\displaystyle v =2div(Δ)1M\displaystyle=-2\operatorname*{div}(-\Delta)^{-1}M (6.31)
w\displaystyle w =2div(Δ)1Q.\displaystyle=-2\operatorname*{div}(-\Delta)^{-1}Q. (6.32)

Note that we then have

M\displaystyle M =symv\displaystyle=\nabla_{sym}v (6.33)
Q\displaystyle Q =symw.\displaystyle=\nabla_{sym}w. (6.34)

Likewise we will take

a\displaystyle a =×v\displaystyle=\nabla\times v (6.35)
b\displaystyle b =×w.\displaystyle=\nabla\times w. (6.36)

Finally we will let

wλ(x)\displaystyle w^{\lambda}(x) =λ1w(λx)\displaystyle=\lambda^{-1}w(\lambda x) (6.37)
bλ(x)\displaystyle b^{\lambda}(x) =b(λx),\displaystyle=b(\lambda x), (6.38)

noting that this implies that

Qλ\displaystyle Q^{\lambda} =symwλ\displaystyle=\nabla_{sym}w^{\lambda} (6.39)
bλ\displaystyle b^{\lambda} =×wλ.\displaystyle=\nabla\times w^{\lambda}. (6.40)

Going back to our linear combination, we can see that

uλ=v+wλ\displaystyle u^{\lambda}=v+w^{\lambda} (6.41)
ωλ=a+bλ.\displaystyle\omega^{\lambda}=a+b^{\lambda}. (6.42)

Applying the triangle inequality we can see that

Pst((uλ)Sλ+13(Sλ)2+14ωλωλ)L2Pst((v)M+13M2+14aa)L2+Pst((wλ)Qλ+13(Qλ)2+14bλbλ)L2+Pst((wλ)M+(v)Qλ+13(QλM+MQλ)+14(bλa+abλ))L2,\left\|P_{st}\left((u^{\lambda}\cdot\nabla)S^{\lambda}+\frac{1}{3}\left(S^{\lambda}\right)^{2}+\frac{1}{4}\omega^{\lambda}\otimes\omega^{\lambda}\right)\right\|_{L^{2}}\leq\left\|P_{st}\left((v\cdot\nabla)M+\frac{1}{3}M^{2}+\frac{1}{4}a\otimes a\right)\right\|_{L^{2}}\\ +\left\|P_{st}\left((w^{\lambda}\cdot\nabla)Q^{\lambda}+\frac{1}{3}\left(Q^{\lambda}\right)^{2}+\frac{1}{4}b^{\lambda}\otimes b^{\lambda}\right)\right\|_{L^{2}}\\ +\left\|P_{st}\left((w^{\lambda}\cdot\nabla)M+(v\cdot\nabla)Q^{\lambda}+\frac{1}{3}\left(Q^{\lambda}M+MQ^{\lambda}\right)+\frac{1}{4}\left(b^{\lambda}\otimes a+a\otimes b^{\lambda}\right)\right)\right\|_{L^{2}}, (6.43)

and applying Hölder’s inequality and our scaling laws from above, we can conclude

Pst((uλ)Sλ+13(Sλ)2+14ωλωλ)L2Pst((v)M+13M2+14aa)L2+λ32Pst((w)Q+13Q2+14bb)L2+λ1wLML2+λ12vLQL2+23λ34ML4QL4+12λ34aL4bL4.\left\|P_{st}\left((u^{\lambda}\cdot\nabla)S^{\lambda}+\frac{1}{3}\left(S^{\lambda}\right)^{2}+\frac{1}{4}\omega^{\lambda}\otimes\omega^{\lambda}\right)\right\|_{L^{2}}\leq\left\|P_{st}\left((v\cdot\nabla)M+\frac{1}{3}M^{2}+\frac{1}{4}a\otimes a\right)\right\|_{L^{2}}\\ +\lambda^{-\frac{3}{2}}\left\|P_{st}\left((w\cdot\nabla)Q+\frac{1}{3}Q^{2}+\frac{1}{4}b\otimes b\right)\right\|_{L^{2}}+\lambda^{-1}\|w\|_{L^{\infty}}\|\nabla M\|_{L^{2}}\\ +\lambda^{-\frac{1}{2}}\|v\|_{L^{\infty}}\|\nabla Q\|_{L^{2}}+\frac{2}{3}\lambda^{-\frac{3}{4}}\|M\|_{L^{4}}\|Q\|_{L^{4}}+\frac{1}{2}\lambda^{-\frac{3}{4}}\|a\|_{L^{4}}\|b\|_{L^{4}}. (6.44)

Likewise we may compute that

ΔSλ+Pst(12(uλ)Sλ+56(Sλ)2+18ωλωλ)L2λ12ΔQL2λ32Pst(12(w)Q+56Q2+18bb)L2ΔM+Pst(12(v)M+56M2+18aa)L253λ34ML4QL414λ34aL4bL412λ12vLQL212λ1wLML2.\left\|-\Delta S^{\lambda}+P_{st}\left(\frac{1}{2}(u^{\lambda}\cdot\nabla)S^{\lambda}+\frac{5}{6}\left(S^{\lambda}\right)^{2}+\frac{1}{8}\omega^{\lambda}\otimes\omega^{\lambda}\right)\right\|_{L^{2}}\geq\lambda^{\frac{1}{2}}\|-\Delta Q\|_{L^{2}}\\ -\lambda^{-\frac{3}{2}}\left\|P_{st}\left(\frac{1}{2}(w\cdot\nabla)Q+\frac{5}{6}Q^{2}+\frac{1}{8}b\otimes b\right)\right\|_{L^{2}}\\ -\left\|-\Delta M+P_{st}\left(\frac{1}{2}(v\cdot\nabla)M+\frac{5}{6}M^{2}+\frac{1}{8}a\otimes a\right)\right\|_{L^{2}}\\ -\frac{5}{3}\lambda^{-\frac{3}{4}}\|M\|_{L^{4}}\|Q\|_{L^{4}}-\frac{1}{4}\lambda^{-\frac{3}{4}}\|a\|_{L^{4}}\|b\|_{L^{4}}\\ -\frac{1}{2}\lambda^{-\frac{1}{2}}\|v\|_{L^{\infty}}\|\nabla Q\|_{L^{2}}-\frac{1}{2}\lambda^{-1}\|w\|_{L^{\infty}}\|\nabla M\|_{L^{2}}. (6.45)

Putting together the inequalities in (6.44) and (6.45), we find that

limλ+Pst((uλ)Sλ+13(Sλ)2+14ωλωλ)L2ΔSλ+Pst(12(uλ)Sλ+56(Sλ)2+18ωλωλ)L2=0,\lim_{\lambda\to+\infty}\frac{\left\|P_{st}\left((u^{\lambda}\cdot\nabla)S^{\lambda}+\frac{1}{3}\left(S^{\lambda}\right)^{2}+\frac{1}{4}\omega^{\lambda}\otimes\omega^{\lambda}\right)\right\|_{L^{2}}}{\left\|-\Delta S^{\lambda}+P_{st}\left(\frac{1}{2}(u^{\lambda}\cdot\nabla)S^{\lambda}+\frac{5}{6}\left(S^{\lambda}\right)^{2}+\frac{1}{8}\omega^{\lambda}\otimes\omega^{\lambda}\right)\right\|_{L^{2}}}=0, (6.46)

and so in particular for sufficiently large λ>0\lambda>0,

Pst((uλ)Sλ+13(Sλ)2+14ωλωλ)L2ΔSλ+Pst(12(uλ)Sλ+56(Sλ)2+18ωλωλ)L2<2.\frac{\left\|P_{st}\left((u^{\lambda}\cdot\nabla)S^{\lambda}+\frac{1}{3}\left(S^{\lambda}\right)^{2}+\frac{1}{4}\omega^{\lambda}\otimes\omega^{\lambda}\right)\right\|_{L^{2}}}{\left\|-\Delta S^{\lambda}+P_{st}\left(\frac{1}{2}(u^{\lambda}\cdot\nabla)S^{\lambda}+\frac{5}{6}\left(S^{\lambda}\right)^{2}+\frac{1}{8}\omega^{\lambda}\otimes\omega^{\lambda}\right)\right\|_{L^{2}}}<2. (6.47)

This completes the proof. ∎

Now that we have established that the perturbative condition (6.2) can hold for initial data, it is straightforward to show that it can hold for at least short times by continuity. This result is Theorem 1.18, which is restated for the reader’s convenience.

Theorem 6.3.

There exists a mild solution of the Navier–Stokes equation uC([0,Tmax);Hdf3)u\in C\left([0,T_{max});H^{3}_{df}\right) and ϵ>0\epsilon>0, such that

3S0H˙124det(S0)>0-3\left\|S^{0}\right\|_{\dot{H}^{1}}^{2}-4\int\det\left(S^{0}\right)>0 (6.48)

and for all 0t<ϵ0\leq t<\epsilon

Pst((u)S+13S2+14ωω)(,t)L2(ΔS+Pst(12(u)S+56S2+18ωω))(,t)L2<2.\frac{\left\|P_{st}\left((u\cdot\nabla)S+\frac{1}{3}S^{2}+\frac{1}{4}\omega\otimes\omega\right)(\cdot,t)\right\|_{L^{2}}}{\left\|\left(-\Delta S+P_{st}\left(\frac{1}{2}(u\cdot\nabla)S+\frac{5}{6}S^{2}+\frac{1}{8}\omega\otimes\omega\right)\right)(\cdot,t)\right\|_{L^{2}}}<2. (6.49)
Proof.

Fix initial data S0Hst2H˙st1S^{0}\in H^{2}_{st}\cap\dot{H}^{-1}_{st} such that

3S0H˙124det(S0)>0-3\left\|S^{0}\right\|_{\dot{H}^{1}}^{2}-4\int\det\left(S^{0}\right)>0 (6.50)

and

Pst((u0)S0+13(S0)2+14ω0ω0)L2ΔS0+Pst(12(u0)S0+56(S0)2+18ω0ω0)L2<2,\frac{\left\|P_{st}\left((u^{0}\cdot\nabla)S^{0}+\frac{1}{3}\left(S^{0}\right)^{2}+\frac{1}{4}\omega^{0}\otimes\omega^{0}\right)\right\|_{L^{2}}}{\left\|-\Delta S^{0}+P_{st}\left(\frac{1}{2}(u^{0}\cdot\nabla)S^{0}+\frac{5}{6}\left(S^{0}\right)^{2}+\frac{1}{8}\omega^{0}\otimes\omega^{0}\right)\right\|_{L^{2}}}<2, (6.51)

where S0=symu0S^{0}=\nabla_{sym}u^{0} and ω0=×u0\omega^{0}=\nabla\times u^{0}. Note that u0Hdf3u^{0}\in H^{3}_{df} by definition, and is given by u0=2div(Δ)1S0u^{0}=-2\operatorname*{div}(-\Delta)^{-1}S^{0}. Let uC([0,Tmax);Hdf3)u\in C\left([0,T_{max});H^{3}_{df}\right) be the unique mild solution of the Navier–Stokes equation with initial data u0u^{0}. Next we will let

c(t)=Pst((u)S+13S2+14ωω)(,t)L22(ΔS+Pst(12(u)S+56S2+18ωω))(,t)L2.c(t)=\left\|P_{st}\left((u\cdot\nabla)S+\frac{1}{3}S^{2}+\frac{1}{4}\omega\otimes\omega\right)(\cdot,t)\right\|_{L^{2}}\\ -2\left\|\left(-\Delta S+P_{st}\left(\frac{1}{2}(u\cdot\nabla)S+\frac{5}{6}S^{2}+\frac{1}{8}\omega\otimes\omega\right)\right)(\cdot,t)\right\|_{L^{2}}. (6.52)

The fact that uC([0,Tmax);Hdf3)u\in C\left([0,T_{max});H^{3}_{df}\right) immediately implies that cC([0,Tmax))c\in C\left([0,T_{max})\right). We also know by hypothesis that

c(0)<0,c(0)<0, (6.53)

so by continuity there must exist ϵ>0\epsilon>0 such that for all 0t<ϵ0\leq t<\epsilon,

c(t)<0.c(t)<0. (6.54)

This completes the proof. ∎

Remark 6.4.

The key to the proof of Proposition 6.2 and Theorem 6.3 rests on the fact that we can add a perturbative term which is small in both H˙1\dot{H}^{1} and L3L^{3}, leaving (6.21) essentially unaffected, but is very large in H˙2\dot{H}^{2}, making the denominator in (6.22) as large as we like. The key is to add a perturbative term that is supported at very high Fourier modes, but with a scaling chosen so that the perturbation remains small in H˙1L3\dot{H}^{1}\cap L^{3}.

Remark 6.5.

There are axisymmetric, swirl free initial data that satisfy Theorem 6.3. To see this, in the context of Proposition 6.2, take MM as in Proposition 5.4 and QQ to be an arbitrary axisymmetric, swirl free strain matrix and the result follows. In this case, however, we know that the perturbative condition can only be satisfied for short times, because there is global regularity for axisymmetric, swirl-free solutions of the Navier–Stokes equation. There is something in the geometry of axisymmetric swirl-free solutions that, when coupled with the dynamics of the equation, guarantees the perturbative condition will fail after short times.

Corollary 6.6.

Suppose uC([0,+);Hdf2)u\in C\left([0,+\infty);H^{2}_{df}\right) is an axisymmetric, swirl-free, mild solution of the Navier–Stokes equation such that

f0:=3S0H˙124det(S0)>0.f_{0}:=-3\left\|S^{0}\right\|_{\dot{H}^{1}}^{2}-4\int\det\left(S^{0}\right)>0. (6.55)

Then there exists 0<t<T:=E0+E02+f0K0f00<t<T^{*}:=\frac{-E_{0}+\sqrt{E_{0}^{2}+f_{0}K_{0}}}{f_{0}}, such that

Pst((u)S+13S2+14ωω)(,t)L2(ΔS+Pst(12(u)S+56S2+18ωω))(,t)L2>2.\frac{\left\|P_{st}\left((u\cdot\nabla)S+\frac{1}{3}S^{2}+\frac{1}{4}\omega\otimes\omega\right)(\cdot,t)\right\|_{L^{2}}}{\left\|\left(-\Delta S+P_{st}\left(\frac{1}{2}(u\cdot\nabla)S+\frac{5}{6}S^{2}+\frac{1}{8}\omega\otimes\omega\right)\right)(\cdot,t)\right\|_{L^{2}}}>2. (6.56)
Proof.

Ladyzhenskaya first proved global regularity for solutions of the Navier–Stokes equation with swirl-free, axisymmetric initial data [LadyzhenskayaAxiSym1, LadyzhenskayaAxiSym2]. The corollary follows immediately from Theorem 6.1 and the global regularity of axisymmetric, swirl-free solutions of the Navier–Stokes equation, because if the perturbative condition from Theorem 6.1 was satisfied up until TT^{*}, then there must be finite-time blowup, which we know cannot occur. ∎

Acknowledgements

This publication was supported in part by the Fields Institute for Research in the Mathematical Sciences while the author was in residence during the Fall 2020 semester. Its contents are solely the responsibility of the author and do not necessarily represent the official views of the Fields Institute. This material is based upon work supported by the National Science Foundation under Grant No. DMS-1440140 while the author participated in a program that was hosted by the Mathematical Sciences Research Institute in Berkeley, California, during the Spring 2021 semester. The material in this publication is also based on research conducted by the author while a PhD student at the University of Toronto and a postdoctoral fellow at McMaster University.

References

  • \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry