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

Self-similar source-type solutions to the three-dimensional Navier-Stokes equations

K. Ohkitani
Research Institute for Mathematical Sciences,
Kyoto University, Kyoto 606-8502 Japan.
   R. Vanon
School of Mathematics, Statistics and Physics,
Newcastle University, Newcastle upon Tyne, NE1 7RU, U.K
Abstract

We formalise a systematic method of constructing forward self-similar solutions to the Navier-Stokes equations in order to characterise the late stage of decaying process of turbulent flows. (i) In view of critical scale-invariance of type 2 we exploit the vorticity curl as the dependent variable to derive and analyse the dynamically-scaled Navier-Stokes equations. This formalism offers the viewpoint from which the problem takes the simplest possible form. (ii) Rewriting the scaled Navier-Stokes equations by Duhamel principle as integral equations, we regard the nonlinear term as a perturbation using the Fokker-Planck evolution semigroup. Systematic successive approximations are introduced and the leading-order solution is worked out explicitly as the Gaussian function with a solenoidal projection. (iii) By iterations the second-order approximation is estimated explicitly up to solenoidal projection and is evaluated numerically. (iv) A new characterisation of nonlinear term is introduced on this basis to estimate its strength NN quantitatively. We find that N=O(102)N=O(10^{-2}) for the 3D Navier-Stokes equations. This should be contrasted with N=O(101)N=O(10^{-1}) for the Burgers equations and N0N\equiv 0 for the 2D Navier-Stokes equations. (v) As an illustration we explicitly determine source-type solutions to the multi-dimensional the Burgers equations. Implications and applications of the current results are given.

Navier-Stokes equations,self-similarity, scale-invariance

1 Introduction

Self-similarity is a tool of fundamental importance in analysing partial differential equations, including the construction of solutions and the determination of their stability. It is also useful for numerical and asymptotic methods of studying partial differential equations. For general aspects of self-similarity and its applications we refer the readers to e.g. [1, 2].

In this paper we will study the so-called source-type self-similar solution to the Navier-Stokes equations. Our motivation for this is as follows. First of all, it will give us a particular self-similar solution to the Navier-Stokes equations, which characterises the decaying process in the late stage of evolution. Second, it is likely to give useful information as to how we may handle more general solutions.

It may be in order to have a look at previous works which are related to this paper. It has been shown under mild conditions that no nontrivial smooth backward self-similar solution exists to the Navier-Stokes equations. On the other hand it is known that nontrivial forward self-similar solutions do exist, but their explicit functional forms are not known, except for some asymptotic results. It is of interest to see how they actually behave because such solutions contain important information regarding more general solutions. This is particularly the case when the governing equations are exactly linearisable, e.g. the Burgers equations. While it is not expected that the Navier-Stokes equations are exactly soluble in general, we might still obtain insights into the nature of their solutions.

In [3] the existence of forward self-similar solutions for small data was proven by using a fixed-point theorem in a Besov space (see below). There, initial data for the self-similar solution (3D) are assumed to be homogeneous of degree 1-1 in velocity

𝒖0(λ𝒙)=λ1𝒖0(𝒙)\bm{u}_{0}(\lambda\bm{x})=\lambda^{-1}\bm{u}_{0}(\bm{x})

and the existence of a self-similar solution of the form

𝒖(𝒙,t)=1t𝑼(𝒙t)\bm{u}(\bm{x},t)=\frac{1}{\sqrt{t}}\bm{U}\left(\frac{\bm{x}}{\sqrt{t}}\right)

has been established under the assumption that initial data is small in some Besov space. Moreover, it has been proved that the self-similar profile UU satisfies (in their notations)

U=S(1)u0+W,U=S(1)u_{0}+W,

where S(1)S(1) denotes a heat operator at time 1 and WL3\|W\|_{L^{3}} is small. In [4] using a locally Hölder class in 3\{0}\mathbb{R}^{3}\backslash\{0\} the smallness assumption has been removed and it is furthermore shown that

|U(x)eu0(x)|C(M)(1+|x|)1+α,|U(x)-e^{\triangle}u_{0}(x)|\leq\frac{C(M)}{(1+|x|)^{1+\alpha}},

where 0<α<10<\alpha<1 and C(M)C(M) denotes a constant with some norm MM of u0u_{0}. Those studies indicate that the self-similar solution is close to the heat flow in the late stage. However, studies on the determination of a specific functional form of self-similar solutions are few and far between, except for an attempt in[5]. We also note that the existence of generalised self-similar solutions (in the sense that the scaling holds only at a set of discrete values of λ\lambda) was studied subsequently, e.g. [7, 6].

Our basic strategy in practice is as follows. After recasting the dynamically-scaled Navier-Stokes equations as integral equations via the Duhamel principle, we regard the nonlinear term as a perturbation using the Fokker-Planck evolution semigroup. Systematic successive approximations are then introduced and the first-order solution is worked out explicitly as the Gaussian function with a solenoidal projection. The second-order approximation is also evaluated numerically to assess the strength of the nonlinear term.

We will construct an approximate solution to the 3D Navier-Stokes equations valid in the long-time limit, using the vorticity curl ×𝝎\nabla\times\bm{\omega}. In Sections 4 we will see why this is the most convenient variable from the reaction of the Navier-Stokes equations under dynamic scaling.

Here we appreciate the suitability of such a choice of the unknown by comparing the source-type solutions to the Navier-Stokes equations and their self-similar solutions at criticality. By definition the source-type solution for nonlinear parabolic PDEs is a solution in a scaled space, which starts from the Dirac mass in some dependent variable and ends up like a near-identity of the Gaussian function in the long-time limit. It serves as an analogue of the fundamental solution to nonlinear PDEs.

The unknown whose L1L^{1}-norm is marginally divergent is suitable for describing the late-stage evolution. This is because this self-similar solution satisfies the same scaling as the Dirac mass and both of them belong to a Besov space near L1L^{1}. In one dimension, in the limit of t0t\to 0, we have roughly

u1x,marginallyL1(1),u\sim\dfrac{1}{x},\;\mbox{marginally}\,\notin L^{1}(\mathbb{R}^{1}),

which suggests that the velocity is convenient in this case.

In two dimensions it is the vorticity which is the most convenient, as can be seen from

ω1r2,marginallyL1(2),\omega\sim\dfrac{1}{r^{2}},\;\mbox{marginally}\,\notin L^{1}(\mathbb{R}^{2}),

where |𝒙|=r|\bm{x}|=r. Recall that those scaling properties of velocity in 1D or vorticity in 2D, are the same as that of the Dirac mass; λdδ(λ𝒙)=δ(𝒙)\lambda^{d}\delta(\lambda\bm{x})=\delta(\bm{x}) in dd-dimensions.

Now consider Besov spaces whose norms are given by

𝒖Bpqs{j=1(2sjΔj(𝒖)Lp)q}1/q,\|\bm{u}\|_{B^{s}_{pq}}\equiv\left\{\sum_{j=1}^{\infty}\left(2^{sj}\|\Delta_{j}(\bm{u})\|_{L^{p}}\right)^{q}\right\}^{1/q},

where 1p,q,s1\leq p,q\leq\infty,s\in\mathbb{R} and Δj(𝒖)\Delta_{j}(\bm{u}) represents band-filtered velocity at frequency 2j2^{j}. It is known that in d\mathbb{R}^{d} the Dirac delta mass is embedded as

δ(𝒙)Bp,d+d/p,forp1,\delta(\bm{x})\in B_{p,\infty}^{-d+d/p},\;\;\mbox{for}\;\;p\geq 1, (1)

see e.g. [8]. In particular we have δ(𝒙)B1,0\delta(\bm{x})\in B_{1,\infty}^{0} for any d.d.

While the velocity u1rL3(3),u\sim\dfrac{1}{r}\notin L^{3}(\mathbb{R}^{3}), we have

u1rB3,0(3),u\sim\frac{1}{r}\in B_{3,\infty}^{0}(\mathbb{R}^{3}),

and correspondingly for 𝝌=×𝝎,\bm{\chi}=\nabla\times\bm{\omega}, the vorticity curl,

χ1r3B1,0(3).\chi\sim\frac{1}{r^{3}}\in B_{1,\infty}^{0}(\mathbb{R}^{3}).

Hence in three dimensions this χ\chi and the Dirac mass belong to the same function class B1,0(3)B_{1,\infty}^{0}(\mathbb{R}^{3}), with p=1p=1 in (1).

The rest of this paper is organised as follows. In Section 2 after reviewing critical scale-invariance of type 2 (to be defined below) using the Burgers equation we introduce successive approximations of determining the self-similar profile. On this basis we introduce and quantify the strength of nonlinearity. Higher-dimensional Burgers equations are also discussed. In Section 3 we have a brief look at the 2D Navier-Stokes equations. In the main Section 4 we study the self-similar solutions of the 3D Navier-Stokes equations utilising the vorticity curl as the unknown to achieve critical scale-invariance of type 2. We carry out successive approximations and determine the strength of nonlinearity as introduced above. Section 5 will be devoted to the Summary and outlook. Some further details and derivations are given in Appendices.

2 Burgers equations

We review the source-type solution of the Burgers equation with an emphasis on the critical scale-invariance. Our approach is novel in the introduction of its successive approximations, in preparation for handling the 3D Navier-Stokes equations, and in employing a new method of estimating the strength of nonlinearity on this basis. Also described are the source-type solutions of the Burgers equations in nn-dimensions.

2.1 Critical scale-invariance

We consider the Burgers equation [9]

ut+uux=ν2ux2,\frac{\partial u}{\partial t}+u\frac{\partial u}{\partial x}=\nu\frac{\partial^{2}u}{\partial x^{2}}, (2)

which satisfies static scale-invariance under

xλx,tλ2t,uλ1u.x\to\lambda x,t\to\lambda^{2}t,u\to\lambda^{-1}u.

This means that if u(x,t)u(x,t) is a solution, so is uλ(x,t)λu(λx,λ2t),u_{\lambda}(x,t)\equiv\lambda u(\lambda x,\lambda^{2}t), for any λ(>0)\lambda(>0). It is readily checked that

uλLp=λp1puLp,\|u_{\lambda}\|_{L^{p}}=\lambda^{\frac{p-1}{p}}\|u\|_{L^{p}},

which shows that the L1L^{1}-norm is scale-invariant.

Let us clarify the two kinds of critical scale-invariance. Type 1 scale-invariance is achieved when we use a dependent variable whose physical dimension is the same as ν\nu. Type 1 is deterministic in nature where the additional term arising in the governing equations under dynamic scaling is minimised in number. Type 2 instead is statistical in nature where the additional terms under dynamic scaling are maximised in number so that a divergence form is completed and the dynamically-scaled equations have the Fokker-Planck operator as the linearisation. In the former the dependent variable has the same physical dimension as kinematic viscosity, whereas in the latter the argument of the Hopf characteristic functional (the independent variable) has the same physical dimension as the reciprocal of kinematic viscosity [15]. This approach provides a viewpoint from which the problem appears in the simplest possible form.

Critical scale-invariance of type 1 is achieved with the velocity potential ϕ,\phi, which is defined by u=xϕu=\partial_{x}\phi. If ϕ(x,t)\phi(x,t) is a solution, so is ϕ(λx,λ2t).\phi(\lambda x,\lambda^{2}t). Under dynamic scaling for the velocity potential ϕ(x,t)=Φ(ξ,τ),\phi(x,t)=\Phi(\xi,\tau), ξ=x2at,τ=12alogt\xi=\frac{x}{\sqrt{2at}},\;\tau=\frac{1}{2a}\log t we have

Φτ+12(Φξ)2=aξΦξ+ν2Φξ2,\frac{\partial\Phi}{\partial\tau}+\frac{1}{2}\left(\frac{\partial\Phi}{\partial\xi}\right)^{2}=a\xi\frac{\partial\Phi}{\partial\xi}+\nu\frac{\partial^{2}\Phi}{\partial\xi^{2}}, (3)

whose linearisation has the Ornstein-Uhlenbeck operator. This is called type 1 (deterministic) scale-invariance where the number of additional terms is minimised, that is, only the drift term remains. Under dynamic scaling for velocity u(x,t)=12atU(ξ,τ)u(x,t)=\frac{1}{\sqrt{2at}}U(\xi,\tau), we find

Uτ+UUξ=aξ(ξU)+ν2Uξ2,\frac{\partial U}{\partial\tau}+U\frac{\partial U}{\partial\xi}=a\frac{\partial}{\partial\xi}\left(\xi U\right)+\nu\frac{\partial^{2}U}{\partial\xi^{2}}, (4)

whose linearisation is the Fokker-Planck equation. Here the zooming-in parameter a(>0)a(>0) has the same physical dimension as ν\nu and is on the same order of it. With this type 2 (statistical) scale-invariance where the number of additional terms is maximised meaning that a divergence form is completed with the addition of aUaU term. As it is a second-order equation it has two independent solutions, of which we will focus on the Gaussian one. See Appendix C for the other non-Gaussian kinds of solutions.

Equation (4) is exactly soluble and its steady solution is called the source-type solution [11, 12]:

U(ξ)=U(0)exp(aξ22ν)1U(0)2ν0ξexp(aη22ν)𝑑η.U(\xi)=\frac{U(0)\displaystyle{\exp\left(-\frac{a\xi^{2}}{2\nu}\right)}}{1-\displaystyle{\frac{U(0)}{2\nu}\int_{0}^{\xi}}\exp\left(-\frac{a\eta^{2}}{2\nu}\right)d\eta}. (5)

The name has come from the time zero asymptotics

limt012atU(ξ)=Mδ(x),\lim_{t\to 0}\frac{1}{\sqrt{2at}}U(\xi)=M\delta(x),

where M1u0(x)𝑑xM\equiv\int_{\mathbb{R}^{1}}u_{0}(x)dx and U(0)=8aνπtanhM4νa2πνM(forM/ν1).U(0)=\sqrt{\frac{8a\nu}{\pi}}\tanh\frac{M}{4\nu}\approx\sqrt{\frac{a}{2\pi\nu}}M\;\;(\mbox{for}\;M/\nu\ll 1). Observe that (5) is a near-identity transformation of the Gaussian function. See [10, 11, 12].

It is also known that for u0L1u_{0}\in L^{1} we have

t12(11p)u(x,t)12atU(ξ)Lp0ast,t^{\frac{1}{2}\left(1-\frac{1}{p}\right)}\left\|u(x,t)-\frac{1}{\sqrt{2at}}U(\xi)\right\|_{L^{p}}\to 0\;\;\mbox{as}\;\;t\to\infty,

where 1p.1\leq p\leq\infty.

The simplest method for solving (4) without linearisation is as follows. Rewrite the equation

U22\displaystyle\frac{U^{2}}{2} =\displaystyle= aξU+νdUdξ\displaystyle a\xi U+\nu\frac{dU}{d\xi}
=\displaystyle= νexp(aξ22ν)ddξ(Uexp(aξ22ν)).\displaystyle\nu\exp\left(-\frac{a\xi^{2}}{2\nu}\right)\frac{d}{d\xi}\left(U\exp\left(\frac{a\xi^{2}}{2\nu}\right)\right).

By changing variables to U~=Uexp(aξ22ν),η=12ν0ξexp(aζ22ν)𝑑ζ,\tilde{U}=U\exp\left(\frac{a\xi^{2}}{2\nu}\right),\eta=\frac{1}{2\nu}\int_{0}^{\xi}\exp\left(-\frac{a\zeta^{2}}{2\nu}\right)d\zeta, we find

dU~dη=U~2,\frac{d\tilde{U}}{d\eta}=\tilde{U}^{2},

which is readily integrable. Alternatively we may solve the equation (2.1) by regarding it as a Bernoulli equation.

It may be in order to comment on the significance of source-type solution. When we recast (5) as

U(ξ)=2νξlog(1U(0)2ν0ξexp(aη22ν)𝑑η),U(\xi)=-2\nu\frac{\partial}{\partial\xi}\log\left({1-\displaystyle{\frac{U(0)}{2\nu}\int_{0}^{\xi}}\exp\left(-\frac{a\eta^{2}}{2\nu}\right)d\eta}\right), (6)

which is reminiscent of the celebrated Cole-Hopf transform. In other words, the source-type solution encodes the vital information of the nonlinear term in the case of the Burgers equation. Note that the error-function itself 0ξexp(aη22ν)𝑑η\int_{0}^{\xi}\exp\left(-\frac{a\eta^{2}}{2\nu}\right)d\eta in (5, 6) is a self-similar solution to the heat equation. This suggests that studying source-type solution of the Navier-Stokes equations may give a hint on how to characterise their long-time evolution by a heat flow.

2.2 Successive approximations

The operator L=+aνξ(ξ)L=\triangle^{*}\equiv\triangle+\frac{a}{\nu}\partial_{\xi}(\xi\cdot) is not self-adjoint. It is possible to find a function GG111With a slight abuse of notation this GG should be distinguished from the Gaussian function used in Section 3. such that LG(ξ)=δ(ξ)L^{\dagger}G(\xi)=-\delta(\xi) holds, where Lξ2aνξξL^{\dagger}\equiv\partial_{\xi}^{2}-\frac{a}{\nu}\xi\partial_{\xi} is the adjoint of LL. In fact G(ξ)D(a2νξ),G(\xi)\propto D\left(\sqrt{\frac{a}{2\nu}}\xi\right), where D()D(\cdot) denotes the Dawson’s integral, defined by D(x)ex20xey2𝑑yD(x)\equiv e^{-x^{2}}\int_{0}^{x}e^{y^{2}}dy [18]. However, because GG decays slowly at large distances G(ξ)1/ξG(\xi)\propto 1/\xi as |ξ||\xi|\to\infty, it cannot be used as a Green’s function, at least, in the usual manner.

The inversion formula for \triangle^{*} can be obtained by an alternative method. Recall that based on 1a=0eat𝑑t(a>0),\frac{1}{a}=\int_{0}^{\infty}e^{-at}dt\;(a>0), the fundamental solution to the Poisson equation in 1D is given by

(ν)10dseνs=|ξ|2ν,(\nu\triangle)^{-1}\equiv-\int_{0}^{\infty}dse^{\nu s\triangle}=\frac{|\xi|}{2\nu}*,

where * denotes convolution. Likewise for the fundamental solution to the Fokker-Planck equation in 1D we write

(ν)10𝑑seνs=𝑑ηg(ξ,η),(\nu\triangle^{*})^{-1}\equiv-\int_{0}^{\infty}dse^{\nu s\triangle^{*}}=\int_{-\infty}^{\infty}d\eta g(\xi,\eta),

where

g(ξ,η)12πνf.p.adσσ2ae12ν(σξησ2a)2g(\xi,\eta)\equiv\frac{-1}{\sqrt{2\pi\nu}}{\rm f.p.}\int_{\sqrt{a}}^{\infty}\frac{d\sigma}{\sigma^{2}-a}e^{-\frac{1}{2\nu}\left(\sigma\xi-\eta\sqrt{\sigma^{2}-a}\right)^{2}}

and f.p. denotes the finite part of Hadamard, e.g. [13, 14]. It can be verified by changing the variable from ss to σ=a1e2aτ,\sigma=\sqrt{\frac{a}{1-e^{-2a\tau}}}, using the solution of the Fokker-Planck equation

eντf=(a2πν(1e2aτ))1/21eaτf(eaτη)exp(a2ν(ξη)21e2aτ)𝑑η.e^{\nu\tau\triangle^{*}}f=\left(\frac{a}{2\pi\nu(1-e^{-2a\tau})}\right)^{1/2}\int_{\mathbb{R}^{1}}e^{a\tau}f(e^{a\tau}\eta)\exp\left(-\frac{a}{2\nu}\frac{(\xi-\eta)^{2}}{1-e^{-2a\tau}}\right)d\eta.

As we will consider the 3D Navier-Stokes equations, for which methods of exact solutions are unavailable, we treat (4) by approximate methods as an illustration. Because the inversion of \triangle^{*} is unwieldy, we will seek a workaround by which we can dispense with it.

First we convert it to an integral equation by the Duhamel principle for the Fokker-Planck operator \triangle^{*}

U(τ)=\displaystyle U(\tau)= eντU(0)0τeν(τs)U(s)22ds\displaystyle e^{\nu\tau\triangle^{*}}U(0)-\int_{0}^{\tau}e^{\nu(\tau-s)\triangle^{*}}\partial\,\frac{U(s)^{2}}{2}ds
=\displaystyle= eντU(0)0τeνsU(τs)22ds.\displaystyle e^{\nu\tau\triangle^{*}}U(0)-\int_{0}^{\tau}e^{\nu s\triangle^{*}}\partial\,\frac{U(\tau-s)^{2}}{2}ds.

The long-time limit U1=limτeνsU(0)U_{1}=\lim_{\tau\to\infty}e^{\nu s\triangle^{*}}U(0) is given by

U1=(a2πν)1/2Mea2νξ2withM=U(0)𝑑ξ.U_{1}=\left(\frac{a}{2\pi\nu}\right)^{1/2}Me^{-\frac{a}{2\nu}\xi^{2}}\;\;\mbox{with}\;\;M=\int_{-\infty}^{\infty}U(0)d\xi.

We may consider a number of different iteration schemes. For example, the following option (1), also known as the Picard iteration, requires the inversion ()1(\triangle^{*})^{-1}:

Successive approximation (1):Un+1=U10eνsUn22ds,\mbox{Successive approximation (1):}\;\;U_{n+1}=U_{1}-\int_{0}^{\infty}e^{\nu s\triangle^{*}}\partial\,\frac{U_{n}^{2}}{2}ds,
in particular, forn=1:U2=U10eνsU122ds.\mbox{in particular, for}\;\;n=1:\;\;U_{2}=U_{1}-\int_{0}^{\infty}e^{\nu s\triangle^{*}}\partial\,\frac{U_{1}^{2}}{2}ds.

Note that UnU_{n} is a steady function at each step.

Alternatively we first consider the steady equation

UU+aν(ξU)ξ=1ν(U22)ξ\triangle^{*}U\equiv\triangle U+\frac{a}{\nu}(\xi U)_{\xi}=\frac{1}{\nu}\left(\frac{U^{2}}{2}\right)_{\xi}

and then introduce iteration schemes:

Iteration scheme (2a):Un+1+aν(ξUn+1)ξ=1ν(Un22)ξ,(n0)\mbox{Iteration scheme (2a):}\;\;\triangle U_{n+1}+\frac{a}{\nu}(\xi U_{n+1})_{\xi}=\frac{1}{\nu}\left(\frac{U_{n}^{2}}{2}\right)_{\xi},\;\;(n\geq 0)
Forn=1:U2+aν(ξU2)ξ=1ν(U122)ξ,\mbox{For}\;\;n=1:\;\;\triangle U_{2}+\frac{a}{\nu}(\xi U_{2})_{\xi}=\frac{1}{\nu}\left(\frac{U_{1}^{2}}{2}\right)_{\xi},

or

Iteration scheme (2b):Un+1=aν(ξUn)ξ+1ν(Un22)ξ,(n1)\mbox{Iteration scheme (2b):}\;\;\triangle U_{n+1}=-\frac{a}{\nu}(\xi U_{n})_{\xi}+\frac{1}{\nu}\left(\frac{U_{n}^{2}}{2}\right)_{\xi},\;\;(n\geq 1)
Forn=1:U2=aν(ξU1)ξ=U1+1ν(U122)ξ.\mbox{For}\;\;n=1:\;\;\triangle U_{2}=\underbrace{-\frac{a}{\nu}(\xi U_{1})_{\xi}}_{=\triangle U_{1}}+\frac{1}{\nu}\left(\frac{U_{1}^{2}}{2}\right)_{\xi}.

Note that iteration schemes (1) and (2a) coincide with each other at n=1n=1.

2.3 Estimation of the strength of nonlinearity

For the Burgers equation we can work out the two kinds of approximations to the second-order analytically. After straightforward algebra they are

(1)U\displaystyle\mbox{(1)}\;\;U Cea2νξ2(1+C2ν0ξea2νη2𝑑η),\displaystyle\approx Ce^{-\frac{a}{2\nu}\xi^{2}}\left(1+\frac{C}{2\nu}\int_{0}^{\xi}e^{-\frac{a}{2\nu}\eta^{2}}d\eta\right),
(2b)U\displaystyle\mbox{(2b)}\;\;U Cea2νξ2+C22ν0ξeaνη2𝑑η,\displaystyle\approx Ce^{-\frac{a}{2\nu}\xi^{2}}+\frac{C^{2}}{2\nu}\int_{0}^{\xi}e^{-\frac{a}{\nu}\eta^{2}}d\eta,

where Ca2πνM.C\approx\sqrt{\frac{a}{2\pi\nu}}M. On this basis we estimate the strength of the nonlinear term N(ξ).N(\xi). The source-type solution is a near identity transform of the Gaussian function. In its series expansion in the Reynolds number Re=M/νRe=M/\nu after non-dimensionalisation, the nonlinear correction term has ReRe as its prefactor. Consider the scheme (1), or (2a) equivalently, taking a/2ν=1a/2\nu=1 without loss of generality. We have

U2=Mπeξ2(1+Re2π0ξeη2𝑑η).U_{2}=\frac{M}{\sqrt{\pi}}e^{-\xi^{2}}\left(1+\frac{Re}{2\sqrt{\pi}}\int_{0}^{\xi}e^{-\eta^{2}}d\eta\right).

Separating out the ReRe-dependence, or equivalently assuming that Re=1,Re=1, we define N(ξ)N(\xi) by

N(ξ)=12π0ξeη2𝑑ηN(\xi)=\frac{1}{2\sqrt{\pi}}\int_{0}^{\xi}e^{-\eta^{2}}d\eta

so that U21+ReN(ξ)U_{2}\propto 1+ReN(\xi) holds. The strength of nonlinearity is given by

N=supξN(ξ)=14.N=\sup_{\xi}N(\xi)=\frac{1}{4}.

The same goes for (2b) from the above expressions for the 1D Burgers equation. Altogether we find

(1)N=14=0.25,\mbox{(1)}\;\;N=\frac{1}{4}=0.25,
(2b)N=1420.2.\mbox{(2b)}\;N=\frac{1}{4\sqrt{2}}\approx 0.2.

We conclude that the typical strength of nonlinearity is N=O(101)N=O(10^{-1}) irrespective of the choice of schemes.

2.4 Burgers equations in several dimensions

The source-type solution is basically a near-identity function of the Gaussian form. It has been seen how the source-type solutions show up in the long-time limit in one and two spatial dimensions in [15]. Here we will take a look at cases in three and higher dimensions. From the Cole-Hopf transform we have

Ui(𝝃,τ)=2νξi3ψ0(λ𝜼)exp(a2ν|𝝃𝜼|21e2aτ)𝑑𝜼3ψ0(λ𝜼)exp(a2ν|𝝃𝜼|21e2aτ)𝑑𝜼.U_{i}(\bm{\xi},\tau)=-2\nu\frac{\partial_{\xi_{i}}\int_{\mathbb{R}^{3}}\psi_{0}(\lambda\bm{\eta})\exp\left(-\dfrac{a}{2\nu}\dfrac{|\bm{\xi}-\bm{\eta}|^{2}}{1-e^{-2a\tau}}\right)d\bm{\eta}}{\int_{\mathbb{R}^{3}}\psi_{0}(\lambda\bm{\eta})\exp\left(-\dfrac{a}{2\nu}\dfrac{|\bm{\xi}-\bm{\eta}|^{2}}{1-e^{-2a\tau}}\right)d\bm{\eta}}.

As we are going for the type 2 scale-invariance, differentiating it twice we find

ξjξkUi(𝝃,τ)=2νξiξjξk3ψ0(λ𝜼)exp(a2ν|𝝃𝜼|21e2aτ)𝑑𝜼3ψ0(λ𝜼)exp(a2ν|𝝃𝜼|21e2aτ)𝑑𝜼+\partial_{\xi_{j}}\partial_{\xi_{k}}U_{i}(\bm{\xi},\tau)=-2\nu\frac{\partial_{\xi_{i}}\partial_{\xi_{j}}\partial_{\xi_{k}}\int_{\mathbb{R}^{3}}\psi_{0}(\lambda\bm{\eta})\exp\left(-\dfrac{a}{2\nu}\dfrac{|\bm{\xi}-\bm{\eta}|^{2}}{1-e^{-2a\tau}}\right)d\bm{\eta}}{\int_{\mathbb{R}^{3}}\psi_{0}(\lambda\bm{\eta})\exp\left(-\dfrac{a}{2\nu}\dfrac{|\bm{\xi}-\bm{\eta}|^{2}}{1-e^{-2a\tau}}\right)d\bm{\eta}}+\ldots
=2νλ33ijkψ0(λ𝜼)exp(a2ν|𝝃𝜼|21e2aτ)d𝜼3ψ0(λ𝜼)exp(a2ν|𝝃𝜼|21e2aτ)𝑑𝜼+=-2\nu\frac{\lambda^{3}\int_{\mathbb{R}^{3}}\partial_{i}\partial_{j}\partial_{k}\psi_{0}(\lambda\bm{\eta})\exp\left(-\dfrac{a}{2\nu}\dfrac{|\bm{\xi}-\bm{\eta}|^{2}}{1-e^{-2a\tau}}\right)d\bm{\eta}}{\int_{\mathbb{R}^{3}}\psi_{0}(\lambda\bm{\eta})\exp\left(-\dfrac{a}{2\nu}\dfrac{|\bm{\xi}-\bm{\eta}|^{2}}{1-e^{-2a\tau}}\right)d\bm{\eta}}+\ldots

The denominator then tends to Kijkexp(a2ν|𝝃|2)K_{ijk}\exp\left(-\frac{a}{2\nu}|\bm{\xi}|^{2}\right) as τ,\tau\to\infty, where Kijk=3ijkψ0(𝜼)d𝜼,(i=1,2,3).K_{ijk}=\int_{\mathbb{R}^{3}}\partial_{i}\partial_{j}\partial_{k}\psi_{0}(\bm{\eta})d\bm{\eta},\;(i=1,2,3). Hence

ξjξkUi(𝝃,)=2ν(Kijkexp(a2ν|𝝃|2)Fijk(𝝃)+),\partial_{\xi_{j}}\partial_{\xi_{k}}U_{i}(\bm{\xi},\infty)=-2\nu\left(\frac{K_{ijk}\exp\left(-\frac{a}{2\nu}|\bm{\xi}|^{2}\right)}{F_{ijk}(\bm{\xi})}+\ldots\right),

where the function FijkF_{ijk} is to be determined such that ijkFijkexp(a2ν|𝝃|2),(no summation implied).\partial_{i}\partial_{j}\partial_{k}F_{ijk}\propto\exp\left(-\frac{a}{2\nu}|\bm{\xi}|^{2}\right),\;\;{\color[rgb]{0,0,0}\mbox{(no summation implied)}}. We can thus take

Fijk(𝝃)=Kijk2ν0ξ1exp(aξ22ν)𝑑ξ0ξ2exp(aη22ν)𝑑η0ξ3exp(aζ22ν)𝑑ζ+1.F_{ijk}(\bm{\xi})=-\frac{K_{ijk}}{2\nu}\int_{0}^{\xi_{1}}\exp\left(-\frac{a\xi^{2}}{2\nu}\right)d\xi\int_{0}^{\xi_{2}}\exp\left(-\frac{a\eta^{2}}{2\nu}\right)d\eta\int_{0}^{\xi_{3}}\exp\left(-\frac{a\zeta^{2}}{2\nu}\right)d\zeta+1.

Therefore after collecting other terms of derivatives we find in three dimensions, say, with (i,j,k)=(1,2,3),(i,j,k)=(1,2,3),

2U1ξ2ξ3=K123exp(a2ν(ξ12+ξ22+ξ32))1+R(ξ1,ξ2,ξ3)(1R(ξ1,ξ2,ξ3))3,\frac{\partial^{2}U_{1}}{\partial\xi_{2}\partial\xi_{3}}=K_{123}\exp\left(-\frac{a}{2\nu}(\xi_{1}^{2}+\xi_{2}^{2}+\xi_{3}^{2})\right)\frac{1+R(\xi_{1},\xi_{2},\xi_{3})}{(1-R(\xi_{1},\xi_{2},\xi_{3}))^{3}}, (7)

where

R(ξ1,ξ2,ξ3)=K1232ν0ξ1exp(aξ22ν)𝑑ξ0ξ2exp(aη22ν)𝑑η0ξ3exp(aζ22ν)𝑑ζR(\xi_{1},\xi_{2},\xi_{3})=\frac{K_{123}}{2\nu}\int_{0}^{\xi_{1}}\exp\left(-\frac{a\xi^{2}}{2\nu}\right)d\xi\int_{0}^{\xi_{2}}\exp\left(-\frac{a\eta^{2}}{2\nu}\right)d\eta\int_{0}^{\xi_{3}}\exp\left(-\frac{a\zeta^{2}}{2\nu}\right)d\zeta

denotes the Reynolds number. Because RR is small the expression (7) is near-Gaussian. It can be verified that K123=32a3π3νtanhM12316ν,K_{123}=\sqrt{\dfrac{32a^{3}}{\pi^{3}\nu}}\tanh\dfrac{M_{123}}{16\nu}, where M123=2U1ξ2ξ3𝑑𝝃.M_{123}=\int\frac{\partial^{2}U_{1}}{\partial\xi_{2}\partial\xi_{3}}d\bm{\xi}. We can also write

2U1ξ2ξ3=2ν3ξ1ξ2ξ3log(1R(ξ1,ξ2,ξ3)),\frac{\partial^{2}U_{1}}{\partial\xi_{2}\partial\xi_{3}}=-2\nu\frac{\partial^{3}}{\partial\xi_{1}\partial\xi_{2}\partial\xi_{3}}\log(1-R(\xi_{1},\xi_{2},\xi_{3})),

which reflects the Cole-Hopf transform more directly, that is, 𝑼=𝝃ϕ,ϕ=2νlog(1R(𝝃)).\bm{U}=\nabla_{\bm{\xi}}\phi,\;\phi=-2\nu\log(1-R(\bm{\xi})). See Appendix A for the general form in nn-dimensions.

3 2D Navier-Stokes equation

We briefly recall the self-similar solution of the 2D Navier-Stokes equations, where the so-called Burgers vortex appears after dynamic scaling.

3.1 Critical scale-invariance

The Burgers vortex was originally introduced to represent the reaction of a vortex under the influence of the collective effect of surrounding vortices in the ambient medium. When we write the steady solution in velocity and vorticity using cylindrical coordinates

𝒖=(ur,uθ,uϕ)=(ar,v(r),2az),𝝎=(0,0,ω(r)),\bm{u}=(u_{r},u_{\theta},u_{\phi})=(-ar,v(r),2az),\;\bm{\omega}=(0,0,\omega(r)),

the solution takes the following forms

ω(r)=aΓ2πνexp(ar22ν),\omega(r)=\frac{a\Gamma}{2\pi\nu}\exp\left(-\frac{ar^{2}}{2\nu}\right),
v(r)=Γ2πr(1exp(ar22ν)),v(r)=\frac{\Gamma}{2\pi r}\left(1-\exp\left(-\frac{ar^{2}}{2\nu}\right)\right),

where Γ2ω0(𝒙)𝑑𝒙\Gamma\equiv\int_{\mathbb{R}^{2}}\omega_{0}(\bm{x})d\bm{x} denotes the velocity circulation.

In two dimensions dynamic scaling transforms take the following form

𝝃=𝒙2at,τ=12alogt,𝝍(𝒙,t)=𝚿(𝝃,τ),\bm{\xi}=\frac{\bm{x}}{\sqrt{2at}},\;\;\tau=\frac{1}{2a}\log t,\bm{\psi}(\bm{x},t)=\bm{\Psi}(\bm{\xi},\tau),
𝒖(𝒙,t)=12at𝑼(𝝃,τ),ω(𝒙,t)=12atΩ(𝝃,τ).\bm{u}(\bm{x},t)=\frac{1}{\sqrt{2at}}\bm{U}(\bm{\xi},\tau),\omega(\bm{x},t)=\frac{1}{2at}\Omega(\bm{\xi},\tau).

In the 2-dimensional case, in order to achieve the critical scale-invariance of type 2, we must choose the second spatial derivative of the stream function, which is the vorticity, as the unknown.

The scaled form of the vorticity equation in two dimensions reads

Ωτ+𝑼Ω=νΩ+a(𝝃Ω),\dfrac{\partial\Omega}{\partial\tau}+\bm{U}\cdot\nabla\Omega=\nu\triangle\Omega+a\nabla\cdot(\bm{\xi}\Omega),

where Ω\Omega satisfies the type 2 scale-invariance. It is known that the self-similar solution under scaling has a mathematically identical form as the Burgers vortex above. Indeed in the scaled variables the above expression can be written

Ω(ξ)=aΓ2πνexp(a|𝝃|22ν),𝝃=𝒙2at.\Omega(\xi)=\frac{a\Gamma}{2\pi\nu}\exp\left(-\frac{a|\bm{\xi}|^{2}}{2\nu}\right),\,\bm{\xi}=\frac{\bm{x}}{\sqrt{2at}}.

Note that 12atΩ(ξ)=Γ4πνtexp(|𝒙|24νt)\frac{1}{2at}\Omega(\xi)=\frac{\Gamma}{4\pi\nu t}\exp\left(-\frac{|\bm{x}|^{2}}{4\nu t}\right) is an exact self-similar decaying solution222When a=0a=0 an exact decaying solution is obtained by formally replacing a12t,a\to\frac{1}{2t}, which is known as the Lamb-Oseen vortex. with the following property

limt0Ω()=Γδ(𝒙).\lim_{t\to 0}\Omega(\cdot)=\Gamma\delta(\bm{x}).

It also satisfies the following asymptotic property, for ω(,0)L1,\omega(\cdot,0)\in L^{1},

t11pω(𝒙,t)12atΩ(𝝃)Lp0ast,t^{1-\frac{1}{p}}\left\|\omega(\bm{x},t)-\frac{1}{2at}\Omega(\bm{\xi})\right\|_{L^{p}}\to 0\;\;\mbox{as}\;\;t\to\infty,

where 1p,1\leq p\leq\infty, see e.g. [16].

3.2 Interpretation

Because the source-type solution coincides with the linearised solution, the inhomogeneous terms on the right-hand side of the approximations at each order vanish identically. Hence there is no way to set up successive approximations that can capture non-zero nonlinear corrections. The strength of nonlinearity is identically zero; N=0N=0.

4 3D Navier-Stokes equations

We will describe two approaches for handling the scaled 3D Navier-Stokes equations perturbatively. First we describe a general framework based on the Green’s function. Second we describe an iterative approach which is specifically suited for calculations associated with the 3D Navier-Stokes problem.

4.1 Critical scale-invariance

We consider the 3D Navier-Stokes equations written in four different dependent variables. Starting from the vector potential and taking a curl successively 𝒖=×𝝍,𝝎=×𝒖,𝝌=×𝝎,\bm{u}=\nabla\times\bm{\psi},\;\bm{\omega}=\nabla\times\bm{u},\;\bm{\chi}=\nabla\times\bm{\omega}, we have

{𝝍t=34πp.v.3𝒓×(×𝝍(𝒚))𝒓(×𝝍(𝒚))|𝒓|5d𝒚+ν𝝍,𝒖t+𝒖𝒖=p+ν𝒖,𝝎t+𝒖𝝎=𝝎𝒖+ν𝝎,𝝌t=(𝒖𝒖+p)+ν𝝌\left\{\begin{array}[]{l}\dfrac{\partial\bm{\psi}}{\partial t}=\dfrac{3}{4\pi}{\rm p.v.}\displaystyle{\int_{\mathbb{R}^{3}}}\dfrac{\bm{r}\times(\nabla\times\bm{\psi}(\bm{y}))\,\bm{r}\cdot(\nabla\times\bm{\psi}(\bm{y}))}{|\bm{r}|^{5}}\;{\rm d}\bm{y}+\nu\triangle\bm{\psi},\\ \vskip 5.69046pt\cr\dfrac{\partial\bm{u}}{\partial t}+\bm{u}\cdot\nabla\bm{u}=-\nabla p+\nu\triangle\bm{u},\\ \vskip 5.69046pt\cr\dfrac{\partial\bm{\omega}}{\partial t}+\bm{u}\cdot\nabla\bm{\omega}=\bm{\omega}\cdot\nabla\bm{u}+\nu\triangle\bm{\omega},\\ \vskip 5.69046pt\cr{\color[rgb]{0,0,0}\dfrac{\partial\bm{\chi}}{\partial t}=\triangle(\bm{u}\cdot\nabla\bm{u}+\nabla p)+\nu\triangle\bm{\chi}}\end{array}\right. (8)

where 𝒓𝒙𝒚\bm{r}\equiv\bm{x}-\bm{y} and p.v. denotes a principal-value integral. We also have 𝝎=𝝍,𝝌=𝒖,\bm{\omega}=-\triangle\bm{\psi},\bm{\chi}=-\triangle\bm{u}, because of the incompressibility condition. The derivation of (8)1 can be found in [17]. The final fourth equation (8)4 is obtained by taking the Laplacian of the velocity equations (8)2. For the 𝝌\bm{\chi} equations we may alternatively take a curl on the vorticity equations to obtain a form of equation different from the final line in (8), which is useful for handling inviscid fluids (details to be found in Appendix B). Under dynamic scaling

𝝃=𝒙2at,τ=12alogt,𝝍(𝒙,t)=𝚿(𝝃,τ),\bm{\xi}=\frac{\bm{x}}{\sqrt{2at}},\;\;\tau=\frac{1}{2a}\log t,\bm{\psi}(\bm{x},t)=\bm{\Psi}(\bm{\xi},\tau),
𝒖(𝒙,t)=1(2at)1/2𝑼(𝝃,τ),𝝎(𝒙,t)=12at𝛀(𝝃,τ),and𝝌(𝒙,t)=1(2at)3/2𝑿(𝝃,τ),\bm{u}(\bm{x},t)=\frac{1}{(2at)^{1/2}}\bm{U}(\bm{\xi},\tau),\bm{\omega}(\bm{x},t)=\frac{1}{2at}\bm{\Omega}(\bm{\xi},\tau),\,\mbox{and}\,\bm{\chi}(\bm{x},t)=\frac{1}{(2at)^{3/2}}\bm{X}(\bm{\xi},\tau),

the 3D Navier-Stokes equations in four different unknowns are transformed respectively to

{𝚿τ=34πp.v.3𝝆×(×𝚿(𝜼))𝝆(×𝚿(𝜼))|𝝆|5d𝜼+ν𝚿+a(𝝃)Ψ,𝑼τ+𝑼𝑼=P+ν𝑼+a(𝝃)𝑼+a𝑼,𝛀τ+𝑼𝛀=𝛀𝑼+ν𝛀+a(𝝃)𝛀+2a𝛀,𝑿τ=(𝑼𝑼+P)+ν𝑿+a(𝝃𝑿),\left\{\begin{array}[]{l}\dfrac{\partial\bm{\Psi}}{\partial\tau}=\dfrac{3}{4\pi}{\rm p.v.}\displaystyle{\int_{\mathbb{R}^{3}}}\dfrac{\bm{\rho}\times(\nabla\times\bm{\Psi}(\bm{\eta}))\,\bm{\rho}\cdot(\nabla\times\bm{\Psi}(\bm{\eta}))}{|\bm{\rho}|^{5}}\;{\rm d}\bm{\eta}+\nu\triangle\bm{\Psi}+a(\bm{\xi}\cdot\nabla)\Psi,\\ \vskip 5.69046pt\cr\dfrac{\partial\bm{U}}{\partial\tau}+\bm{U}\cdot\nabla\bm{U}=-\nabla P+\nu\triangle\bm{U}+a(\bm{\xi}\cdot\nabla)\bm{U}+a\bm{U},\\ \vskip 5.69046pt\cr\dfrac{\partial\bm{\Omega}}{\partial\tau}+\bm{U}\cdot\nabla\bm{\Omega}=\bm{\Omega}\cdot\nabla\bm{U}+\nu\triangle\bm{\Omega}+a(\bm{\xi}\cdot\nabla)\bm{\Omega}+2a\bm{\Omega},\\ \vskip 5.69046pt\cr{\color[rgb]{0,0,0}\dfrac{\partial\bm{X}}{\partial\tau}=\triangle\left(\bm{U}\cdot\nabla\bm{U}+\nabla P\right)+\nu\triangle\bm{X}+a\nabla\cdot(\bm{\xi}\otimes\bm{X}),}\end{array}\right. (9)

where 𝝆𝝃𝜼.\bm{\rho}\equiv\bm{\xi}-\bm{\eta}.

It is to be noted that the coefficient of the linear term increases in number with the increasing order of derivatives and for the variable 𝝌\bm{\chi} a divergence is completed in the convective term. Observe that type 1 scale-invariance is achieved with 𝚿\bm{\Psi} and type 2 scale-invariance with 𝑿\bm{X}.

4.2 Successive approximations

Using the Duhamel principle we convert the scaled Navier-Stokes equations (9)4 into integral equations

𝑿(𝝃,τ)=eντ𝑿0(𝝃)+0τeνs(𝑼𝑼+P)(𝝃,τs)𝑑s.\bm{X}(\bm{\xi},\tau)=e^{\nu\tau\triangle^{*}}\bm{X}_{0}(\bm{\xi})+\int_{0}^{\tau}e^{\nu s\triangle^{*}}\triangle\left(\bm{U}\cdot\nabla\bm{U}+\nabla P\right)(\bm{\xi},\tau-s)ds.

Here +aν(𝝃)\triangle^{*}\equiv\triangle+\frac{a}{\nu}\nabla\cdot(\bm{\xi}\otimes\cdot) and the action of whose exponential operator is given by

exp(ντ)f()=(a2πν(1e2aτ))3/23e3aτf(eaτ𝒚)exp(a2ν|𝝃𝒚|21e2aτ)𝑑𝒚\exp(\nu\tau\triangle^{*})f(\cdot)=\left(\frac{a}{2\pi\nu(1-e^{-2a\tau})}\right)^{3/2}\int_{\mathbb{R}^{3}}e^{3a\tau}f(e^{a\tau}\bm{y})\exp\left(-\frac{a}{2\nu}\frac{|\bm{\xi}-\bm{y}|^{2}}{1-e^{-2a\tau}}\right)d\bm{y} (10)

for any function f,f, as can be verified by combining the heat kernel and dynamic scaling transforms.

The inverse operator associated with the fundamental solution to the Fokker-Planck equation in 3D is defined by

(ν)10𝑑seνs=𝑑𝜼g(𝝃,𝜼),(\nu\triangle^{*})^{-1}\equiv-\int_{0}^{\infty}dse^{\nu s\triangle^{*}}=\int d{\color[rgb]{0,0,0}\bm{\eta}}g(\bm{\xi},\bm{\eta}),

where the Green’s function is given by

g(𝝃,𝜼)1(2πν)3/2f.p.aσ2dσσ2ae12ν|σ𝝃𝜼σ2a|2.g(\bm{\xi},\bm{\eta})\equiv\frac{-1}{(2\pi\nu)^{3/2}}{\rm f.p.}\int_{\sqrt{a}}^{\infty}\frac{\sigma^{2}d\sigma}{\sigma^{2}-a}e^{-\frac{1}{2\nu}|\sigma\bm{\xi}-\bm{\eta}\sqrt{\sigma^{2}-a}|^{2}}.

This can be verified by changing the variable from ss to σ=a1e2aτ\sigma=\sqrt{\frac{a}{1-e^{-2a\tau}}} in the solution (10) to the Fokker-Planck equation.

We consider the steady solution 𝑿(𝝃)\bm{X}(\bm{\xi}) in the long-time limit of τ\tau\to\infty

𝑿(𝝃)=𝑿1(𝝃)+limτ0τeνs(𝑼𝑼+P)(𝝃,τs)𝑑s,\bm{X}(\bm{\xi})=\bm{X}_{1}(\bm{\xi})+\lim_{\tau\to\infty}\int_{0}^{\tau}e^{\nu s\triangle^{*}}\triangle\left(\bm{U}\cdot\nabla\bm{U}+\nabla P\right)(\bm{\xi},\tau-s)ds,

where 𝑿1=𝑴G\bm{X}_{1}=\mathbb{P}\bm{M}G denotes the leading-order approximation (to be made more explicit in next subsection). This is one form of integral equations we are supposed to handle.

On the other hand, steady equations are obtained by assuming /τ=0\partial/\partial\tau=0 in (9)4

𝑿𝑿+aν(𝝃𝑿)=1ν(𝑼𝑼+P),\triangle^{*}\bm{X}\equiv\triangle\bm{X}+\frac{a}{\nu}\nabla\cdot(\bm{\xi}\otimes\bm{X})=-\frac{1}{\nu}\triangle\left(\bm{U}\cdot\nabla\bm{U}+\nabla P\right),

or

𝑿=1ν(𝑼𝑼+P)aν1(𝝃𝑿).\bm{X}=-\frac{1}{\nu}\left(\bm{U}\cdot\nabla\bm{U}+\nabla P\right)-\frac{a}{\nu}\triangle^{-1}\nabla\cdot(\bm{\xi}\otimes\bm{X}).

This is yet another form of the steady Navier-Stokes equations after dynamic scaling. It is noted that one of the potential problems associated with the nonlinear term has been eliminated without having a recourse to the Green’s function. It is this virtually trivial fact that allows us to set up a simple successive approximation.

To summarise, the steady Navier-Stokes equations after dynamic scaling can be written as

𝑿=1ν(𝑼𝑼)aν1(𝝃𝑿),\bm{X}=-\frac{1}{\nu}\mathbb{P}\left(\bm{U}\cdot\nabla\bm{U}\right)-\frac{a}{\nu}\triangle^{-1}\nabla\cdot(\bm{\xi}\otimes\bm{X}), (11)

or, by 𝑿=𝑼,\bm{X}=-\triangle\bm{U}, we can express it solely in terms of 𝑿\bm{X} as

𝑿=1ν(1𝑿1𝑿)aν1(𝝃𝑿).\bm{X}=-\frac{1}{\nu}\mathbb{P}\left(\triangle^{-1}\bm{X}\cdot\nabla\triangle^{-1}\bm{X}\right)-\frac{a}{\nu}\triangle^{-1}\nabla\cdot(\bm{\xi}\otimes\bm{X}). (12)

This is the set of equations that we need to solve.

In passing we note the following facts before proceeding to the specific results. By the definition of scaled variables it is easily seen that for p1p\geq 1

t32(11p)𝝌(𝒙,t)𝑿(𝝃)(2at)3/2Lp=𝑿(𝝃,τ)𝑿(𝝃)Lp.t^{\frac{3}{2}\left(1-\frac{1}{p}\right)}\left\|\bm{\chi}(\bm{x},t)-\frac{\bm{X}(\bm{\xi})}{(2at)^{3/2}}\right\|_{L^{p}}=\left\|\bm{X}(\bm{\xi},\tau)-\bm{X}(\bm{\xi})\right\|_{L^{p}}.

This means that if 𝑿(𝝃,τ)𝑿(𝝃)Lp0\left\|\bm{X}(\bm{\xi},\tau)-\bm{X}(\bm{\xi})\right\|_{L^{p}}\to 0 as τ,\tau\to\infty, we have

t32(11p)𝝌(𝒙,t)𝑿(𝝃)(2at)3/2Lp0ast.t^{\frac{3}{2}\left(1-\frac{1}{p}\right)}\left\|\bm{\chi}(\bm{x},t)-\frac{\bm{X}(\bm{\xi})}{(2at)^{3/2}}\right\|_{L^{p}}\to 0\;\;\mbox{as}\;\;t\to\infty.

That is about the long-time asymptotics. On the other hand as time-zero asymptotics we have

𝑿(𝝃)(2at)3/2𝑴δast0,\frac{\bm{X}(\bm{\xi})}{(2at)^{3/2}}\to\mathbb{P}\bm{M}\delta\;\;\mbox{as}\;\;t\to 0,

where δ()\delta(\cdot) is the Dirac mass.

4.3 Leading-order approximations

Before discussing the second-order approximation we derive expressions of the first-order solutions in several different variables.

We will derive the basic formulas by solving the heat equation

𝝍1t=ν𝝍1.\frac{\partial\bm{\psi}_{1}}{\partial t}=\nu\triangle\bm{\psi}_{1}.

The first-order approximation is given by

𝝍1(𝒙,t)=1(4πνt)3/23𝝍0(𝒚)exp(|𝒙𝒚|24νt)𝑑𝒚.\bm{\psi}_{1}(\bm{x},t)=\frac{1}{(4\pi\nu t)^{3/2}}\int_{\mathbb{R}^{3}}\bm{\psi}_{0}(\bm{y})\exp\left(-\frac{|\bm{x}-\bm{y}|^{2}}{4\nu t}\right)d\bm{y}.

After applying the dynamic scaling

𝝍1(𝒙,t)=𝚿1(𝝃,τ),𝝃=𝒙2at,τ=12alogt,\bm{\psi}_{1}(\bm{x},t)=\bm{\Psi}_{1}(\bm{\xi},\tau),\bm{\xi}=\frac{\bm{x}}{\sqrt{2at}},\;\tau=\frac{1}{2a}\log t,

the linearised equations for the vector potential read

𝚿1τ=a𝝃𝚿1+ν𝚿1.\frac{\partial\bm{\Psi}_{1}}{\partial\tau}=a\bm{\xi}\cdot\nabla\bm{\Psi}_{1}+\nu\triangle\bm{\Psi}_{1}.

After dynamic scaling their solution is given by

𝚿1(𝝃,τ)=e3aτ(a2πν(1e2aτ))3/23𝚿0(𝜼)exp(a2ν|𝝃𝜼eaτ|21e2aτ)𝑑𝜼\bm{\Psi}_{1}(\bm{\xi},\tau)=e^{-3a\tau}\left(\frac{a}{2\pi\nu(1-e^{-2a\tau})}\right)^{3/2}\int_{\mathbb{R}^{3}}\bm{\Psi}_{0}(\bm{\eta})\exp\left(-\frac{a}{2\nu}\frac{|\bm{\xi}-\bm{\eta}e^{-a\tau}|^{2}}{1-e^{-2a\tau}}\right)d\bm{\eta}
=(a2πν(1e2aτ))3/23𝚿0(eaτ𝒚)exp(a2ν|𝝃𝒚|21e2aτ)𝑑𝒚,=\left(\frac{a}{2\pi\nu(1-e^{-2a\tau})}\right)^{3/2}\int_{\mathbb{R}^{3}}\bm{\Psi}_{0}(e^{a\tau}\bm{y})\exp\left(-\frac{a}{2\nu}\frac{|\bm{\xi}-\bm{y}|^{2}}{1-e^{-2a\tau}}\right)d\bm{y},

where 𝚿1\bm{\Psi}_{1} denotes the first-order approximation and 𝚿0\bm{\Psi}_{0} the initial data. (The same convention applies to 𝑿1\bm{X}_{1} and 𝑿0\bm{X}_{0} in the following.) Taking a curl with respect to 𝝃\bm{\xi} three times we find the expressions for the vorticity curl

𝑿1(𝝃,τ)=(a2πν(1e2aτ))3/23e3aτ𝑿0(eaτ𝒚)exp(a2ν|𝝃𝒚|21e2aτ)𝑑𝒚.\bm{X}_{1}(\bm{\xi},\tau)=\left(\frac{a}{2\pi\nu(1-e^{-2a\tau})}\right)^{3/2}\int_{\mathbb{R}^{3}}e^{3a\tau}\bm{X}_{0}(e^{a\tau}\bm{y})\exp\left(-\frac{a}{2\nu}\frac{|\bm{\xi}-\bm{y}|^{2}}{1-e^{-2a\tau}}\right)d\bm{y}.

For well-localised initial data we make use of the formula λ3𝑿0(λ𝒚)𝑴δ(𝒚)\lambda^{3}\bm{X}_{0}(\lambda\bm{y})\to\bm{M}\delta(\bm{y}) where 𝑴=𝑿0𝑑𝒚.\bm{M}=\int\bm{X}_{0}d\bm{y}. Noting that 𝑿0=𝑿0\mathbb{P}\bm{X}_{0}=\bm{X}_{0} we have333If the initial condition satisfies the similarity condition λ3𝑿0(λ𝒚)=𝑿0(𝒚)forλ(>0),\lambda^{3}\bm{X}_{0}(\lambda\bm{y})=\bm{X}_{0}(\bm{y})\;\mbox{for}\,\forall\lambda(>0), we have 𝑿1(𝝃,τ)(a2πν)3/23𝑿0(𝒚)exp(a2ν|𝝃𝒚|2)𝑑𝒚=𝑿0G.\bm{X}_{1}(\bm{\xi},\tau)\to\left(\frac{a}{2\pi\nu}\right)^{3/2}\int_{\mathbb{R}^{3}}\bm{X}_{0}(\bm{y})\exp\left(-\frac{a}{2\nu}|\bm{\xi}-\bm{y}|^{2}\right)d\bm{y}=\bm{X}_{0}*G. In this case 𝑿0(𝝃)\bm{X}_{0}(\bm{\xi}) is singular like |𝝃|3\sim|\bm{\xi}|^{-3}.

𝑿1(𝝃,τ)\displaystyle\bm{X}_{1}(\bm{\xi},\tau) =\displaystyle= (a2πν(1e2aτ))3/23e3aτ(𝑿0(eaτ𝒚))exp(a2ν|𝝃𝒚|21e2aτ)𝑑𝒚\displaystyle\left(\frac{a}{2\pi\nu(1-e^{-2a\tau})}\right)^{3/2}\int_{\mathbb{R}^{3}}e^{3a\tau}(\mathbb{P}\bm{X}_{0}(e^{a\tau}\bm{y}))\exp\left(-\frac{a}{2\nu}\frac{|\bm{\xi}-\bm{y}|^{2}}{1-e^{-2a\tau}}\right)d\bm{y}
=\displaystyle= (a2πν(1e2aτ))3/23e3aτ𝑿0(eaτ𝒚)exp(a2ν|𝝃𝒚|21e2aτ)𝑑𝒚\displaystyle\left(\frac{a}{2\pi\nu(1-e^{-2a\tau})}\right)^{3/2}\int_{\mathbb{R}^{3}}e^{3a\tau}\bm{X}_{0}(e^{a\tau}\bm{y})\mathbb{P}\exp\left(-\frac{a}{2\nu}\frac{|\bm{\xi}-\bm{y}|^{2}}{1-e^{-2a\tau}}\right)d\bm{y}
\displaystyle\to 𝑴Gasτ,\displaystyle\mathbb{P}\bm{M}G\>\>\mbox{as}\;\;\tau\to\infty,

where G(𝝃)(a2πν)3/2exp(a2ν|𝝃|2)G(\bm{\xi})\equiv\left(\frac{a}{2\pi\nu}\right)^{3/2}\exp\left(-\frac{a}{2\nu}|\bm{\xi}|^{2}\right) and 𝑴=𝑿0𝑑𝝃.\bm{M}=\int\bm{X}_{0}d\bm{\xi}. This is the leading-order approximation for the scaled 3D Navier-Stokes equations.

The first-order (that is, the leading-order) approximation obtained above can be calculated explicitly because the Gaussian function is a radial function. Care should be taken that the leading-order approximations themselves are not radial because of the incompressibility condition. Indeed, in terms of the vorticity curl the first-order approximation is given by

Xi\displaystyle X_{i} =\displaystyle= Mj(a2πν)3/2(δijij1)exp(a2νr2),(i=1,2,3,)\displaystyle M_{j}\left(\frac{a}{2\pi\nu}\right)^{3/2}\left(\delta_{ij}-\partial_{i}\partial_{j}\triangle^{-1}\right)\exp\left(-\frac{a}{2\nu}r^{2}\right),\;\;(i=1,2,3,)
=\displaystyle= Mi(μπ)3/2exp(μr2)+Mjjierf(μr)4πr\displaystyle M_{i}\left(\frac{\mu}{\pi}\right)^{3/2}\exp(-\mu r^{2})+M_{j}\partial_{j}\partial_{i}\frac{{\rm erf}(\sqrt{\mu}r)}{4\pi r}
=\displaystyle= Mj(δijξiξjr2)(μπ)3/2eμr2Mj(δijr33ξiξjr5){erf(μr)4πr2μ(μπ)3/2eμr2},\displaystyle M_{j}\left(\delta_{ij}-\frac{\xi_{i}\xi_{j}}{r^{2}}\right)\left(\frac{\mu}{\pi}\right)^{3/2}e^{-\mu r^{2}}-M_{j}\left(\frac{\delta_{ij}}{r^{3}}-\frac{3\xi_{i}\xi_{j}}{r^{5}}\right)\left\{\frac{{\rm erf}(\sqrt{\mu}r)}{4\pi}{\color[rgb]{0,0,0}-}\frac{r}{2\mu}\left(\frac{\mu}{\pi}\right)^{3/2}e^{-\mu r^{2}}\right\},\noindent

where μ=a/(2ν),\mu=a/(2\nu), r=|𝝃|,r=|\bm{\xi}|, Mj=Xj𝑑𝝃,(j=1,2,3)M_{j}=\int X_{j}d\bm{\xi},\;(j=1,2,3) and summation is implied on repeated indices. In the second line we computed 1\triangle^{-1} for the Gaussian function using =1r2ddr(r2ddr)\triangle=\frac{1}{r^{2}}\frac{d}{dr}\left(r^{2}\frac{d}{dr}\right) and the final line by direct computations. Clearly the final expression is not radial.

Hereafter in this subsection we take μ=1\mu=1 for simplicity. Note that ner2\triangle^{-n}e^{-r^{2}} for n=1,2,3n=1,2,3 can be evaluated by quadratures and their explicit form are as follows, which can be obtained most conveniently with the assistance of computer algebra. The results are

1er2\displaystyle\triangle^{-1}e^{-r^{2}} =\displaystyle= 12r0res2𝑑s=π4rerf(r),\displaystyle-\frac{1}{2r}\int_{0}^{r}e^{-s^{2}}ds=-\frac{\sqrt{\pi}}{4r}{\rm erf}(r), (14)
2er2\displaystyle\triangle^{-2}e^{-r^{2}} =\displaystyle= 12r0r𝑑s0s𝑑s0ses′′2𝑑s′′\displaystyle-\frac{1}{2r}\int_{0}^{r}ds\int_{0}^{s}ds^{\prime}\int_{0}^{s^{\prime}}e^{-s^{\prime\prime 2}}ds^{\prime\prime} (15)
=\displaystyle= er28π8(r+12r)erf(r),\displaystyle-\frac{e^{-r^{2}}}{8}-\frac{\sqrt{\pi}}{8}\left(r+\frac{1}{2r}\right){\rm erf}(r),
3er2\displaystyle\triangle^{-3}e^{-r^{2}} =\displaystyle= 12r0r𝑑s0s𝑑s0s𝑑s′′0s′′𝑑s′′′0s′′′𝑑s′′′′es′′′′2,\displaystyle-\frac{1}{2r}\int_{0}^{r}ds\int_{0}^{s}ds^{\prime}\int_{0}^{s^{\prime}}ds^{\prime\prime}\int_{0}^{s^{\prime\prime}}ds^{\prime\prime\prime}\int_{0}^{s^{\prime\prime\prime}}ds^{\prime\prime\prime\prime}e^{-s^{\prime\prime\prime\prime 2}}, (16)
=\displaystyle= 1384r[(4r3+10r)er2+4π(r4+3r2+34)erf(r)].\displaystyle-\frac{1}{384r}\left[(4r^{3}+10r)e^{-r^{2}}+4\sqrt{\pi}\left(r^{4}+3r^{2}+\frac{3}{4}\right){\rm erf}(r)\right].

With these at hand the expressions for the several different unknowns, including the vorticity curl above, are (i=1,2,3i=1,2,3)

Bi\displaystyle B_{i} =\displaystyle= 1π3/2(Mi2er2Mjij3er2),\displaystyle\frac{1}{\pi^{3/2}}\left(M_{i}\triangle^{-2}e^{-r^{2}}-M_{j}\partial_{i}\partial_{j}\triangle^{-3}e^{-r^{2}}\right), (17)
Ψi\displaystyle{\color[rgb]{0,0,0}\Psi_{i}} =\displaystyle= ϵijkMjξk8π3/2J(r),\displaystyle\frac{\epsilon_{ijk}M_{j}{\color[rgb]{0,0,0}\xi_{k}}}{8\pi^{3/2}}{\color[rgb]{0,0,0}J(r)}, (18)
Ui\displaystyle U_{i} =\displaystyle= 1π3/2(Mi1er2Mjij2er2),\displaystyle-\frac{1}{\pi^{3/2}}\left(M_{i}\triangle^{-1}e^{-r^{2}}-M_{j}\partial_{i}\partial_{j}\triangle^{-2}e^{-r^{2}}\right), (19)
Ωi\displaystyle\Omega_{i} =\displaystyle= ϵijkMjξk4π3/2H(r),\displaystyle\frac{\epsilon_{ijk}M_{j}{\color[rgb]{0,0,0}\xi_{k}}}{4\pi^{3/2}}{\color[rgb]{0,0,0}H(r)}, (20)
Xi\displaystyle X_{i} =\displaystyle= 1π3/2(Mier2Mjij1er2),\displaystyle\frac{1}{\pi^{3/2}}\left(M_{i}e^{-r^{2}}-M_{j}\partial_{i}\partial_{j}\triangle^{-1}e^{-r^{2}}\right), (21)

where 𝚿=×𝑩,𝑼=×𝚿,𝛀=×𝑼,𝑿=×𝛀{\color[rgb]{0,0,0}\bm{\Psi}}=\nabla\times\bm{B},\bm{U}=\nabla\times{\color[rgb]{0,0,0}\bm{\Psi}},\bm{\Omega}=\nabla\times\bm{U},\bm{X}=\nabla\times\bm{\Omega} and erf(r)2π0ret2𝑑t{\rm erf}(r)\equiv\frac{2}{\sqrt{\pi}}\int_{0}^{r}e^{-t^{2}}dt denotes the error function. Here for convenience we have introduced two functions

H(r)πerf(r)2rer2r3,J(r)rer2+πerf(r)(r212)r3,H(r)\equiv\frac{\sqrt{\pi}{\rm erf}(r)-2re^{-r^{2}}}{r^{3}},\;J(r)\equiv\frac{re^{-r^{2}}+\sqrt{\pi}{\rm erf}(r)\left(r^{2}-\frac{1}{2}\right)}{r^{3}},

both of which are continuous at r=0r=0 (See Figure 1 below). Because all the fields are incompressible we also have 𝑼=𝑩,𝛀=𝚿,𝑿=𝑼.\bm{U}=-\triangle\bm{B},\bm{\Omega}=-\triangle{\color[rgb]{0,0,0}\bm{\Psi}},\bm{X}=-\triangle\bm{U}.

Regarding the derivations of (17)-(21), applying 1\triangle^{-1} to (21) repeatedly we get (19) and (17), respectively. Then, taking a curl of (17) we get (18) and taking a curl of (19) we get (20). Finally simpler forms of (19) and (21) are obtained by taking a curl of (18) and (20), respectively. The final results in vectorial form read

𝚿\displaystyle\bm{\Psi} =\displaystyle= 𝑴×𝝃8π3/2J(r),\displaystyle\frac{\bm{M}\times\bm{\xi}}{8\pi^{3/2}}J(r), (22)
𝑼\displaystyle\bm{U} =\displaystyle= erf(r)4πr(𝑴(𝑴𝝃)𝝃r2)J(r)8π3/2(𝑴3(𝑴𝝃)𝝃r2),\displaystyle\frac{\rm{erf}(r)}{4\pi r}\left(\bm{M}-\frac{(\bm{M}\cdot\bm{\xi})\bm{\xi}}{r^{2}}\right)-\frac{J(r)}{8\pi^{3/2}}\left(\bm{M}-\frac{3(\bm{M}\cdot\bm{\xi})\bm{\xi}}{r^{2}}\right), (23)
𝛀\displaystyle\bm{\Omega} =\displaystyle= 𝑴×𝝃4π3/2H(r),\displaystyle\frac{\bm{M}\times\bm{\xi}}{4\pi^{3/2}}H(r), (24)
𝑿\displaystyle\bm{X} =\displaystyle= er2π3/2(𝑴(𝑴𝝃)𝝃r2)H(r)4π3/2(𝑴3(𝑴𝝃)𝝃r2).\displaystyle\frac{e^{-r^{2}}}{\pi^{3/2}}\left(\bm{M}-\frac{(\bm{M}\cdot\bm{\xi})\bm{\xi}}{r^{2}}\right)-\frac{H(r)}{4\pi^{3/2}}\left(\bm{M}-\frac{3(\bm{M}\cdot\bm{\xi})\bm{\xi}}{r^{2}}\right). (25)

It is of interest to observe that 𝑼𝛀=𝛀𝑿=0\bm{U}\cdot\bm{\Omega}=\bm{\Omega}\cdot\bm{X}=0. Using the above formula (20) with 𝑴=(1,1,1)\bm{M}=(1,1,1) it is instructive to compare a component of the Gaussian function 1π3/2eξ2\frac{1}{\pi^{3/2}}e^{-{\color[rgb]{0,0,0}\xi}^{2}} with that of the vorticity curl

X1(ξ,0,0)=πerf(ξ)2xeξ22π3/2ξ3(=H(ξ)2π3/2).X_{1}({\color[rgb]{0,0,0}\xi},0,0)=\frac{\sqrt{\pi}{\rm erf}({\color[rgb]{0,0,0}\xi})-2xe^{-{\color[rgb]{0,0,0}\xi}^{2}}}{2\pi^{3/2}{\color[rgb]{0,0,0}\xi}^{3}}\;\;\;\;{\color[rgb]{0,0,0}\left(=\frac{H(\xi)}{2\pi^{3/2}}\right)}. (26)

Figure 1 shows how X1(ξ,0,0)X_{1}(\xi,0,0) is affected by the incompressible condition (solenoidality), in particular the peak value at ξ=0\xi=0 is reduced by a factor of 2/32/3.

Refer to caption
Figure 1: Comparison of exp(ξ2)\exp(-\xi^{2}) (solid), π3/2χ1(ξ)=H(ξ)/2\pi^{3/2}\chi_{1}(\xi)=H(\xi)/2 (dashed) and J(ξ)/2J(\xi)/2 (dotted). See the text for their definitions. Note that H(0)=J(0)=2/3.H(0)=J(0)=2/3.

4.4 Estimation of the strength of nonlinearity

We first describe the numerical methods employed to obtain the approximate solutions. We use a centered finite-difference scheme in a box [L,L]3[-L,L]^{3} of size L,L, with discretised coordinates (xi,yj,zk)=(iΔh,jΔh,kΔh)(x_{i},y_{j},z_{k})=(i\Delta h,j\Delta h,k\Delta h) where i,j,k=N,,Ni,j,k=-N,\ldots,N and Δh=L/N\Delta h=L/N.

An important step in the determination of the second-order approximation is the inversion of the Laplacian operator. This was done by using a Poisson solver in the Intel MKL library. Parameters used are L=40,N=800,L=40,N=800, hence Δh=0.05\Delta h=0.05. The box size LL was chosen large enough to capture the decay of 𝑿(𝝃)\bm{X}(\bm{\xi}) near the boundaries and NN was chosen large enough to resolve spatial structure in the center of the domain. As a code validation we calculated (4.6) numerically and compared against the analytical expression (26) and confirmed their agreement (figure omitted).

To evaluate the perturbation we make use of the iteration scheme (2b) illustrated in the previous section for simplicity. Consider a series expansion of the source-solution in ReRe. In comparison to the leading-order the nonlinear correction is proportional to ReRe.444After full non-dimensionalisation 𝑿^=ν𝝌(𝒙,t)(2at)3/2\widehat{\bm{X}}=\dfrac{\nu\bm{\chi}(\bm{x},t)}{(2at)^{3/2}} we have 𝑿^2=𝑴^G(1𝑴^G1𝑴^G),\widehat{\bm{X}}_{2}=\mathbb{P}\widehat{\bm{M}}G-\mathbb{P}\left(\triangle^{-1}\mathbb{P}\widehat{\bm{M}}G\cdot\nabla\triangle^{-1}\mathbb{P}\widehat{\bm{M}}G\right), where 𝑴^=M/ν.\widehat{\bm{M}}=M/\nu. As |𝑴^|=Re|\widehat{\bm{M}}|=Re it is clear that the nonlinear term has a factor of ReRe in comparison to the leading-order approximation. Separating out Re,Re, or equivalently assuming that Re=1,Re=1, we define the strength of nonlinearity by the remaining factor. To be more specific the second-order solution in this case is given by

𝑿2\displaystyle\bm{X}_{2} =\displaystyle= 𝑿11ν(1𝑿11𝑿1)\displaystyle\bm{X}_{1}-\frac{1}{\nu}\mathbb{P}\left(\triangle^{-1}\bm{X}_{1}\cdot\nabla\triangle^{-1}\bm{X}_{1}\right)
=\displaystyle= 𝑴G1ν(1𝑴G1𝑴G)\displaystyle\mathbb{P}\bm{M}G-\frac{1}{\nu}\mathbb{P}\left(\triangle^{-1}\mathbb{P}\bm{M}G\cdot\nabla\triangle^{-1}\mathbb{P}\bm{M}G\right)
=\displaystyle= 𝑴GRe(1𝑴G1𝑴G)/|𝑴|.\displaystyle\mathbb{P}\bm{M}G-Re\,\mathbb{P}\left(\triangle^{-1}\mathbb{P}\bm{M}G\cdot\nabla\triangle^{-1}\mathbb{P}\bm{M}G\right)/|\bm{M}|.

Separating out the ReRe-dependence we define N(𝝃)N(\bm{\xi}) by

N(𝝃)=|(1𝑴G1𝑴G)|/|𝑴|sup𝝃|𝑴G|N(\bm{\xi})=\frac{\left|\mathbb{P}\left(\triangle^{-1}\mathbb{P}\bm{M}G\cdot\nabla\triangle^{-1}\mathbb{P}\bm{M}G\right)\right|/|\bm{M}|}{\sup_{\bm{\xi}}|\mathbb{P}\bm{M}G|} (27)

and put N=sup𝝃N(𝝃)N=\sup_{\bm{\xi}}N(\bm{\xi}).

Refer to caption
Figure 2: The first order term, i.e. the ξ1\xi_{1}-component of 𝑿1=𝑴G\bm{X}_{1}=\mathbb{P}\bm{M}G as a function of ξ2\xi_{2}.
Refer to caption
Figure 3: The ξ1\xi_{1}-component of the second order correction, i.e. the nonlinear term 1ν(1𝑿11𝑿1)\frac{1}{\nu}\mathbb{P}\left(\triangle^{-1}\bm{X}_{1}\cdot\nabla\triangle^{-1}\bm{X}_{1}\right) depicted as in Fig.2.

It turns out that the ξ1\xi_{1}-component of the second-order correction along the line (ξ1,0,0)(\xi_{1},0,0) is identically equal to zero owing to the radial symmetry of the Gaussian function. For this reason we show in Figure 3 the ξ1\xi_{1}-component of the first-order approximation 𝑿1\bm{X}_{1} as a function of ξ2\xi_{2} along the line (0,ξ2,0)(0,\xi_{2},0). It has a peak at the origin whose height is approximately 0.12. Accordingly we show in Figure 3 the ξ1\xi_{1}-component of the second-order correction due to the nonlinearity 1ν(1𝑿11𝑿1)\frac{1}{\nu}\mathbb{P}\left(\triangle^{-1}\bm{X}_{1}\cdot\nabla\triangle^{-1}\bm{X}_{1}\right) as a function of ξ2\xi_{2}. It has double peaks near the origin, but their value is small and is about 0.0015. Noting Re=Mν=11/2=2,Re=\frac{M}{\nu}=\frac{1}{1/2}=2, by (27) we can estimate the strength of nonlinearity in that cross-section as

N0.00152×0.126×103.N\approx\frac{0.0015}{2\times 0.12}\approx 6\times 10^{-3}.

Actually the maximum value of the nonlinear term in the above sense in 3\mathbb{R}^{3} is 0.0022, not much different from the above value. Thus the strength of nonlinearity is at most N9×103N\approx 9\times 10^{-3} and we conclude

N=O(102)for the 3D Navier-Stokes equations.N=O(10^{-2})\;\;\mbox{for the 3D Navier-Stokes equations.}

It should be noted that it is much smaller than the value of NN for the Burgers equations, whose solutions are known to remain regular all time. Because the difference between the Navier-Stokes and Burgers equations is the presence or absence of the incompressibility condition, it is the incompressibility that makes the value of NN smaller for the Navier-Stokes equations. On a practical side this also means that even if we add the second-order correction to the first-order term at low Reynolds number, say Re=1,Re=1, the superposed solution is virtually indistinguishable from the first-order approximation. In the final period of decay the Navier-Stokes flows are very close to (25), to within 1%1\% at Re=1,Re=1, which may be regarded as the building blocks for representing the late-stage of evolution, that is, as the counterpart of the Burgers vortex in two dimensions.

5 Summary and outlook

We have studied self-similar solutions to the fluid dynamical equations with particular focus on the so-called source-type solutions. As an illustration of successive approximation schemes we have discussed the 1D Burgers equation which is exactly soluble. In this case the velocity is the most convenient choice for its analysis. We have introduced a method of quantitatively assessing the strength of nonlinearity NN using the source-type solutions. Similar analyses have been carried out for higher-dimensional Burgers equations. For the Burgers equations we find N=O(0.1)N=O(0.1) in any dimensions.

We then move on to consider the 2D and 3D Navier-Stokes equations. In two dimensions we review the known results done using the vorticity. In three dimensions the most convenient choice of the unknown is the vorticity curl. We have formulated the dynamically-scaled equations using that variable and set up the successive approximation schemes. We have found that the second-order correction stemming from the nonlinear term gives rise to N0.01N\approx 0.01, an order of magnitude smaller than that for the Burgers equations. We are led to conclude that the incompressible condition makes NN smaller for the Navier-Stokes equations than for the Burgers equations.

The current approach relies on perturbative treatments. It may be challenging, but worthwhile to study the functional form of the solution by non-perturbative methods for further theoretical developments. It is also of interest to seek a fully non-linear solution by numerical methods. It is noted that this is at least one order of magnitude smaller than NN found for the Burgers equations whose solutions are known to remain regular all the time. As an application of the source-type solution, it is useful to characterise the late stage of statistical solutions of the Navier-Stokes equations [15].

Appendix A Source solution to the Burgers equations in nn dimensions

The source-type solution to the nn-dimensional Burgers equations takes the following form

n1U1ξ2ξn=K1nexp(a2ν|𝝃|2)Pn(R(𝝃))(1R(𝝃))n,\frac{\partial^{n-1}U_{1}}{\partial\xi_{2}\ldots\partial\xi_{n}}=K_{1\ldots n}\exp\left(-\frac{a}{2\nu}|\bm{\xi}|^{2}\right)\frac{P_{n}(-R(\bm{\xi}))}{(1-R(\bm{\xi}))^{n}},

where Pn(s)P_{n}(s)’s are polynomials to be constructed below,

R(𝝃)=\displaystyle R(\bm{\xi})= K1n2νΠj=1n0ξjexp(a2νξ2)𝑑ξ,\displaystyle\frac{K_{1\ldots n}}{2\nu}\Pi_{j=1}^{n}\int_{0}^{\xi_{j}}\exp\left(-\frac{a}{2\nu}\xi^{2}\right)d\xi,
K1n=\displaystyle K_{1\ldots n}= 2ν(2aπν)n/2tanhM1n2n+1ν(a2πν)n/2M1n(forM1n/ν1)\displaystyle 2\nu\left(\frac{2a}{\pi\nu}\right)^{n/2}\tanh\frac{M_{1\ldots n}}{2^{n+1}\nu}\approx\left(\frac{a}{2\pi\nu}\right)^{n/2}M_{1\ldots n}\;\;(\mbox{for}\;\;M_{1\ldots n}/\nu\ll 1)

and M1n=n1U1ξ2ξn𝑑𝝃.M_{1\ldots n}=\int\dfrac{\partial^{n-1}U_{1}}{\partial\xi_{2}\ldots\partial\xi_{n}}d\bm{\xi}.

Construction
Consider a function s(x1,x2,,xn)s(x_{1},x_{2},\ldots,x_{n}) separable in nn variables x1,x2,,xnx_{1},x_{2},\ldots,x_{n}

s(x1,x2,,xn)=F(x1)F(x2)F(xn),s(x_{1},x_{2},\ldots,x_{n})=F(x_{1})F(x_{2})\ldots F(x_{n}),

where F()F(\cdot) is a smooth function and nn\in\mathbb{N}. Define another function ϕ\phi by

ϕ(x1,x2,,xn)=log(1+s),\phi(x_{1},x_{2},\ldots,x_{n})=\log(1+s),

then we have

nϕx1x2xn=nsx1x2xnPn(s)(1+s)n,\frac{\partial^{n}\phi}{\partial x_{1}\partial x_{2}\ldots\partial x_{n}}=\frac{\partial^{n}s}{\partial x_{1}\partial x_{2}\ldots\partial x_{n}}\frac{P_{n}(s)}{(1+s)^{n}},

where Pn(s)P_{n}(s) is a sequence of polynomials in ss of degree not exceeding n2.n-2. In fact, it is given by

Pn(s)=(1+s)n(ddss)nlog(1+s)s,for(n=1,2,)P_{n}(s)=(1+s)^{n}\left(\frac{d}{ds}s\right)^{n}\frac{\log(1+s)}{s},\;\;\mbox{for}\;\;(n=1,2,\ldots)

or equivalently,

Pn(s)=(1+s)nk=0(k+1)n1(s)k.P_{n}(s)=(1+s)^{n}\sum_{k=0}^{\infty}(k+1)^{n-1}(-s)^{k}.

The first four of them are P1(s)=P2(s)=1,P3(s)=1s,P4(s)=14s+s2.P_{1}(s)=P_{2}(s)=1,P_{3}(s)=1-s,P_{4}(s)=1-4s+s^{2}.

Proof (Due to Yuji Okitani.)

ϕx1=sx111+s.\phi_{x_{1}}=s_{x_{1}}\frac{1}{1+s}.

Noting sx2sx1=ssx1x2,s_{x_{2}}s_{x_{1}}=ss_{x_{1}x_{2}}, we have

ϕx1x2=sx1x211+ssx11(1+s)2sx2=sx1x2(11+ss(1+s)2)=sx1x21(1+s)2,\phi_{x_{1}x_{2}}=s_{x_{1}x_{2}}\frac{1}{1+s}-s_{x_{1}}\frac{1}{(1+s)^{2}}s_{x_{2}}=s_{x_{1}x_{2}}\left(\frac{1}{1+s}-\frac{s}{(1+s)^{2}}\right)=s_{x_{1}x_{2}}\frac{1}{(1+s)^{2}},

while the penultimate expression of which may also be written

=sx1x2(11+s+sdds11+s)f2(s).=s_{x_{1}x_{2}}\underbrace{\left(\frac{1}{1+s}+s\frac{d}{ds}\frac{1}{1+s}\right)}_{\equiv f_{2}(s)}.

Likewise we have

ϕx1x2x3=sx1x2x3(1(1+s)2+sdds1(1+s)2)f3(s).\phi_{x_{1}x_{2}x_{3}}=s_{x_{1}x_{2}x_{3}}\underbrace{\left(\frac{1}{(1+s)^{2}}+s\frac{d}{ds}\frac{1}{(1+s)^{2}}\right)}_{\equiv f_{3}(s)}.

Hence in general the recursion relationship is

fn+1(s)=fn(s)+sddsfn(s)=dds(sfn(s)),f_{n+1}(s)=f_{n}(s)+s\frac{d}{ds}f_{n}(s)=\frac{d}{ds}(sf_{n}(s)),

where fn=Pn(1+s)n.f_{n}=\frac{P_{n}}{(1+s)^{n}}.

Alternatively, by ϕ(s)=k=1(1)k1skk,\phi(s)=\sum_{k=1}^{\infty}(-1)^{k-1}\dfrac{s^{k}}{k}, we compute

ϕx1=k=1(1)k1sk1sx1,\phi_{x_{1}}=\sum_{k=1}^{\infty}(-1)^{k-1}s^{k-1}s_{x_{1}},
ϕx1x2=k=1(1)k1{(k1)sk2sx2sx1+sk1sx1x2}=k=1(1)k1ksk1sx1x2,\phi_{x_{1}x_{2}}=\sum_{k=1}^{\infty}(-1)^{k-1}\{(k-1)s^{k-2}s_{x_{2}}s_{x_{1}}+s^{k-1}s_{x_{1}x_{2}}\}=\sum_{k=1}^{\infty}(-1)^{k-1}ks^{k-1}s_{x_{1}x_{2}},

and

ϕx1x2x3=k=1(1)k1{k(k1)sk2sx3sx1x2+ksk1sx1x2x3}=sx1x2x3k=1(1)k1k2sk1.\phi_{x_{1}x_{2}x_{3}}=\sum_{k=1}^{\infty}(-1)^{k-1}\{k(k-1)s^{k-2}s_{x_{3}}s_{x_{1}x_{2}}+ks^{k-1}s_{x_{1}x_{2}x_{3}}\}=s_{x_{1}x_{2}x_{3}}\sum_{k=1}^{\infty}(-1)^{k-1}k^{2}s^{k-1}.

In general we find

ϕx1x2xn=sx1x2xnk=1(1)k1kn1sk1,\phi_{x_{1}x_{2}\ldots x_{n}}=s_{x_{1}x_{2}\ldots x_{n}}\sum_{k=1}^{\infty}(-1)^{k-1}k^{n-1}s^{k-1},

and hence

fn(s)=k=1kn1(s)k1.f_{n}(s)=\sum_{k=1}^{\infty}k^{n-1}(-s)^{k-1}.\;\;\square

Using the general expression we can estimate the nonlinearity NN in nn-dimensional cases. Write Pn(R)=1+cnR+P_{n}(-R)=1+c_{n}R+\ldots for small R,R, we have

Pn(R)(1R)n(1+cnR+)(1+nR+)1+(cn+n)R+,\frac{P_{n}(-R)}{(1-R)^{n}}\approx(1+c_{n}R+\ldots)(1+nR+\ldots)\approx 1+(c_{n}+n)R+\ldots,
R(𝝃)(a2πν)n/2M1n2ν(πν2a)n/2Re2n+1,R(\bm{\xi})\approx\left(\frac{a}{2\pi\nu}\right)^{n/2}\frac{M_{1\ldots n}}{2\nu}\left(\frac{\pi\nu}{2a}\right)^{n/2}\approx\frac{Re}{2^{n+1}},

so we find

n1U1ξ2ξn1(a2πν)n/2M1n(1+cn+n2n+1Re+).\frac{\partial^{n-1}U_{1}}{\partial\xi_{2}\ldots\partial\xi_{n-1}}\approx\left(\frac{a}{2\pi\nu}\right)^{n/2}M_{1\ldots n}\left(1+\frac{c_{n}+n}{2^{n+1}}Re+\ldots\right).

As cn=2n1n,c_{n}=2^{n-1}-n, we deduce N=cn+n2n+1=14N=\frac{c_{n}+n}{2^{n+1}}=\frac{1}{4} for all n(1).n\;(\geq 1). Hence the estimate obtained in the iteration (1) in Section 2(c) holds valid in any dimensions.

Appendix B Derivation of the vorticity curl equations

Recalling vector identities

(𝑨𝑩)=𝑨𝑩+𝑩𝑨+𝑨×rot𝑩+𝑩×rot𝑨\nabla(\bm{A}\cdot\bm{B})=\bm{A}\cdot\nabla\bm{B}+\bm{B}\cdot\nabla\bm{A}+\bm{A}\times{\rm rot}\bm{B}+\bm{B}\times{\rm rot}\bm{A}
rot(𝑨×𝑩)=𝑨𝑩+𝑩𝑨+𝑨div𝑩+𝑩div𝑨{\rm rot}(\bm{A}\times\bm{B})=-\bm{A}\cdot\nabla\bm{B}+\bm{B}\cdot\nabla\bm{A}+\bm{A}{\rm div}\bm{B}+\bm{B}{\rm div}\bm{A}

and adding them and solving for 𝑩𝑨,\bm{B}\cdot\nabla\bm{A}, we have

𝑩𝑨=12{(𝑨𝑩)+rot(𝑨×𝑩)𝑨×rot𝑩𝑨div𝑩𝑩×rot𝑨+𝑩div𝑨}.\bm{B}\cdot\nabla\bm{A}=\frac{1}{2}\left\{\nabla(\bm{A}\cdot\bm{B})+{\rm rot}(\bm{A}\times\bm{B})-\bm{A}\times{\rm rot}\bm{B}-\bm{A}{\rm div}\bm{B}-\bm{B}\times{\rm rot}\bm{A}+\bm{B}{\rm div}\bm{A}\right\}.

Taking 𝑨=𝝎,𝑩=𝒖,\bm{A}=\bm{\omega},\bm{B}=\bm{u}, we find

(𝒖)𝝎=12{(𝒖𝝎)+×(𝝎×𝒖)𝒖×𝝌}.(\bm{u}\cdot\nabla)\bm{\omega}=\frac{1}{2}\left\{\nabla(\bm{u}\cdot\bm{\omega})+\nabla\times(\bm{\omega}\times\bm{u})-\bm{u}\times\bm{\chi}\right\}.

We then compute

𝝌t\displaystyle\frac{\partial\bm{\chi}}{\partial t} =\displaystyle= ×{×(𝒖×𝝎)}\displaystyle\nabla\times\left\{\nabla\times(\bm{u}\times\bm{\omega})\right\}
=\displaystyle= ×(𝝎)𝒖×(𝒖)𝝎\displaystyle\nabla\times(\bm{\omega}\cdot\nabla)\bm{u}-\nabla\times(\bm{u}\cdot\nabla)\bm{\omega}
=\displaystyle= ×(𝝎)𝒖12×{(𝒖𝝎)+×(𝝎×𝒖)𝒖×𝝌}\displaystyle\nabla\times(\bm{\omega}\cdot\nabla)\bm{u}-\frac{1}{2}\nabla\times\left\{\nabla(\bm{u}\cdot\bm{\omega})+\nabla\times(\bm{\omega}\times\bm{u})-\bm{u}\times\bm{\chi}\right\}
=\displaystyle= ×(𝝎)𝒖+12××(𝒖×𝝎)¯+12×(𝒖×𝝌),\displaystyle\nabla\times(\bm{\omega}\cdot\nabla)\bm{u}+\frac{1}{2}\underline{\nabla\times\nabla\times(\bm{u}\times\bm{\omega})}+\frac{1}{2}\nabla\times(\bm{u}\times\bm{\chi}),

Identifying the underlined part as the right-hand side, we obtain

𝝌t=2×(𝝎)𝒖+×(𝒖×𝝌),\frac{\partial\bm{\chi}}{\partial t}=2\nabla\times(\bm{\omega}\cdot\nabla)\bm{u}+\nabla\times(\bm{u}\times\bm{\chi}),

that is,

D𝝌Dt=𝝌𝒖+2×(𝝎)𝒖.\frac{D\bm{\chi}}{Dt}=\bm{\chi}\cdot\nabla\bm{u}+2\nabla\times(\bm{\omega}\cdot\nabla)\bm{u}.

Appendix C Steady Fokker-Planck equation

C.1 One-dimensional case

Because

2Uξ2+aνξ(ξU)=0\frac{\partial^{2}U}{\partial\xi^{2}}+\frac{a}{\nu}\frac{\partial}{\partial\xi}\left(\xi U\right)=0 (28)

is a second-order equation, its general solution has two constants of integration. We solve it paying attention to the boundary conditions. First, upon integration we have

Uξ+aνξU=C,\frac{\partial U}{\partial\xi}+\frac{a}{\nu}\xi U=C,

for some constant CC. If UU decays sufficiently rapidly as |ξ||\xi|\to\infty, then C=0C=0. Otherwise it’s possible to have C0C\neq 0. Let us proceed keeping CC and write

ea2νξ2ξ(ea2νξ2U(ξ))=C.e^{-\frac{a}{2\nu}\xi^{2}}\partial_{\xi}(e^{\frac{a}{2\nu}\xi^{2}}U(\xi))=C.

A further integration gives

U(ξ)\displaystyle U(\xi) =\displaystyle= Cexp(a2νξ2)+Cexp(a2νξ2)0ξexp(a2νη2)𝑑η,\displaystyle C^{\prime}\exp\left(-\frac{a}{2\nu}\xi^{2}\right)+C\exp\left(-\frac{a}{2\nu}\xi^{2}\right)\int_{0}^{\xi}\exp\left(\frac{a}{2\nu}\eta^{2}\right)d\eta,
=\displaystyle= Cexp(a2νξ2)+C2νaD(a2νξ),\displaystyle C^{\prime}\exp\left(-\frac{a}{2\nu}\xi^{2}\right)+C\sqrt{\frac{2\nu}{a}}D\left(\sqrt{\frac{a}{2\nu}}\xi\right),

with another constant CC^{\prime}. Here we have assumed that |U|,|ξU|0|U|,|\partial_{\xi}U|\to 0 as |ξ|.|\xi|\to\infty. However, if ξU(ξ)const,\xi U(\xi)\to\mbox{const}, the second term survives with C0C\neq 0. The function

D(x)ex20xey2𝑑yD(x)\equiv e^{-x^{2}}\int_{0}^{x}e^{y^{2}}dy

is the Dawson’s integral [18], which behaves 12x\approx\frac{1}{2x} as |x|.|x|\to\infty. It also satisfies

D(x)=π2H[ex2],D(x)=\frac{\sqrt{\pi}}{2}H\left[e^{-x^{2}}\right],

where H[]H[\cdot] denotes the Hilbert transform

H[f]=1πp.v.f(y)xy𝑑y.H[f]=\frac{1}{\pi}{\rm p.v.}\int_{-\infty}^{\infty}\frac{f(y)}{x-y}dy.

As exp(a2νξ2)0ξexp(a2νη2)𝑑η=πν2aH[exp(a2νξ2)],\exp\left(-\frac{a}{2\nu}\xi^{2}\right)\int_{0}^{\xi}\exp\left(\frac{a}{2\nu}\eta^{2}\right)d\eta=\sqrt{\frac{\pi\nu}{2a}}H\left[\exp\left(-\frac{a}{2\nu}\xi^{2}\right)\right], we can write

U(ξ)=Ca2πνexp(a2νξ2)+CH[a2πνexp(a2νξ2)],U(\xi)=C^{\prime}\sqrt{\frac{a}{2\pi\nu}}\exp\left(-\frac{a}{2\nu}\xi^{2}\right)+CH\left[\sqrt{\frac{a}{2\pi\nu}}\exp\left(-\frac{a}{2\nu}\xi^{2}\right)\right],

with suitably redefined constants C,CC,C^{\prime}. This shows the general solution consists of two kinds of solutions, which we call a source-type solution and a kink-type one. The former converges to the Dirac mass and the latter to the Cauchy kernel 1/x1/x in the limit of a/ν.a/\nu\to\infty.

It is of interest to note that if a solution is found, then its Hilbert transform gives the other solution. In fact applying H[]H[\cdot] to (28)

ξ2H[U]+aνξ(H[ξU])=0.\partial_{\xi}^{2}H[U]+\frac{a}{\nu}\partial_{\xi}(H[\xi U])=0.

By H[ξU]=ξH[U]1π1U(ξ)𝑑ξ,H[\xi U]=\xi H[U]-\frac{1}{\pi}\int_{\mathbb{R}^{1}}U(\xi)d\xi, e.g. [19], it follows that

ξ2H[U]+aνξ(ξH[U])=0,\partial_{\xi}^{2}H[U]+\frac{a}{\nu}\partial_{\xi}(\xi H[U])=0,

which shows that H[U]H[U] also a solves the same equation. See the Fig.4 for a comparison of those fundamental solutions.

Refer to caption
Figure 4: Comparison of exp(ξ2)\exp(-\xi^{2}) (dashed) with D(ξ)=π2H[exp(ξ2)]D(\xi)=\frac{\sqrt{\pi}}{2}H[\exp(-\xi^{2})] (solid).
initial data non-Gaussian steady solutions
1D Burgers u1xB1,0,L1u\propto\frac{1}{x}\in B^{0}_{1,\infty},\,\notin L^{1} uxLloc1u\propto x\in L^{1}_{\rm loc}, continuous
2D Navier-Stokes ω1r2B1,0,L1(u1rB2,0,L2){\omega\propto\frac{1}{r^{2}}\in B^{0}_{1,\infty},\,\notin L^{1}\atop\left(u\propto\frac{1}{r}\in B^{0}_{2,\infty},\,\notin L^{2}\right)} ωlogrLloc1\omega\propto\log r\in L^{1}_{\rm loc}, discontinuous
3D Navier-Stokes χ1r3B1,0,L1(u1rB3,0,L3){\chi\propto\frac{1}{r^{3}}\in B^{0}_{1,\infty},\notin L^{1}\atop\left(u\propto\frac{1}{r}\in B^{0}_{3,\infty},\,\notin L^{3}\right)} χ1rLloc1\chi\propto\frac{1}{r}\in L^{1}_{\rm loc}, discontinuous
Table 1: The behaviour of the initial data and that of non-Gaussian steady solutions near the origin. Note that the latter behave just like Green’s functions for the Poisson equations.

C.2 Two-dimensional case

In the case of the 1D Burgers equation the singular self-similar initial condition becomes continuous after infinitesimal time evolution. The second steady solution, which is not Gaussian, is also continuous but not in L1(1)L^{1}(\mathbb{R}^{1}). We should take it into account when we discuss long-time evolution.

For the 2D (and 3D) Navier-Stokes equations, the other steady solutions of non-Gaussian form, are in fact neither continuous nor in L1L^{1}. They are discontinuous at the origin. Hence care should be taken in considering them when we discuss long-time evolution in a larger function space such as B1,0B^{0}_{1,\infty}.

Refer to caption
Figure 5: Comparison of exp(r2)\exp(-r^{2}) (dashed) with 12exp(r2)(Ei(r2)γ)\frac{1}{2}\exp(-r^{2})({\rm Ei}(r^{2})-\gamma) (solid) for a2ν=1.\frac{a}{2\nu}=1., where γ=0.57721\gamma=0.57721\ldots is the Euler’s constant. The latter behaves as 1/r21/r^{2} as rr\to\infty and logr\log r as r0+r\to 0+.
Refer to caption
Figure 6: Comparison of exp(r2)\exp(-r^{2}) (dashed) with 2D(r)1r2D(r)-\frac{1}{r} (solid) for a2ν=1.\frac{a}{2\nu}=1. The latter behaves 1/r31/r^{3} as rr\to\infty and 1/r-1/r as r0+r\to 0+.

Under the assumption of radial symmetry the Fokker-Planck equation in two dimensions takes the following form

1rddr(rdfdr)+aνrddr(r2f)=0,\frac{1}{r}\frac{d}{dr}\left(r\frac{df}{dr}\right)+\frac{a}{\nu r}\frac{d}{dr}\left(r^{2}{\color[rgb]{0,0,0}f}\right)=0,

which is equivalent to

(ddr+1r)(dfdr+aνrf)=0.\left(\frac{d}{dr}+\frac{1}{r}\right)\left(\frac{df}{dr}+\frac{a}{\nu}rf\right)=0.

Indeed,

f′′+1rf+aν(2f+rf)=0,f^{\prime\prime}+\frac{1}{r}f^{\prime}+\frac{a}{\nu}\left(2f+rf^{\prime}\right)=0,
f′′+aν(f+rf)+1r(f+aνrf)=0,f^{\prime\prime}+\frac{a}{\nu}\left(f+rf^{\prime}\right)+\frac{1}{r}\left(f^{\prime}+\frac{a}{\nu}rf\right)=0,
(f+aνrf)+1r(f+aνrf)=0.\left(f^{\prime}+\frac{a}{\nu}rf\right)^{\prime}+\frac{1}{r}(f^{\prime}+\frac{a}{\nu}rf)=0.\;\;\square

Setting h(r)=f+aνrf,h(r)=f^{\prime}+\frac{a}{\nu}rf, we have h(r)=C/r,h(r)=C/r, that is,

exp(a2νr2){fexp(a2νr2)}=Cr.\exp\left(-\frac{a}{2\nu}r^{2}\right)\left\{f\exp\left(\frac{a}{2\nu}r^{2}\right)\right\}^{\prime}=\frac{C}{r}.

Hence

f\displaystyle f =\displaystyle= c1exp(ar22ν)+c2exp(ar22ν)f.p.0rexp(as22ν)dss\displaystyle c_{1}\exp\left(-\frac{ar^{2}}{2\nu}\right)+c_{2}\exp\left(-\frac{ar^{2}}{2\nu}\right){\rm f.p.}\int_{0}^{r}\exp\left(\frac{as^{2}}{2\nu}\right)\frac{ds}{s}
=\displaystyle= c1exp(ar22ν)+c22exp(ar22ν)(Ei(ar22ν)log(a2ν)γ),\displaystyle c_{1}\exp\left(-\frac{ar^{2}}{2\nu}\right)+\frac{c_{2}}{2}\exp\left(-\frac{ar^{2}}{2\nu}\right)\left({\rm Ei}\left(\frac{ar^{2}}{2\nu}\right)-\log\left(\frac{a}{2\nu}\right)-\gamma\right),

where Ei(x)=p.v.xett𝑑t{\rm Ei}(x)=-{\rm p.v.}\int_{-x}^{\infty}\frac{e^{-t}}{t}dt denotes the exponential integral and γ0.577\gamma\approx 0.577 the Euler’s constant.

Derivation
Define

Iλeλr2ϵreλs2s𝑑s.I\equiv\lambda e^{-\lambda r^{2}}\int_{\epsilon}^{r}\frac{e^{-\lambda s^{2}}}{s}ds.

We have

ϵreλs2s𝑑s=[logseλs2]ϵrϵrlogs 2λseλs2ds\int_{\epsilon}^{r}\frac{e^{-\lambda s^{2}}}{s}ds=\left[\log se^{-\lambda s^{2}}\right]_{\epsilon}^{r}-\int_{\epsilon}^{r}\log s\,2\lambda se^{-\lambda s^{2}}ds
=eλr2logrlogϵ2λϵrslogseλs2=(A)ds.=e^{-\lambda r^{2}}\log r-\log\epsilon-\underbrace{2\lambda\int_{\epsilon}^{r}s\log se^{-\lambda s^{2}}}_{=(A)}ds.

Now, with u=eλs2,ϵ=eλϵ21u=e^{-\lambda s^{2}},\epsilon^{\prime}=e^{\lambda\epsilon^{2}}\approx 1

(A)=eλϵ2eλr2log{1λ(logu)1/2}𝑑u(A)=\int_{e^{\lambda\epsilon^{2}}}^{e^{-\lambda r^{2}}}\log\left\{\frac{1}{\sqrt{\lambda}}(\log u)^{1/2}\right\}du
=eλϵ2eλr2log1λdu+12eλϵ2eλr21loglogudu.=\int_{e^{\lambda\epsilon^{2}}}^{e^{-\lambda r^{2}}}\log\frac{1}{\sqrt{\lambda}}du+\frac{1}{2}\int_{e^{\lambda\epsilon^{2}}}^{e^{-\lambda r^{2}}}1\log\log udu.

As

eλϵ2eλr21loglogudu=[uloglogu]eλϵ2eλr2eλϵ2eλr2dulogu\int_{e^{\lambda\epsilon^{2}}}^{e^{-\lambda r^{2}}}1\log\log udu=\left[u\log\log u\right]_{e^{\lambda\epsilon^{2}}}^{e^{-\lambda r^{2}}}-\int_{e^{\lambda\epsilon^{2}}}^{e^{-\lambda r^{2}}}\frac{du}{\log u}
eλr2log(λr2)logλ2logϵli(eλr2)+li(eλϵ2),\approx e^{\lambda r^{2}}\log(\lambda r^{2})-\log\lambda-2\log\epsilon-{\rm li}(e^{\lambda r^{2}})+{\rm li}(e^{\lambda\epsilon^{2}}),

we find

(A)12logλ+eλr2logr12li(eλr2)+γ2.(A)\approx\frac{1}{2}\log\lambda+e^{\lambda r^{2}}\log r-\frac{1}{2}{\rm li}(e^{\lambda r^{2}})+\frac{\gamma}{2}.

Hence

ϵreλs2s𝑑s=12logλ+12li(eλr2)γ2logϵ,\int_{\epsilon}^{r}\frac{e^{-\lambda s^{2}}}{s}ds=-\frac{1}{2}\log\lambda{\color[rgb]{0,0,0}+}\frac{1}{2}{\rm li}(e^{\lambda r^{2}})-\frac{\gamma}{2}-\log\epsilon,

that is,

I=λeλr2ϵreλs2s𝑑s=λ2eλr2{li(eλr2)logλγ2logϵ}.I=\lambda e^{-\lambda r^{2}}\int_{\epsilon}^{r}\frac{e^{-\lambda s^{2}}}{s}ds=\frac{\lambda}{2}e^{-\lambda r^{2}}\left\{{\rm li}(e^{\lambda r^{2}})-\log\lambda-\gamma-2\log\epsilon\right\}.

Taking its f.p.,

I=λ2eλr2{li(eλr2)logλγ}.I=\frac{\lambda}{2}e^{-\lambda r^{2}}\left\{{\rm li}(e^{\lambda r^{2}})-\log\lambda-\gamma\right\}.

Here

li(x)p.v.0xdtlogt,(x>1){\rm li}(x)\equiv{\rm p.v.}\int_{0}^{x}\frac{dt}{\log t},\;(x>1)

denotes the logarithmic integral and

Ei(x)γ+logx+xasx0+{\rm Ei}(x)\approx\gamma+\log x+x\;\;\,\mbox{as}\;\;x\to 0+
Ei(x)(li(ex))=O(exx)asx.{\rm Ei}(x)\left(\equiv{\rm li}(e^{x})\right)=O\left(\dfrac{e^{x}}{x}\right)\>\>\,\mbox{as}\;\;x\to\infty.

C.3 Three-dimensional case

Under the assumption of radial symmetry, the Fokker-Planck equation f+aν(𝝃f)=0\triangle f+\frac{a}{\nu}\nabla\cdot(\bm{\xi}f)=0 in three dimensions takes the following form

1r2ddr(r2dfdr)+aν1r2ddr(r3f)=0,\frac{1}{r^{2}}\frac{d}{dr}\left(r^{2}\frac{df}{dr}\right)+\frac{a}{\nu}\frac{1}{r^{2}}\frac{d}{dr}\left(r^{3}{\color[rgb]{0,0,0}f}\right)=0,

where r=|𝝃|.r=|\bm{\xi}|. It is equivalent to

(ddr+2r)(dfdr+aνrf)=0.\left(\frac{d}{dr}+\frac{2}{r}\right)\left(\frac{df}{dr}+\frac{a}{\nu}rf\right)=0.

Indeed,

f′′+2rf+aν(3f+rf)=0,f^{\prime\prime}+\frac{2}{r}f^{\prime}+\frac{a}{\nu}(3f+rf^{\prime})=0,
(f′′+aν(f+rf))+(2rf+2aνf)=0,\left(f^{\prime\prime}+\frac{a}{\nu}(f+rf^{\prime})\right)+\left(\frac{2}{r}f^{\prime}+\frac{2a}{\nu}f\right)=0,
(f+aνrf)+2r(f+aνrf)=0.\left(f^{\prime}+\frac{a}{\nu}rf\right)^{\prime}+\frac{2}{r}(f^{\prime}+\frac{a}{\nu}rf)=0.\;\;\square

Setting h(r)=dfdr+aνrf,h(r)=\frac{df}{dr}+\frac{a}{\nu}rf, we have h(r)=C/r2.h(r)=C/r^{2}. Thus we can write

exp(a2νr2){fexp(a2νr2)}=Cr2.\exp\left(-\frac{a}{2\nu}r^{2}\right)\left\{f\exp\left(\frac{a}{2\nu}r^{2}\right)\right\}^{\prime}=\frac{C}{r^{2}}.

A further integration gives

f=c1exp(ar22ν)+c2exp(ar22ν)f.p.0rexp(as22ν)dss2f=c_{1}\exp\left(-\frac{ar^{2}}{2\nu}\right)+c_{2}\exp\left(-\frac{ar^{2}}{2\nu}\right){\rm f.p.}\int_{0}^{r}\exp\left(\frac{as^{2}}{2\nu}\right)\frac{ds}{s^{2}}
=c1exp(ar22ν)+c2(2aνD(a2νr)1r),=c_{1}\exp\left(-\frac{ar^{2}}{2\nu}\right)+c_{2}\left(\sqrt{\dfrac{2a}{\nu}}D\left(\sqrt{\frac{a}{2\nu}}r\right)-\frac{1}{r}\right),

where D()D(\cdot) denotes the Dawson’s integral.

Derivation
As

ϵreλs2dss2=[eλs2s]ϵr+2λϵreλs2𝑑s,\int_{\epsilon}^{r}e^{\lambda s^{2}}\frac{ds}{s^{2}}=\left[-\frac{e^{\lambda s^{2}}}{s}\right]_{\epsilon}^{r}+2\lambda\int_{\epsilon}^{r}e^{\lambda s^{2}}ds,

we have

eλr2ϵreλs2dss2=1r+eλr2ϵ+2λeλr2ϵreλs2𝑑s.e^{-\lambda r^{2}}\int_{\epsilon}^{r}e^{\lambda s^{2}}\frac{ds}{s^{2}}=-\frac{1}{r}+\frac{e^{-\lambda r^{2}}}{\epsilon}+2\lambda e^{-\lambda r^{2}}\int_{\epsilon}^{r}e^{\lambda s^{2}}ds.

Noting

eλr20reλs2𝑑s=1λD(λr)e^{-\lambda r^{2}}\int_{0}^{r}e^{\lambda s^{2}}ds=\frac{1}{\sqrt{\lambda}}D\left(\sqrt{\lambda}r\right)

and dropping the O(ϵ1)O(\epsilon^{-1}) term, we obtain the desired form.

Acknowledgement
This work has been supported by EPSRC grant EP/N022548/1. Its major part was carried out when one of the authors (K.O.) was affiliated with School of Mathematics and Statistics, the University of Sheffield.


References

  • [1] Barenblatt GI. 2003 Scaling Cambridge: Cambridge University Press.
  • [2] Dresner L. 1983 Similarity Solutions of Nonlinear Partial Differential Equations Boston: Pitman Publishing.
  • [3] Cannone M and Planchon F. 1996. Self-similar solutions for Navier-Stokes equations in 3\mathbb{R}^{3}. Commun. in P.D.E. 21, 179–193.
  • [4] Jia H and Šverák V. 2014. Local-in-space estimates near initial time for weak solutions of the Navier-Stokes equations and forward self-similar solutions. Invent. Math. 196, 233–265.
  • [5] Brandolese L. 2009 Fine properties of self-similar solutions of the Navier–Stokes equations. Arch. Rat. Mech. Anal. 192, 375–401.
  • [6] Bradshaw Z and Tsai T-P. 2019. Discretely self-similar solutions to the Navier–Stokes equations with data in Lloc2 satisfying the local energy inequality. Anal. PDE 12, 1943–1962.
  • [7] Chae D and Wolf J. 2017. Removing discretely self-similar singularities for the 3D Navier–Stokes equations Commun. Partial Diff. Eqs. 42, 1359–1374.
  • [8] Hairer M and Labbé C. 2017. The reconstruction theorem in Besov spaces. J. Funct. Anal. 273, 2578–2618.
  • [9] Burgers JM. 1948. A mathematical model illustrating the theory of turbulence Adv. Appl. Mech. 1, 171–199.
  • [10] Escobedo M and Zuazua E. 1991. Large time behavior for convection-diffusion equations in RNR^{N}. J. Func. Anal. 100, 119–161.
  • [11] Biler P, Karch G and Woyczyński WA. 1999. Asymptotics for multifractal conservation laws. Studia Mathematica 135, 231-252
  • [12] Biler P, Karch G and Woyczyński WA. 2001. Critical nonlinearity exponent and self-similar asymptotics for Lévy conservation laws. Annal. IHP Anal. nonlin. 18, 613–637.
  • [13] Bureau FJ. 1955. Divergent integrals and partial differential equations. Commun Pure Appl. Math. 8 143–202.
  • [14] Yosida K. 1956 An operator-theoretical integration of the wave equation. J. Math. Soc. Jpn. 8 79–92.
  • [15] Ohkitani K. 2020. Study of the Hopf functional equation for turbulence: Duhamel principle and dynamical scaling. Phys. Rev. E 101, 013104-1–15.
  • [16] Gallay T and Wayne CE. 2005. Global stability of vortex solutions of the two-dimensional Navier-Stokes equation. Commun. Math. Phys. 255, 97–129.
  • [17] Ohkitani K. 2015. Dynamical equations for the vector potential and the velocity potential in incompressible irrottational Euler flows: a refined Bernoulli theorem. Phys. Rev. E. 92, 033010-1–8.
  • [18] Olver FW, Lozier DW, Boisvert RF and Clark CW. (eds.) 2010 NIST Handbook of Mathematical Functions Cambridge: Cambridge University Press.
  • [19] King FW. 2009. Hilbert transforms Cambridge: Cambridge University Press.