Navier-Stokes equations paradox

https://doi.org/10.1016/S0034-4877(21)00054-9Get rights and content

It is proved that if the Navier-Stokes problem is considered in ℝ3, then the initial velocity has to be zero. This is the paradox. The consequences of this paradox are: a) The NSP is contradictory physically and mathematically; it is not a correct formulation of the problem of motion of the incompressible viscous fluid. b) The solution to the NSP does not exist on [0, ∞). This solves one of the millennium problems.

Introduction

It is of general interest to find out whether the equations describing the fluid are correct. What does it mean that the equations are correct? It means that: a) their solution is unique (in a suitable space), b) the solution does exist (in this space), c) there are no consequences of these equations that contradict physical assumptions. Existence of such consequences we call paradoxes.

The Navier-Stokes (NS) equations (NS Problem-NSP) describe the non-compressible viscous fluid. It is proved that the solution to NSP is unique (in various spaces) [3, 5, 7, 8]. The aim of this paper is to describe a paradox in the NSP and its consequences: the NSP is contradictory and the solution to the NSP does not exists on [0, ∞), see also [5, 11, 12].

First we outline the theory. The NSP in ℝ3 consists of solving the equations: v+(v,)v=f-p+vΔv,v=0,v(x,0)=v0(x).Here v = v(x, t) is the fluid velocity, v’ := dv/dt, f = f (x, t) is the exterior force, p = p (x, t) is the pressure, the fluid density ρ = 1, v = const > 0 is the viscosity coefficient, x3, t > 0. The f, v0 are known, v, p are unknown. It is assumed that f, v0 are smooth and rapidly decaying functions of their arguments, v, p decay

at infinity. It is proved in [5], that Eqs. (1) are equivalent to v=F-0tds3G(x-y,t-s)(v-)vdy, F=0tds3G(x-y,t-s)f(y,s)dy+3g(x-y,t-s)v0(y)dy, g(x,t)=e-|x|24vt(4vπt)3/2,g(x,0)=δ(x);g(x,t)=0ift<0.

Here δ(x) is the delta-function and G is the tensor solving the equations: G-vΔG=δ(x)δ(t)δjm-pm,G=0,G=0ift<0,where δjm is the Kronecker symbol, pm = p(x, t)em, ej, 1 ≤ j ≤ 3, is an orthonormal basis in ℝ3, p = 0 if t < 0, ∇ pm = pxj, ej em is a tensor. Tensor G is calculated in [5] explicitly and the equivalence of (1) and (2) is proved there. Uniqueness of the solution to (2) in various functional spaces is proved in [5, 8]. Let us outline some ideas of our proofs. Fourier transform (2) with respect to x and get v˜=F˜-0tdsG˜(ξ,t-s)v˜*iξv˜,where * is the convolution in ℝ3, v˜=(2π)-33e-iξxv(x,t)dx, G˜=(2π)-3(δjm-ξjξm(ξ2)-1)e-vξ2t,see [5, p. 11]. One has |v*iξv˜|v˜|ξ|v˜,where=L2(3). Also, |G˜|ce-vξ2t where c > 0 stands for various constants independent of t and ξ, G˜(ξ,t-s)c(t-s)-34, |ξ|G˜(ξ,t-s)c(t-s)-54.

Denote |ξ|v˜:=b(t)0,|ξ|F˜:=b0(t)0. Multiply (6) by |ξ|, take the norm of both sides of (6) and use (8) to get b(t)b0(t)+c0t(t-s)-54b(s)ds.

The integral 0t(t-s)λ-1b(s)ds:=tλ-1*b is the convolution of the distribution tλ-1 and b(t) in ℝ+ = [0, ∞). We assume that b = b(t) is continuous on ℝ+ and tλ = 0 for t < 0. The integral tλ-1 *b diverges classically if λ ≤ 0. Integral (9) with λ = -1/4 diverges classically. In [5, 9, 10, 11, 12] such integrals are defined in a new way. Let us define it for λ ≤ 0 by the formula L(tλ-1*b)=L(tλ-1)L(b).

Here L is the Laplace transform, L(b)=0e-ptb(t)dt. One has L(tλ-1)=Γ(λ)pλ,λ0,-1,-2,see [5, 1]. Here Γ(λ) is the gamma-function, which is analytic with respect to λ ∈ ℂ except for λ = -n, n = 0, 1, 2,…. At the point λ = -n the Γ(λ) has a simple pole with the residue (-1)n/n!, 1/Γ(λ) is an entire function, Γ(z + 1) = zΓ(z), Γ(z(1 - z) = π/sin(πz), see [4]. Let Φλ:= tλ-1/Γ(A). Then Lλ) = p−λ for all λ ∈ ℂ. For λ ≤ 0 one has L(Φλ*b)=L(b)/pλ,Φλ*b=L-1(L(b)/pλ).

This formula defines the convolution for λ < 0, that is, for the values of λ for which the convolution is not defined classically, in particular, for λ=-14. We avoided the usage of distribution theory, cf. [2].

Lemma 1. One has Φλ*Φμ=Φλ+μ,limλ+0Φλ=δ(t),where δ(t) is the Dirac distribution.

Proof: Since Lλ * Φμ) = Lλ)Lμ) = 1/pλ+μ and L−1(1/pλ+μ) = Φλ+μ, the first part of Lemma 1 is proved. To prove the second part we use the formula Φλ*Φμ=1Γ(λ)Γ(μ)0t(t-s)λ-1sμ-1ds=tλ+μ-1Γ(λ)Γ(μ)B(λ,μ),where B(x, y) is the beta-function, B(x,y)=Γ(x)Γ(y)Γ(x+y)=01(1-u)x-1uy-1du,see [4]. Formula (14) gives an alternative derivation of the first formula (13) because the right side of (14) equals to Φλ+μ(t).

Let ϕ ∈ C∞(ℝ+) be an arbitrary function vanishing near infinity. Note that 1/Γ(λ) ~ λ as λ → 0. Therefore limλ+00Φλ(t)ϕ(t)dt=limλ+0λ0tλ-1ϕ(t)dt=ϕ(0).

Lemma 1 is proved. See [5, p. 24] for a different proof of Lemma 1.

Let us write (9) as b(t)b0(t)-cc1Φ-14*b,c1:=|Γ(-14)|>0.

Apply the operator Φ1/4* to (16) and use (13) to get Φ1/4*bΦ1/4*b0-cc1b,cc1>0.

If λ > 0 and g(t) ≥ 0, then Φ1/4* g ≥ 0. Since b ≥ 0 it follows from (17) that b(t)1cc1Φ1/4*b0.

Since b0(t) is a smooth bounded function on ℝ+, it follows that b(t) has these properties. We have proved the following result.

Theorem 1. If supt≥0 b0(t) < ∞, then supt>0b(t)<,and b(0)=0.

Proof: Formula (20) holds because limt→0 Φ¼ * b0 = 0.

Formula (20) is the NSP paradox: originally it was not assumed that v(x, 0) = v0(x) = 0, so b(0) ≠ 0, but we have derived that b(0) = 0, so v0(x) = 0.

Lemma 2. If b(t) ∈ C([0, T]), then 0t(t-s)54b(s)dst-14b(0)Γ(14)/Γ(3/4),t0.where the sign ~ stands for asymptotic equivalence.

Proof: One has 0t(t-s)54b(s)ds=t-140t(1-u)-54b(tu)dut-14b(0)B(-14,1).

Since b(tu) → b(0) uniformly with respect to u ∈ [0, 1] and B( -14, 1) = Γ( -14)/Γ(3/4), Lemma 2 is proved.

The integral 01(1-u)-54du diverges classically. We define it as B(-14,1). The beta-function B(x, y) is well defined for all x, y ∈ ℂ except for x, y = 0, -1, -2,…, so at the values x = -14 = 1 the function B(x, y) is well defined.

Lemma 3. Let b(t) ≥ 0 satisfy (9) and β(t) ≥ 0 satisfy the equation β(t)=b0(t)+c0t(t-s)-54β(s)ds.

Then b(t)β(t).

Proof: Let z(t):= β(t) - b(t). Subtract from (9) (23) and get 0z(t)-c0t(t-s)-54z(s)ds.

By formula (21) one gets from (25) that 0z(t)+cc1t-14(0)/Γ(3/4),t0.

Thus, it follows from (25) that z(0) ≥ 0 as t → 0.

Similarly, if there is a point t0 at which z(t0) < 0, then we have a contradiction with (25). A detailed and different proof of Lemma 3 is given in [5, p. 20].

Lemma 3 is proved.

Eq. (23) can be solved analytically, see [6, 9]. Take the Laplace transform of (23) and get L(β) = L(b0) - cc1p¼ L(β). Thus β(t)=L-1(L(b0)1+cc1p1/4).

Let us state a priori estimates, proved in [5], for the solution to (1) assuming that |ξ||F˜|<c: supt>0(v+0tb(s)ds)<c,supt>0b(t):=|ξ|v˜<c, |ξ|v˜<c,|v˜|c(1+t1/2).

References (12)

  • A.G. Ramm

    Solution of the Navier-Stokes problem

    Appl. Math. Lett.

    (2019)
  • Yu. Brychkov et al.

    Integral transforms of generalized functions

    (1977)
  • I. Gel'fandand et al.

    Generalized functions

    (1959)
  • O.A. Ladyzhenskaya

    The Mathematical Theory of Viscous Incompressible Flow

    (1969)
  • N. Lebedev

    Special Functions and their Applications

    (1972)
  • A.G. Ramm

    The Navier-Stokes Problem

    (2021)
There are more references available in the full text version of this article.
View full text