Open access peer-reviewed chapter

Relaxation Dynamics of Point Vortices

Written By

Ken Sawada and Takashi Suzuki

Reviewed: 22 September 2021 Published: 09 November 2021

DOI: 10.5772/intechopen.100585

From the Edited Volume

Vortex Dynamics - From Physical to Mathematical Aspects

Edited by İlkay Bakırtaş and Nalan Antar

Chapter metrics overview

256 Chapter Downloads

View Full Metrics

Abstract

We study a model describing relaxation dynamics of point vortices, from quasi-stationary state to the stationary state. It takes the form of a mean field equation of Brownian point vortices derived from Chavanis, and is formulated by our previous work as a limit equation of the patch model studied by Robert-Someria. This model is subject to the micro-canonical statistic laws; conservation of energy, that of mass, and increasing of the entropy. We study the existence and nonexistence of the global-in-time solution. It is known that this profile is controlled by a bound of the negative inverse temperature. Here we prove a rigorous result for radially symmetric case. Hence E/M2 large and small imply the global-in-time and blowup in finite time of the solution, respectively. Where E and M denote the total energy and the total mass, respectively.

Keywords

  • point vortex
  • quasi-equilibrium
  • relaxation dynamics

1. Introduction

Our purpose is to study the system

ωt+ωψ=ω+βωψinΩ×0T,ων+βωψν∂Ω=0,ωt=0=ω0xE1

with

Δψ=ωinΩ,ψ∂Ω=0,β=ΩωψΩωψ2,E2

where ΩR2 is a bounded domain with smooth boundary ∂Ω, ν is the outer unit normal vector on ∂Ω, and

=x1x2,=x2x1,x=x1x2.E3

The unknown ω=ωxtR stands for a mean field limit of many point vortices,

ωxtdx=i=1Nαiδxitdx.E4

It was derived, first, for Brownian point vortices by [1, 2], with β=βt standing for the inverse temperature. Then, [3, 4] reached it by the Lynden-Bell theory [5] of relaxation dynamics, that is, as a model describing the movement of the mean field of many point vortices, from quasi-stationary state to the stationary state. This model is consistent to the Onsager theory [6, 7, 8, 9, 10, 11, 12] on stationary states and also the patch model proposed by [13, 14], that is,

ωxt=i=1Npσi1Ωitx,E5

where Np, σi, and Ωit denote the number of patches, the vorticity of the i-th patch, and the domain of the i-th patch, respectively [15, 16, 17].

This chapter is concerned on the one-sided case of

ω0=ω0x>0.E6

If this initial value is smooth, there is a unique classical solution to (1)(4) local in time, denoted by ω=ωxt, with the maximal existence time T=Tmax0+. More precisely, the strong maximum principle to (1) guaranttes

ω=ωxt>0onΩ¯×0T.E7

Then, the Hopf lemma to the Poisson equation in (2) ensures

ψν∂Ω<0,E8

and hence the well-definedness of

β=ΩωψΩωψ2.E9

We confirm that system (1)(3) satisfies the requirements of isolated system of thermodynamics. First, the mass conservation is derived from (1) as

ddtΩω=0,E10

because

νψ∂Ω=0E11

holds by (2). Second, the energy conservation follows as

12ddtψ22=ψψt=ωtψ=ωψψω+βωψψ=ωψβΩωψ2=0E12

by (1) and (2), because

ψψ=0,E13

where , denotes the L2 inner product. Third, the entropy increasing is achieved, writing (1) as

ωt=ωψ+logω+βψ,νlogω+βψ∂Ω=0.E14

In fact, it then follows that

Ωωtlogω+βψ=Ωωψlogψ+βψωlogω+βψ2dxE15

with

Ωωψlogω+βψ=Ωωψ=∂ΩωνψΩωψ=0E16

from (11) and

==0.E17

Since

Ωωtlogω=ddtΩωlogω1,Ωωtψ=12ddtψ22=0,E18

We thus end up with the mass conservation

M=Ωω,E19

the energy conservation

E=ψ22=ψω,E20

and the entropy increasing

ddtΩωlogω1=Ωωlogω+βψ20.E21

Henceforth, C>0 stands for a generic constant. In the previous work [4] we studied radially symmetric solutions and obtained a criterion for the existence of the solution global in time. Here, we refine the result as follows, where B01 denotes the unit ball.

Theorem 1 Let

Ω=B01,ω0=ω0r,ω0r<0,0<r=x1.E22

Then there is C0>0 such that

C0ω023Eω¯T=+,ωtC,t0,E23

where

ω¯=minΩ¯ω0>0.E24

Theorem 2 Under the assumption of (22) there is δ0>0 such that

EM2<δ0T<+.E25

Remark 1 Since

ω023=Ωω023/2ω¯2/3Ωω04/33/2=ω¯Ωω04/33/2ω¯Ω1/2Ωω02=ω¯Ω1/2M2E26

the assumption (23) implies

EM2C0Ω1/2.E27

Therefore, roughly, the conditions E/M21 and E/M21 imply T=+ and T<+, respectively.

Remark 2 The assumption (22) implies

β=βt<0,0t<T,E28

and then we obtain Theorem 1. In other words, the conclusion of this theorem arises from (28), without (22).

Remark 3 Since

EM2=Ωψ2Ωω2E29

it holds that

EM2=c22,c=Δ1ω0Ωω0=ψ0∂Ωψ0ν,E30

where ψ0=Δ1ω0.

The system (1)(4) thus obays a profile of the micro-canonical ensemble. In a system associated with the canonical ensemble, the inverse temperature β is a constant in (1) independent of t, with the third equality in (2) elimiated:

ωt+ωψ=ω+βωψ,ων+βωψν∂Ω=0,ωt=0=ω0x>0Δψ=ω,ψ∂Ω=0.E31

Then there arise the mass conservation

ddtΩω=0,E32

and the free energy decreasing

ddtΩωlogω1+β2ψ2dx=Ωωlogω+βψ20.E33

The system (31) without vortex term,

ωt=ω+βωψ,ων+βωψν∂Ω=0,ωt=0=ω0x>0Δψ=ω,ψ∂Ω=0.E34

is called the Smoluchowski-Poisson equation. This model is concerned on the thermodynamics of self-gravitating Brownian particles [18] and has been studied in the context of chemotaxis [19, 20, 21, 22, 23]. We have a blowup threshold to (34) as a consequence of the quantized blowup mechanism [19, 23]. The results on the existence of the bounded global-in-time solution [24, 25, 26] and blowup of the solution in finite time [27] are valid even to the case that β is a function of t as in β=βt. provided with the vortex term ωψ on the right-hand side. We thus obtain the following theorems.

Theorem 3 It holds that

βtδ,ω01<8πδ1T=+,ωtCE35

in (31), where δ>0 is arbitrary.

Theorem 4 It holds that

βtδ,ω01>8πδ1ω0>0,ω01>8πδ1suchthatT<+E36

in (31), where δ>0 is arbitrary.

Remark 4 In the context of chemotaxis in biology, the boundary condition of ψ is required to be the form of Neumann zero. The Poisson equation in (34) is thus replaced by

Δψ=ω1ΩΩω,ψν∂Ω=0E37

or

Δψ+ψ=ω,ψν∂Ω=0E38

by [28] and [29], respectively. In this case there arises the boundary blowup, which reduces the value 8π in Theorems 3–4 to 4π. The value 8π in Theorems 3–4, therefore, is a consequence of the exclusion of the boundary blowup [30]. This property is valid even for (37) or (38) of the Poisson part, if (22) is assumed.

Remark 5 The requirement to ω0 in Theorem 4 is the concentration at an interior point, which is not necessary in the case of (22). Hence Theorems 3 and 4 are refined as

βtδ,ω01<8πδ1T=+,ωtCE39

and

βtδ,ω01>8πδ1T<+,E40

if (22) holds in (35). The main task for the proof of Theorems 1 and 2, therefore, is a control of β=βt in (1).

This paper is composed of four sections and an appendix. Section 2 is devoted to the study on the stationary solutions, and Theorems 1 and 2 are proven in Sections 3 and 4, respectively. Then Theorem 4 is confirmed in Appendix.

Advertisement

2. Stationary states

First, we take the canonical system (31) with β independent of t. By (32) and (33), its stationary state is defined by

logω+βψ=constant,ω=ωx>0,Ωω=M.E41

Then it holds that

ω=MeβψΩeβψE42

and hence

Δψ=MeβψΩeβψ,ψ∂Ω=0.E43

There arises an oredered structure arises in β<0, as observed by [11], as a consequence of a quantized blowup mechanism [19, 20, 31]. In the micro-canonical system (1) and (2), the value β in (43) has to be determined by E besides M.

Equality (21), however, still ensures (41) and hence (42) in the stationary state even for (1)(3). Writing

v=βψ,μ=βMΩeβψ,E44

we obtain

Δv=μevinΩ,v∂Ω=0,EM2=v22Ωvν2E45

by (30) and (43).

This system is the stationary state of (1) and (2) introduced by [4]. The first two equalities

Δv=μev,v∂Ω=0E46

comprise a nonlinear elliptic eigenvalue problem and the unknown eigenvalue μ is determined by the third equality,

EM2=v22Ωvν2.E47

The elliptic theory ensures rather deailed features of the set of solutions to (46). Here we note the following facts [31].

  1. There is μ¯=μΩ>0 such that the problem (46) does not admit a solution for μ>μ¯.

  2. Each μ0 admits a unique solution.

  3. Each 0<δ<μ¯ admits a constant C=Cδ>0 such that vC for any solution v=vx.

  4. There is a family of solutions μv such that μ0 and v+.

We show the following theorem, consistent to Theorem 2.

Theorem 5 If Ω=B01R2, there is δ>0 such that any solution vμ to (45) admits

EM2δ.E48

Proof: If μ=0, it holds that v=0. We have ±v>0 exclusively in Ω, provided that ±μ>0, respectively. By the elliptic theory [32], therefore, any solution v to (46) is radially symmetric as in v=vr, r=x. We have, furthermore, ±vr<0 in 0<r1, if ±μ>0, respectively.

Then it holds that ψ=ψr, and hence

1rrψrr=ωin0<r1,ψr=1=0E49

by (42) and (43), which implies

rψrr=0rsds>0,0<r1.E50

We thus obtain μ0, in particular.

If μ<0 we have β>0 by (44), and therefore, ψr>0 in 0<r1 by vr>0 there. It is a contradiction, and hence μ>0. In this case, the solution v=vr to (46) is explicit [31]. The numbers of the solution is 0, 1, and 2, according to μ>2, μ=2, and 0<μ<2, respectively, and if 0<μ2 the solutions v=v± are given as

v±r=log8γ±r1+γ±r22,γ±=4μ1μ4±1μ21/2.E51

In fact, we have γ+=γ for μ=2.

This solution is parametrized by

σ=Ωμev08π.E52

Hence each 0<σ<8π admits a unique solution vμ to (46), and v=v+ and v=v according as σ4π and σ4π, respectively. It holds also that μ0 if either σ8π or σ0. Thus we have only to confirm that E/M2 is bounded, both as σ8π and σ0.

As σ8π, we have

v=v+x4log1xlocallyuniformlyonΩ¯\0E53

and hence

v22+,∂Ωvν8π,E54

which implies

limσ8πEM2=+.E55

As σ0, on the other hand, we have

v=vx0uniformlyinΩ¯.E56

Since μ0, furthermore, there arises that

γ=γ=4μ1μ41μ21/2=μ1+o1.E57

It holds also that

vr=log8γμ2log1+μr2E58

and hence

vrr=4μr1+μr22=4μr1+o1uniformlyonΩ¯.E59

Then, (59) implies

v22=2π01vr2rdr=2π16μ201r3dr1+o1=8πμ21+o1E60

as well as

∂Ωvν2=16μ22π1+o1.E61

It thus follows that

limσ0EM2=14E62

and hence the conclusion.

Advertisement

3. Proof of Theorem 1

The first observation is the following lemma.

Lemma 1 Under the assumption of (22), it holds that

β=βt<0,ωrrt<0,0<r1,0t<T.E63

Proof: We have (7) and hence

ψrrt<0,0<r1,0t<TE64

by (49), which implies, in particular,

β=ωψΩωψ2<0E65

at t=0 by (22).

Since ω=ωrt and ψ=ψrt, we obtain ψ=0, and hence

ωt=ωrr+1rωr+βψrωrβω2E66

by (1). Then z=ωr satisfies

zt=zrr1r2z+1rzr+βψrrz+βψrzr2βωz,0<r1,0t<Tzr=0=0,zt=0=ω0rr<0,0<r1E67

and

z=βωψr,r=1,0t<T.E68

Putting

mt=min∂Ωzt=ωrtr=1,E69

we obtain m0<0 from the assumption. If there is 0<t0< such that

mt<0,0t<t0<T,mt0=0,E70

we obtain zrt>0 for 0t<t0, 0<r1, and t=t0, 0<r<1 by the strong maximum principle. By (64), we have (65) for 0tt0, that is,

β=01ψrzrdr01ωψr2rdr<0,0tt0,E71

and hence

z=βωψr<0r=1,t=t0,E72

a contradiction. It holds that z=ωr<0 for 0t<T, r=1, and hence

β=01ψrωrrdr01ωψr2rdr<0,0t<T.E73

The proof of Theorem 3 relies on the fact

βC,Ωωlogω1CT=+,ωtC.E74

This property is known for the Smoluchoski-Poisson equation (34), but the proof is valid even to (31) with vortex term. Having (21), therefore, we have to provide the inequality βC.

The inequality β<0, on the other hand, is sufficient for the following arguments.

Lemma 2 If β0, 0t<T, it holds that

ωω¯minΩ¯ω0>0onΩ¯×0T.E75

Proof: Since (17) we obtain

ωt+ψω=Δω+βψω+βΔψ=Δω+βψωβω2Δω+βψωinΩ×0TE76

with

ων=βωψν>0on∂Ω×0TE77

by (8). Then the result follows from the comparison theorem.

Lemma 3 Under the assumption of the previous lemma, there is C0=C0Ω>0 such that

C0ω023Eω¯ωt2ω02,βtαω022Eω¯,0t<T.E78

Proof: Using (11) and (17), we obtain

Ωωψω=Ωωωψ=12Ωω2ψ=12Ωω2ψ=0.E79

Hence (1) with (2) implies

12ddtω22+ω22=βΩωψω=β2ψω2=β2∂Ωω2ψν+β2Δψω2β2ω33E80

by β0 and (88). Since

Ωωψ=∂Ωωψν+ΩωΔψΩω2E81

follows from (8), furthermore, it holds that

β=ΩωψΩωψ2ω¯1ω22ψ22=1Eω¯ω22.E82

Then ineqality (80) induces

12ddtω22+ω2212Eω¯ω22ω33.E83

Here we use the Gagliardo-Nirenberg inequality (see (4.16) of [19]) in the form of

ω33CωH1ω22=Cω22ω2+ω2,E84

to obtain

12ddtω22+ω22CEω¯ω24ω2+ω212ω22+C28Eω¯2ω28+C2Eω¯ω25E85

and hence

ddtω22+ω22CEω¯ω25CEω¯ω23+1.E86

Then, Poincaré-Wirtinger’s inequality ensures

ddtω22+μω22CEω¯CEω¯ω26+ω23ω22,E87

where μ=μΩ>0 is a constant.

Writing

yt=CEω¯ω23,E88

we obtain

ddtω22+μω22y2+yω22,E89

and therefore, if

y2+y<μ/2E90

holds at t=0, it keeps to hold that

ddtω220E91

and (90) for 0t<T. Then, we obtain

ωt2ω02,0t<T,E92

and hence

βtω022Eω¯=α,0t<TE93

by (82).

The condition y0<μ2 means

C0ω02Eω¯E94

for C0>0 sufficiently large, and hence we obtain the conclusion.

Proof of Theorem 1: By the parabolic regularity, it suffices to show that

ωtC,0t<TE95

under the assumption. We have readily shown

ωt2C,0βtC,0t<TE96

by Lemma 3. Then, the conclusion (95) is obtained similarly to (34). See [26] for more details.

In fact, we have

Ωωψωp=Ωωψωp=pΩωpψω=pp+1Ωψωp+1=pp+1Ωωp+1ψ=0E97

for p>0 by (11) and (34). Then it follows that

1p+1ddtΩωp+1+4pp+12ωp+1222=βΩωψωp=βpp+1Ωψωp+1βpp+1Ωωp+1Δψ=βpp+1Ωωp+1CΩωp+2E98

by β<0 and (8). Then, Moser’s iteration scheme ensures (95) as in [33].

Advertisement

4. Proof of Theorem 2

We begin with the following lemma.

Lemma 4 Under the assumption of (22), it holds that

βtδ,0t<T,M=ω01>8πδT<+E99

in (31), where δ>0 is a constant.

Proof: We have ω=ωrt and ψ=ψrt for r=x under the assumption, which implies ψ=0. Then we obtain

ωψ=ωψ=0E100

by (17). It holds also that

ωψ=ωψrxr=xrωψr+xrωψr=1rωψr+ωψrr=1rrωψrr,E101

and therefore, there arises that

ωt=1rrωr+βrωψrr,ωr+βωψrr=1=0.E102

from (31).

Then (102) implies

ddt01ωr3dr=01ωtr3dr=01rωr+βrωψrrr2dr=012r2ωr+βωψrdr=2r2ωr=0r=1+0142βωψrr2dr.E103

Here we use (50) derived from the Poisson part of (31), that is,

rψrrt=Art0rstds.E104

Putting

λ=01ωrdr=M2π,E105

we obtain

ddt01ωr3dr=2ωr=1+4λ+2β01AArdr=2ωr=1+4λ+βA2r=0r=1=2ωr=1+4λ+βλ2<4λβ+M8π4λδ+M8π.E106

Since δ+M8π<0, therefore, T=+ is impossible, and we obtain T<+.

Lemma 5 Under the assumption (22), there is δ>0 such that

EM2<δ,βt0,0t<Tβt1CE1/20t<T.E107

Proof: First, Lemma 1 implies

ωωωr=1.E108

Second, we have

Ωψω=∂Ωψνω+ΩΔψω=ω∂Ωψν+ω22=ωΩΔψ+ω22=ω22ωM,E109

and hence

β=ΩψωΩωψ2=ω22ωMΩωψ2.E110

Here, we use the Gagliardo-Nirenberg inequality in the form of

w42Cw2wH1,E111

which implies

Ωωψ2ω2ψ42Cω2ψ2ψH1CE1/2ω22E112

by the elliptic estimate of the Poisson equation in (2),

ψH2Cω2.E113

We have, on the other hand,

ωMMEΩωψ2E114

by (110), and therefore,

β1CE1/2EM12CE1/2,E115

provided that

EM2<12C2.E116

Then the conclusion follows.

Proof of Theorem 2: By Lemma 5, there is δ0> such that

EM2<δβ1CE1/2δ1,E117

and then, Lemma 4 ensures

M>8πδ1T<+.E118

The assumption in (118) means

EM2<18πc2,E119

and hence we obtain the conclusion.

Advertisement

Acknowledgments

This work was supported by JSPS Grand-in-Aid for Scientific Research 19H01799.

Advertisement

This theorem is valid to the general case of Ω and ω0 without (22). We assume δ=1 without loss of generation, so that

β1.E120

We follow the argument [27] concerning (34) with the Poisson part replaced by (42) or (43). Thus we have to take case of the vortex term ωψ, time varying β=βt, and the Dirichlet boundary condition in (31).

We recall the cut-off function used in [34] (see also Chapter 5 of [19]). Hence each x0Ω¯ and 0<R1 admit φ=φx0,RC2Ω¯ with

φν∂Ω=0,0φ1,φ=1inΩBx0R/2,φ=0inΩ\Bx0R,E121

and

φCR1φ1/2,2φCR2φ1/2.E122

In more details, we take a cut-off function, denoted by ψ, satisfying (121), using a local conformal mapping, and then put φ=ψ4.

Let

φC2Ω¯,φν∂Ω=0.E123

be given. First, we have

ddtΩωφ=Ωωψφω+βωψφdx=Ωωψφ+ωΔφβωψφdxE124

by (11). It holds that

Ωωψφ=Ω×ΩωxtxGxxφxωxtdxdx=ΩΩωxtφx0,2RxxGxxφxωxtdxdx+Ω×Ωωxt1φx0,2RxxGxxφxωxtdxdx=I+II.E125

Let, furthermore, x0Ω and 0<R1 in the above equality. Then,

φ=xx02φx0,.RE126

satisfies the requirement (123).

It holds that

φ=2xx0φx0,R+xx02φx0,RE127

and hence

φCxx0φx0,R+xx0R1φx0,R1/2Cxx0φx0,R1/2.E128

We obtain, furthermore,

xx02R,xx0RxxR,E129

and hence

xGxxCR1E130

in this case. Then it follows that

IICR1MΩxx0φx0,R1/2ωxtdxCR1M3/2A1/2,E131

where

A=Ωxx02φx0,Rω.E132

We have, on the other hand,

I=Ω×Ωωxtφx0,2RxxGxxφxωxtdxdx=12ΩΩφx0,2RxφxxGxx+φx0,2RxφxxGxxωω,E133

where G=Gxx is the Green’s function to

Δψ=ω,ω∂Ω=0E134

and

ωω=ωxtωxtdxdx.E135

Here we use the local property of the Green’s function

Gxx=Γxx+Kxx,KC2Ω¯×ΩC2Ω×Ω¯,E136

where

Γx=12πlog1xE137

stands for the fundamental solution to Δ.

Let

ρx0,R2(x,x=φx0,2RxφxxKxx+φx0,2RφxxKx.x.E138

Since (128) implies

φx0,2RxφxCφx0,2Rxxx0φx0,R1/2xCxx0φx0,R1/2x,E139

it holds that

ρx0,R1xxCxx0φx0,R1/2x+xx0φx0,R1/2x.E140

Then, we obtain

I=12Ω×Ωρx0,R0xxωω+IIIE141

with

IIICM3/2A1/2CR1M3/2A1/2,E142

where

ρx0,R0xx=Γxxφx0,2Rxφxφx0,2Rxφx.E143

Here, we have

Γx=x2πx2,E144

and therefore,

ρx0,R0xx=ρx0,R2xx+ρx0,R3xxE145

fo

ρx0,R2xx=12πxxxx2φx0,2RxφxφxE146
ρx0,R3xx=12πxxxx2φx0,2Rxφx0,2Rxφx.E147

Since (128) implies

ρx0,R3xxCR1φxCR1xx0φx0,R1/2x,E148

there arises that

I=12Ω×Ωρx0,R2xxωω+IV,E149

with

IVCR1M3/2A1/2,E150

similarly.

We have, furthermore,

φxφx=2xxφx0,Rx+2xx0φx0,Rxφx0,Rx+xx02φx0,Rxφx0,Rx+xx02xx02φx0,Rx,E151

and hence

ρx0,R2xx=1πφx0,2Rxφx0,Rx+ρx0,R4xx+ρx0,R5xx+ρx0,R6xxE152

with

ρx0,R4xxCxx1φx0,2Rxxx0φx0,Rxφx0,RxCR1xx0φx0,2Rx,E153
ρx0,R5xxCxx1φx0,2Rxxx02φx0,Rxφx0,RxCR2xx02φx0,2RxCR1xx0φx0,2Rx,E154

and

ρx0,R6xxCxxφx0,2Rxxx02xx02φx0,RxCR1xx0+xx0φx0,Rxφx0,2RxCR1xx0φx0,Rx+R1xx0φx0,2RxE155

by

xx02xx02=xxx+x2x0xxxx0+xx0.E156

The residual terms are thus treated similarly, and it follows that

I+12πΩωφx0,RΩωφx0,2RCR1M3/2A1/2,E157

which results in

Ωωψφ+12πΩωφx0,RΩωφx0,2RCR1M3/2A1/2.E158

We can argue similarly to the vortex term in (124). This time, from

Γxx=0E159

it follows that

ΩωψφCR1M3/2A1/2.E160

Concerning the principal term of (124), we use

Δφ=4φx0,R+4xx0φx0,R+xx02Δφx0,R.E161

From

xx0φx0,RCR1xx0φx0,R1/2E162

and

x02Δφx0,RCR2xx02φx0,R1/2CR1xx0φx0,R1/2,E163

it follows that

ΩωΔφ4Ωωφx0,RCΩR1xx0φx0,RωCR1M1/2A1/2.E164

Let M1=Mx0,R and M2=Mx0,2R for

Mx0,R=Ωωφx0,R.E165

Then, using (120), we end up with

dAdt4M1M122π+CR1M3/2+M1/2A1/2+CM2M1.E166

Inequalilty (166) implies T<+ if A01, as is observed by [27] (see also Chapter 5 of [19]). Here we describe the proof for completeness.

The first observation is the monotoniity formula

ddtΩωφCM+M2φC1,E167

derived from (124) and the symmetry of the Green’s function: Gxx=Gxx. The proof is the same as in (34) and is omitted.

Second, we put I1=Ix0,R and I2=Ix0,2R for

Ix0,R=Ωxx02ωφx0,R.E168

Then it holds that

M2M1R<xx0<2Rφx0,2Rω2R1Ωxx0φx0,2Rω2M1/2R1I21/2E169

and

A2=A1+Ωxx02φx0,2Rφx0,RωA1+4R2Ωφx0,2Rφx0,Rω,E170

which implies

dA1dt4M1M122π+CR1M3/2+M1/2A11/2+CM3/2+M1/2Ωφx0,2Rφx0,Rω1/2.E171

Here, we use (167) to ensure

ddt(4M1M22πCM+M2R2E172

and

ddtΩφx0,2Rφx0,RωCM+M2R2.E173

Then, it follows that

4M1M122π4M10M1022π+CBaR1t1/2E174

and

Ωφx0,2Rφx0,RωΩφx0,2Rφx0,Rω0+CBaR1t1/22R2A20+CBaR1t1/2E175

for

B=M3/2+M1/2,as=s2+s.E176

Thus we obtain

dA1dt4M10M1022π+CR1BA11/2+CBA201/2+CBaR1t1/2=J0+CBaR1t1/2+CBR1A11/2E177

for

J=4M1M124π+CBR1A21/2.E178

Assume M10>8π, and put

4δ=4M10M1022π<0.E179

Let, furthemore,

1R2Ωxx02φx0,2Rω0η.E180

Now we define s0 by

CBas0=δE181

in (177), and take 0<η1 such that

ηδs02.E182

Then, if R and T0 satisfy R2T0=ηδ1, it holds that

A10R2η<2δT0.E183

Making 0<η1, furthermore, we may assume

J0+CBR1A101/24δ+CBR1A201/24δ+CBη1/23δ,E184

which results in

dA1dtJ0+CBaR1T01/2+BR1A1t1/2
=J0+δ+CBR1A11/2,0t<T0,E185

provided that TT0.

A continuation argument to (184)(185) guarantees

dA1dt2δ,0t<T0,E186

and then we obtain

A1T0A102δT0<0E187

by (183), a contradiction.

References

  1. 1. P.H. Chavanis, Generalized thermodynamics and Fokker-Planck equations: Applications to stellar dynamics and two-dimensional turbulence, Phys. Rev. E68, (2003) 036108.
  2. 2. P.-H. Chavanis, Two-dimensional Brownian vortices, Physica A 387 (2008) 6917-6942.
  3. 3. P.H. Chavanis, J. Sommeria, R. Robert, Statistical mechanics of two-dimensional vortices and collisionless stellar systems, Astrophys. J. 471 (1996) 385-399.
  4. 4. K. Sawada, T, Suzuki, Relaxation theory for point vortices, In; Vortex Structures in Fluid Dynamic Problems (H. Perez-de-Tejada ed.), INTECH 2017, Chapter 11, 205-224.
  5. 5. D. Lynden-Bell, Statistical mechanics of violent relaxation in stellar systems, Monthly Notices of Royal Astronmical Society 136 (1967) 101-121.
  6. 6. E. Caglioti, P.L. Lions, C. Marchioro, M. Pulvirenti, A special class of stationary flows for two-dimensional Euler equations: a statistical mechanics description Comm. Math. Phys. 143 (1992) 501.
  7. 7. G.L. Eyink, H. Spohn, Negative-temperature states and large scale, long-visited vortices in two-dimensional turbulence, Statistical Physics 70 (1993) 833.
  8. 8. G. Joyce, D. Montgomery, Negative temperature states for two-dimensional guiding-centre plasma, J. Plasma Phys. 10 (1973) 107.
  9. 9. M.K.H. Kiessling, Statistical mechanics of classical particles with logarithmic interaction, Comm. Pure Appl. Math. 46 (1993) 27-56.
  10. 10. K. Nagasaki, T. Suzuki, Asymptotic analysis for two-dimensional elliptic eigenvalue problems with exponentially dominated nonlinearities, Asymptotic Analysis 3 (1990) 173-188.
  11. 11. L. Onsager, Statistical hydrodynamics, Suppl. Nuovo Cimento 6 (1949) 279-287.
  12. 12. Y.B. Pointin, T.S. Lundgren, Statistical mechanics of two-dimensional vortices in a bounded container, Phys. Fluids 19 (1976) 1459-1470.
  13. 13. R. Robert, J. Sommeria, Statistical equilibrium states for two-dimensional flows, J. Fluid Mech. 229 (1991) 291-310.
  14. 14. R. Robert, J. Sommeria, Relaxation towards a statistical equilibrium state in two-dimensional perfect fluid dynamics, Phys. Rev. Lett. 69 (1992) 2776-2779.
  15. 15. R. Robert, A maximum-entropy principle for two-dimensional perfect fluid dynamics J. Stat. Phys. 65 (1991) 531-553.
  16. 16. R. Robert, C. Rosier, The modeling of small scales in two-dimensional turbulent flows: A statistical Mechanics Approach, J. Stat. Phys. 86 (1997) 481-515.
  17. 17. K. Sawada, T. Suzuki, Derivation of the equilibrium mean field equations of point vortex system and vortex filament system, Theor. Appl. Mech. Japan 56 (2008) 285-290.
  18. 18. C. Sire, P.-H. Chavanis, Thermodynamics and collapse of self-gravitating Brownian particles in D dimensions, Phys. Rev. E 66 (2002) 046133.
  19. 19. T. Suzuki, Free Energy and Self-Interacting Particles, Birkhäuser, Boston, 2005.
  20. 20. T. Suzuki, Mean Field Theories and Dual Variation - Mathematical Structures of Mesoscopic Model, 2nd edition, Atlantis Press, Paris, 2015.
  21. 21. T. Suzuki, Chemotaxis, Reaction, Network - Mathematics for Self-Organization, World Scientific, Singapore, 2018.
  22. 22. T. Suzuki, Liouville’s Theory in Linear and Nonlinear Partial Differential Equations - Interaction of Analysis, Geometry, Physics, Springer, Berlin, (to appear).
  23. 23. T. Suzuki, Applied Analysis - Mathematics for Science, Engineering, Technology, 3rd edition, Imperial College Press, London, (to appear).
  24. 24. P. Biler, Local and global solvability of some parabolic systems modelling chemotaxis, Adv. Math. Sci. Appl. 8 (1998) 715-743.
  25. 25. H. Gajewski, K. Zacharias, Global behaviour of a reaction-diffusion system modelling chemotaxis, Math. Nachr. 195 (1998) 77-114.
  26. 26. T. Nagai, T. Senba, and K. Yoshida, Application of the Trudinger-Moser inequality to a parabolic system of chemotaxis, Funkcial. Ekvac. 40 (1997) 411-433.
  27. 27. T. Senba and T. Suzuki, Parabolic system of chemotaxis; blowup in a finite and in the infinite time, Meth. Appl. Anal. 8 (2001) 349-368.
  28. 28. W. Jäger, S. Luckhaus, On explosions of solutions to a system of partial differential equations modelling chemotaxis, Trans. Amer. Math. Soc. 329 (1992) 819-824.
  29. 29. T. Nagai, Blow-up of radially symmetric solutions to a chemotaxis system, Adv. Math. Sci. Appl. 5 (1995) 581-601.
  30. 30. T. Suzuki, Exclusion of boundary blowup for 2D chemotaxis system provided with Dirichlet condition for the Poisson part, J. Math. Pure Appl. 100 (2013) 347-367.
  31. 31. T. Suzuki, Semilinear Elliptic Equations - Classical and Modern Theories, De Gruyter, Berlin, 2020.
  32. 32. B. Gidas, W.M. Ni, L. Nirenberg, Symmetry and related properties via the maximum principle, Comm. Math. Phys. 68 (1979) 209-243.
  33. 33. N.D. Alikakos, Lp bounds of solutions of reaction-diffusion equations, Comm. Partial Differential Equations 4 (1979) 827-868.
  34. 34. T. Senba and T. Suzuki, Chemotactic collapse in a parabolic-elliptic system of mathematical biology, Adv. Differential Equations 6 (2001) 21-50.

Written By

Ken Sawada and Takashi Suzuki

Reviewed: 22 September 2021 Published: 09 November 2021