• Aucun résultat trouvé

Long-time dynamics of 2d double-diffusive convection: analysis and/of numerics

N/A
N/A
Protected

Academic year: 2022

Partager "Long-time dynamics of 2d double-diffusive convection: analysis and/of numerics"

Copied!
26
0
0

Texte intégral

(1)

Revised

Proof

Numer. Math.

DOI 10.1007/s00211-014-0670-9

Numerische Mathematik

Long-time dynamics of 2d double-diffusive convection:

analysis and/of numerics

Florentina Tone · Xiaoming Wang · Djoko Wirosoetisno

Received: 5 March 2014 / Revised: 23 September 2014

© Springer-Verlag Berlin Heidelberg 2014

Abstract We consider a two-dimensional model of double-diffusive convection and

1

its time discretisation using a second-order scheme (based on backward differentiation

2

formula for the time derivative) which treats the non-linear term explicitly. Uniform

3

bounds on the solutions of both the continuous and discrete models are derived (under

4

a timestep restriction for the discrete model), proving the existence of attractors and

5

invariant measures supported on them. As a consequence, the convergence of the

6

attractors and long time statistical properties of the discrete model to those of the

7

continuous one in the limit of vanishing timestep can be obtained following established

8

methods.

9

Mathematics Subject Classification 65M12·35B35·35K45

10

1 Introduction

11

The phenomenon of double-diffusive convection, in which two properties of a fluid

12

are transported by the same velocity field but diffused at different rates, often occurs

13

F. Tone

Department of Mathematics and Statistics,

University of West Florida, Pensacola, FL 32514, USA e-mail: ftone@uwf.edu

X. Wang

Department of Mathematics, Florida State University, Tallahassee, FL 32306-4510, USA

e-mail: wxm@math.fsu.edu D. Wirosoetisno (

B

)

Mathematical Sciences, Durham University, Durham DH1 3LE, UK e-mail: djoko.wirosoetisno@durham.ac.uk

(2)

Revised

Proof

in nature [12]. Perhaps the best known example is the transport throughout the world’s

14

oceans of heat and salinity, which has been recognised as an essential part of cli-

15

mate dynamics [15,21]. In contrast to simple convections (cf. [3]), double-diffusive

16

convections support a richer set of physical regimes, e.g., a stably stratified initial

17

state rendered unstable by diffusive effects. Although in this paper we shall be refer-

18

ring to the oceanographic case, the mathematical theory is essentially identical for

19

astrophysical [14,16] and industrial [4] applications.

20

In this paper, we consider a two-dimensional double-diffusive convection model,

21

which by now-standard techniques [18] can be proved to have a global attractor and

22

invariant measures supported on it, and its temporal discretisation. We use a backward

23

differentiation formula for the time derivative and a fully explicit method for the non-

24

linearities, resulting in an accurate and efficient numerical scheme. Of central interest,

25

here and in many practical applications, is the ability of the discretised model to cap-

26

ture long-time behaviours of the underlying PDE. This motivates the main aim of this

27

article: to obtain bounds necessary for the convergence of the attractor and associated

28

invariant measures of the discretised system to those of the continuous system. We do

29

this using the framework laid down in [19,20], with necessary modifications for our

30

more complex model.

31

For motivational concreteness, one could think of our system as a model for the

32

zonally-averaged thermohaline circulation in the world’s oceans. Here the physical

33

axes correspond to latitude and altitude, and the fluid is sea water whose internal

34

motion is largely driven by density differentials generated by the temperature T and

35

salinity S, as well as by direct wind forcing on the surface. Both T and S are also

36

driven from the boundary—by precipitation/evaporation and ice melting/formation

37

for the salinity, and by the associated latent heat release and direct heating/cooling

38

for the temperature. Physically, one expects the boundary forcing for T , S and the

39

momentum to have zonal (latitude-dependent) structure, so we include these in our

40

model. Furthermore, one may also wish to impose a quasi-periodic time dependence

41

on the forcing; although this is eminently possible, we do not do so in this paper to

42

avoid technicalities arising from time-dependent attractors.

43

Taking as our domainD = [0,L] × [0,H]which is periodic in the horizontal

44

direction, we consider a temperature field Tand a salinity field S, both transported

45

by a velocity fieldv=(u, w)which is incompressible,∇·v =0, and diffused

46

at ratesκT andκS, respectively,

47

∂T/∂t+v·∇T=κTT

∂S/∂t+v·∇S=κSS. (1.1)

48

Here the stardenotes dimensional variables. Taking the Boussinesq approximation

49

and assuming that the density is a linear function of T and S, which is a good

50

approximation for sea water (although not for fresh water near its freezing point), the

51

velocity field evolves according to

52

∂v/∂t+v·∇v+ ∇p=κvv+TTαSS)ez (1.2)

53

for some positive constantsαT andαS.

54

(3)

Revised

Proof

Our system is driven from the boundary by the heat and salinity fluxes (which could

55

be seen to arise from direct contact with air and latent heat release in the case of heat,

56

and from precipitation, evaporation and ice formation/melt in the case of salinity),

57

∂T/∂n=QT and ∂S/∂n=QS. (1.3)

58

Here ndenotes the outward normal, n =zat the top boundary and n= −zat

59

the bottom boundary. We also prescribe a wind-stress forcing,

60

∂u/∂n=Qu (1.4)

61

along with the usual no-flux conditionw=0 on z=0 and z=H.

62

Largely following standard practice, we cast our system in non-dimensional form

63

as follows. Using the scalest ,˜ l,˜ T and˜ S, we define the non-dimensional variables˜

64

t =tt , x = x/l,˜ v = vt˜/l, T˜ = T/T and S˜ = S/S, in terms of which our˜

65

system reads

66

p1(∂tv+v·∇v)= −∇p+v+(TS)ez

tT+v·∇T =T

tS+v·∇S =βS.

(1.5)

67

To arrive at this, we have putl˜=Hand taken the thermal diffusive timescale for

68

˜

t= ˜l2T, (1.6)

69

as well as scaled the dependent variables as

70

T˜ =pl˜/(αTt˜2) and S˜=pl˜/(αSt˜2), (1.7)

71

where the non-dimensional Prandtl number and diffusivity ratio (also known as the

72

Lewis number in the engineering literature) are

73

p=κvT and β =κTS. (1.8)

74

Another non-dimensional quantity is the domain aspect ratioξ = L/l. The sur-˜

75

face fluxes are non-dimensionalised in the natural way: QT = pQT/(αTt˜2),

76

QS=pQS/(αSt˜2)and Qu=Qut.˜

77

For clarity and convenience, keeping in mind the oceanographic application, we

78

assume that the fluxes vanish on the bottom boundary z=0,

79

Qu(x,0)=QT(x,0)=QS(x,0)=0. (1.9)

80

(4)

Revised

Proof

For boundedness of the solution in time, the net fluxes must vanish, so (1.9) then

81

implies that the net fluxes vanish on the top boundary z=1,

82

ξ

0

Qu(x,1)dx= ξ

0

QT(x,1)dx = ξ

0

QS(x,1)dx =0. (1.10)

83

These boundary conditions can be seen to imply that the horizontal velocity flux is

84

constant in time, which we take to be zero, viz.,

85

1

0

u(x,z,t)dz= 1

0

u(x,z,0)dz≡0 for all x∈ [0, ξ]. (1.11)

86

For some applications (e.g., the classical Rayleigh–Bénard problem), the fluxes on the

87

bottom boundary may not vanish, which must then be balanced by the fluxes on the

88

top boundary,

89 ξ

0

[QT(x,1)−QT(x,0)]dx=0 (1.12)

90

and similarly for Qu and QS. With some modifications (by subtracting background

91

profiles from u, T and S), the analysis of this paper also apply to this more general

92

case. This involves minimal conceptual difficulty but adds to the clutter, so we do not

93

treat this explicitly here.

94

Defining the vorticityω := xwzu, the streamfunctionψ byψ = ωwith

95

ψ=0 on∂D[this is consistent with (1.11)], and the Jacobian determinant∂(f,g):=

96

xf∂zgxg∂zf = −∂(g, f), our system reads

97

p1{∂tω+∂(ψ, ω)} =ω+xTxS

tT +∂(ψ,T)=T

tS+∂(ψ,S)=βS.

(1.13)

98

The boundary conditions are,

99

zT =QT, ∂zS=QS, ω=Qu and ψ=0 on∂D. (1.14)

100

We note that for the solution to be smooth at t =0, the initial data and the boundary

101

conditions must satisfy a compatibility condition; cf. e.g., [17, Thm. 6.1] in the case

102

of Navier–Stokes equations. In the rest of this paper, we will be working with (1.13)–

103

(1.14) and its discretisation. We assume thatω, T and S all have zero integral overD

104

at t=0. Thanks to the no-net-flux condition (1.10), this persists for all t ≥0.

105

Another dimensionless parameter often considered in studies of (single-species)

106

convection is the Rayleigh number Ra. When the top and bottom temperatures are

107

held at fixed values T1and T0, Ra is proportional to T0T1. The relevant parameters

108

in our problem would be RaT ∝ |QT|L2(∂D)and RaS∝ |QS|L2(∂D), but we will not

109

consider them explicitly here; see, e.g., (2.11) in [2]. For notational conciseness, we

110

denote the variables U :=(ω,T,S), the boundary forcing Q:=(Qu,QT,QS)and

111

the parametersπ :=(p, β, ξ).

112

(5)

Revised

Proof

We do not provide details on the convergence of the global attractors and long time

113

statistical properties. Such kind of convergence can be obtained following established

114

methods once we have the uniform estimates derived here. See [10] for the convergence

115

of the global attractors and [20] for the convergence of long time statistical properties.

116

The rest of this paper is structured as follows. In Sect. 2 we review briefly the

117

properties of the continuous system, setting up the scene and the notation for its

118

discretisation. Next, we describe the time discrete system and derive uniform bounds

119

for the solution. In the appendix, we present an alternate derivation of the boundedness

120

results in [20], without using Wente-type estimates but requiring slightly more regular

121

initial data.

122

2 Properties of the continuous system

123

In this section, we obtain uniform bounds on the solution of our system and use them

124

to prove the existence of a global attractorA. For the single diffusion case (of T only,

125

without S), this problem has been treated in [6] which we follow in spirit, though not

126

in detail in order to be closer to our treatment of the discrete case.

127

We start by noting that the zero-integral conditions onω, T and S imply the Poincaré

128

inequalities

129

|ω|2L2(D)c0|∇ω|2L2(D), (2.1)

130

as well as the equivalence of the norms

131

|ω|H1(D)c|∇ω|L2(D), (2.2)

132

with analogous inequalities for T and S. The boundary conditionψ=0 implies that

133

(2.1)–(2.2) also hold forψ, while an elliptic regularity estimate [7, Cor. 8.7] implies

134

that

135

|∇ψ|2L2(D)c0|ω|2L2(D). (2.3)

136

Following the argument in [8], this also holds for functions, such as our T and S, with

137

zero integrals inD.

138

LetΩbe an H2extension of QutoD¯(further requirements will be imposed below)

139

and letωˆ := ωΩ; we also defineψˆ := ˆωandΨ :=Ω with homogeneous

140

boundary conditions. Nowωˆ satisfies the homogeneous boundary conditionsωˆ =0

141

on∂D, and thus the Poincaré inequality (2.1)–(2.2). Furthermore, let TQ ∈ ˙H2(D)

142

be such thatzTQ =QT on∂D(with other constraints to be imposed below) and let

143

Tˆ :=TTQ; analogously for SQ andSˆ :=SSQ. We note that since bothT andˆ

144

S have zero integrals overˆ D, they satisfy the Poincaré inequality (2.1)–(2.2).

145

We start with weak solutions of (1.13). For conciseness, unadorned norms and inner

146

products are understood to be L2(D),| · | := | · |L2(D)and(·,·):=(·,·)L2(D). With

147

ˆ

ω,T andˆ S as defined above, we haveˆ

148

(6)

Revised

Proof

tωˆ+∂(Ψ + ˆψ, Ω+ ˆω)=p

ωˆ+Ω+xTQ+xTˆ−xSQxSˆ

tTˆ+∂(Ψ + ˆψ,TQ+ ˆT)=TQ+Tˆ

tSˆ+∂(Ψ + ˆψ,SQ+ ˆS)=β (SQ+Sˆ).

(2.4)

149

On a fixed time interval[0,T), a weak solution of (2.4) are

150

ˆ

ωC0(0,T;L2(D))L2(0,T;H01(D)) Tˆ ∈C0(0,T;L2(D))L2(0,T;H1(D)) Sˆ∈C0(0,T;L2(D))L2(0,T;H1(D))

(2.5)

151

such that, for allω˜ ∈H01(D),T ,˜ S˜∈ H1(D), the following holds in the distributional

152

sense,

153

d

dt(ω,ˆ ω)˜ +(∂(Ψ + ˆψ, Ω+ ˆω),ω)˜

154

+p

(∇Ω+ ∇ ˆω,∇ ˜ω)(∂xTQ+xTˆ,ω)˜ +(∂xSQ+xS,ˆ ω)˜

=0

155

d

dt(Tˆ,T˜)+(∂(Ψ + ˆψ,TQ+ ˆT),T˜)+(∇ ˆT,∇ ˜T)(TQ,T˜)=0 (2.6)

156

d

dt(S,ˆ S)˜ +(∂(Ψ + ˆψ,SQ+ ˆS),S)˜ +β (∇ ˆS,∇ ˜S)β (SQ,S)˜ =0.

157 158

The existence of such solutions can be obtained by Galerkin approximation together

159

with Aubin–Lions compactness argument [17, §3.3], which we do not carry out explic-

160

itly here.

161

Next, we derive L2inequalities for T , S andω. Multiplying (2.4a) byωˆ in L2(D)

162

and noting that(∂(ψ,ω),ˆ ω)ˆ =0, we find

163

1 2 d

dt| ˆω|2+p|∇ ˆω|2 = −(∂(Ψ, Ω),ω)ˆ −(∂(ψ, Ω),ˆ ω)ˆ

164

+p

(Ω,ω)ˆ +(∂xT,ω)ˆ −(∂xS,ω)ˆ

. (2.7)

165

We bound the rhs as

166

(Ω,ω)ˆ = |∇Ω| |∇ ˆω| ≤1

8|∇ ˆω|2+2|∇Ω|2

167

(∂xT,ω)ˆ = |∂xω| |Tˆ | ≤ 1

8|∇ ˆω|2+2|T|2≤ 1

8|∇ ˆω|2+4c0|∇ ˆT|2+4|TQ|2

168

(∂xS,ω)ˆ = |∂xω| |S| ≤ˆ 1

8|∇ ˆω|2+2|S|2≤ 1

8|∇ ˆω|2+4c0|∇ ˆS|2+4|SQ|2,

169 170

(7)

Revised

Proof

and the “nonlinear” terms as (this defines c1)

171

(∂(ψ,ˆ ω), Ω)ˆ ≤c|∇ ˆψ|L|∇ ˆω|L2|Ω|L2c1

2 |Ω|L2|∇ ˆω|2

172

(∂(Ψ,ω), Ω)ˆ ≤c|∇Ψ|L|∇ ˆω|L2|Ω|L2 ≤ p

8|∇ ˆω|2+c

p|∇Ψ|2L|Ω|2.(2.8)

173

This brings us to

174

d

dt| ˆω|2+(pc1|Ω|)|∇ ˆω|2≤4pc0(|∇ ˆT|2+ |∇ ˆS|2)

175

+c

p|∇Ψ|2L|Ω|2+4p(|∇Ω|2+ |TQ|2+ |SQ|2).

176

(2.9)

177178

As usual, in the above and henceforth, c denotes generic constants which may take

179

different values each time it appears. Numbered constants such as c0have fixed values;

180

they are independent of the parameterspandβunless noted explicitly.

181

Now forS, we multiply (1.13c), or equivalently,ˆ

182

tSˆ+∂(ψ,Sˆ+SQ)=β (Sˆ+SQ), (2.10)

183

byS in Lˆ 2(D)and use(∂(ψ,S),ˆ S)ˆ =0 to find

184

1 2 d

dt| ˆS|2+β|∇ ˆS|2= −(∂(Ψ,SQ),S)ˆ −(∂(ψ,ˆ SQ),S)ˆ +β (SQ,S).ˆ (2.11)

185

The last term on the rhs requires some care,

186

(SQ,S)ˆ =(QS,S)ˆ L2(∂D)(∇SQ,∇ ˆS)

c|QS|H1/2(∂D)| ˆS|H1/2(∂D)+ |∇QS| |∇ ˆS|

≤ 1

8|∇ ˆS|2+c(QS2+ |∇SQ|2)

(2.12)

187

where we have used the trace theorem [1, Thm. 4.12] for the last inequality and

188

denoted QS := |QS|H1/2(∂D). We note that|∇SQ|L2(D) ultimately depends on

189

|QS|H1/2(∂D)plus the constraint (2.16) below. Bounding the “nonlinear” terms as

190

(∂(Ψ,S),ˆ SQ)c|∇Ψ|L|∇ ˆS|L2|SQ|L2β

8 |∇ ˆS|2+ c

β |∇Ψ|2L|SQ|2

191

(∂(ψ,ˆ S),ˆ SQ)c|∇ ˆψ|L|∇ ˆS|L2|SQ|L2β

8 |∇ ˆS|2+ c

β |∇ ˆω|2|SQ|2,

192 193

(8)

Revised

Proof

we arrive at (this defines c2)

194

d

dt| ˆS|2+β|∇ ˆS|2c2

8c0β |∇ ˆω|2|SQ|2

195

+c

β |∇Ψ|2L|SQ|2+cβ (|∇SQ|2+ QS2). (2.13)

196

Analogously, we have forT (withˆ QT := |QT|H1/2(∂D)),

197

d

dt| ˆT|2+ |∇ ˆT|2c2

8c0|∇ ˆω|2|TQ|2

198

+c|∇Ψ|2L|TQ|2+c(|∇TQ|2+ QT2). (2.14)

199

Adding 8pc0times (2.14) and 8pc0times (2.13) to (2.9), we find

200

d dt

| ˆω|2+8pc0| ˆT|2+8pc0 β | ˆS|2

+4pc0(|∇ ˆT|2+ |∇ ˆS|2)

201

+

p−c1|Ω| −c2p|TQ|2c2p β2|SQ|2

|∇ ˆω|2 (2.15)

202

cp|∇Ψ|2L

|Ω|2/p2+ |TQ|2+ |SQ|22

203

+cp(|∇Ω|2+ |∇TQ|2+ |∇SQ|2+ QT2+ QS2).

204205

If we now chooseΩ, TQ and SQsuch that

206

|Ω|L2 ≤p/(8c1), |TQ|2L2 ≤1/(8c2) and |SQ|2L2β2/(8c2), (2.16)

207

(given the BC (1.14), this can always be done at the price of making∇Ω,∇TQ and

208

SQ large) we obtain the differential inequality

209

d dt

| ˆω|2+8pc0| ˆT|2+8pc0 β | ˆS|2

+p

2|∇ ˆω|2+4pc0(|∇ ˆT|2+|∇ ˆS|2)≤ F2, (2.17)

210

with F2 denoting the purely “forcing” terms on the rhs of (2.15). Integrat-

211

ing this using the Gronwall lemma, we obtain the uniform bounds, with | ˆU|2 =

212

| ˆω|2+8pc0| ˆT|2+8pc0| ˆS|2,

213

| ˆU(t)|2≤e−λt| ˆU(0)|2+ F2

214

c3p t+1

t

|∇ ˆω|2+ |∇ ˆT|2+ |∇ ˆS|2

(t)dt≤e−λt|U(0)|2+(1+1/λ)F2

215

(2.18)

216217

valid for all t ≥ 0, for some λ(π) > 0. It is clear from (2.18a) that we have an

218

absorbing ball, i.e.|U(t)|2M0(Q;π)for all tt0(|U(0)|;π).

219

(9)

Revised

Proof

On to H1, we multiply (2.4a) by−ωˆ in L2to find

220

1 2 d

dt|∇ ˆω|2+p|ω|ˆ 2 = −(∂(∇ψ,ω),ˆ ∇ ˆω)+(∂(ψ, Ω), ω)ˆ

221

−p(Ω, ω)ˆ −p(∂xT, ω)ˆ +p(∂xS, ω).ˆ (2.19)

222

Bounding the linear terms in the obvious way, and the nonlinear terms as

223

(∂(∇ψ,ω),ˆ ∇ ˆω)c|∇ ˆω|2L4|∇2ψ|L2c|∇ ˆω| |ω| |ψ|ˆ

224

≤ p

8|ω|ˆ 2+c

p|∇ ˆω|2(| ˆω|2+ |Ω|2)

225

(∂(ψ, Ω), ω)ˆ ≤ p

8|ω|ˆ 2+c p|∇Ω|2

|∇ ˆω|2+ |∇Ψ|2L ,

226 227

we find

228

d

dt|∇ ˆω|2+p|ω|ˆ 2c

p|∇ ˆω|2(| ˆω|2+ |Ω|2+ |∇Ω|2)+c

p|∇Ψ|2L|∇Ω|2

229

+8p

|∇ ˆT|2+ |∇ ˆS|2+ |∇TQ|2+ |∇SQ|2+ |Ω|2 . (2.20)

230231

Sinceω,ˆ T andˆ S have been bounded uniformly in Lˆ 2t,1Hx1in (2.18b), we can integrate

232

(2.20) using the uniform Gronwall lemma to obtain a uniform bound for|∇ ˆω|2,

233

|∇ ˆω(t)|2M1(· · ·) and

t+1

t

|ω(tˆ )|2dt≤ ˜M1(· · ·). (2.21)

234

Similarly, multiplying (2.10) by−S in Lˆ 2, we find

235

1 2 d

dt|∇ ˆS|2+β|Sˆ|2= −β (SQ, Sˆ)

236

−(∂(∇ψ,Sˆ),∇ ˆS)+(∂(ψ,SQ), Sˆ). (2.22)

237

Bounding as we did forω, we arrive atˆ

238

d

dt|∇ ˆS|2+β|Sˆ|2≤8β|SQ|2

239

+ c

β |∇ ˆS|2(| ˆω|2+ |Ω|2)+ c

β |∇SQ|2

|∇ ˆω|2+ |∇Ψ|2L ,

240

(2.23)

241242

which can be integrated using the uniform Gronwall lemma to obtain

243

|∇ ˆS(t)|2M1(· · ·) and

t+1

t |Sˆ(t)|2dt≤ ˜M1(· · ·). (2.24)

244

(10)

Revised

Proof

Obviously one has the analogous bound forT ,ˆ

245

|∇ ˆT(t)|2M1(· · ·) and

t+1 t

|Tˆ(t)|2dt≤ ˜M1(· · ·). (2.25)

246

These bounds allow us to conclude [18] the existence of a global attractorAand

247

of an invariant measureμsupported onA. The convergence of the global attractors

248

can be deduced following an argument similar to that in [11], while the convergence

249

of the the invariant measures can be inferred from an argument similar to that in [20].

250

In particular, any generalised long-time average generates an invariant measure in the

251

sense that for any given bounded continuous functionalΦ(whose domain is the phase

252

space H and rangeR), we have

253

tlim→∞

1 t

t

0

Φ(S(t)U0)dt=

H

Φ(U)dμ(U) (2.26)

254

where U(t)=S(t)U0is the solution of (1.13) with initial data U0. It is known thatAis

255

unique whileμmay depend on the initial data U0and the definition of the generalised

256

limitlim.

257

Due to the boundary conditions, one cannot simply multiply by2ω, etc., to obtainˆ

258

a bound in H2, but following [17, §6.2], one takes time derivative of (1.13a) and uses

259

the resulting bound on|∂tω|to bound|ω|, etc. We shall not do this explicitly here,

260

although similar ideas are used for the discrete case below (Proof of Theorem2).

261

3 Numerical scheme: boundedness

262

Fixing a timestep k > 0, we discretise the system (1.13) in time by the following

263

two-step explicit–implicit scheme,

264

n+1−4ωn+ωn1

2k +∂(2ψnψn1,nωn1)

=p

ωn+1+xTn+1xSn+1 3Tn+14Tn+Tn1

2k +∂(2ψnψn1,2TnTn1)=Tn+1 3Sn+14Sn+Sn1

2k +∂(2ψnψn1,2SnSn1)=βSn+1,

(3.1)

265

plus the boundary conditions (1.14). Writing Un = n,Tn,Sn), we assume that

266

the second initial data U1has been obtained from U0using some reasonable one-step

267

method, but all we shall need for what follows is that U1H1(D). The time derivative

268

term is that of the backward differentiation formula (BDF2) and the explicit form of

269

the nonlinear term is chosen to preserve the order of the scheme. This results in a

270

method that is essentially explicit yet second order in time, and as we shall see below,

271

preserves the important invariants of the continuous system.

272

(11)

Revised

Proof

Subject to some restrictions on the timestep k, we can obtain uniform bounds

273

and absorbing balls for the solution of the discrete system analogous to those of the

274

continuous system. Our first result is the following:

275

Theorem 1 With QH3/2(∂D), the scheme (3.1) defines a discrete dynamical

276

system in H1(D)×H1(D). Assuming U0, U1H1(D)and the timestep restriction

277

given in (3.20) below,

278

kk1

|U0|H1,|U1|H1; |Q|H1/2(∂D), π , (3.2)

279

the following bounds hold

280

|Un|2L2 ≤40 e−νnk/4

|U0|2L2 + |U1|2L2 +M0

|Q|H1/2(∂D);π

281

+c

|Q|H1/2(∂D);π k e−νnk/4

|U0|2H1+ |U1|2H1 , (3.3)

282

|Un|2H1N1

nk; |U0|H1,|U1|H1,|Q|H1/2(∂D), π +M1(|Q|H3/2(∂D);π), (3.4)

283 284

whereν(π) >0 and N1(t; · · ·)=0 for tt1

|U0|H1,|U1|H1;Q, π .

285

We note that the last term in (3.3) has no analogue in the continuous case; we believe

286

this is an artefact of our proof, but have not been able to circumvent it. Here one can

287

choose the bounds M0 and M1 to hold for both the continuous and discrete cases,

288

although the optimal bounds (likely very laborious to compute) may be different.

289

Unlike in [20], H2bounds do not follow as readily due to the boundary conditions,

290

so we proceed by first deriving bounds for|Un+1Un|, using an approach inspired

291

by [17, §6.2]. We state our result without the transient terms:

292

Theorem 2 Assume the hypotheses of Theorem 1. Then for sufficiently large time,

293

nkt2(U0,U1;Q, π), one has

294

n+1ωn|2+ |Tn+1Tn|2+ |Sn+1Sn|2k2Mδ(|Q|H3/2(∂D);π). (3.5)

295

Furthermore, for large time nkt2the solution is bounded in H2as

296

n|2+ |Tn|2+ |Sn|2M2(|Q|H3/2(∂D);π). (3.6)

297

We remark that these difference and H2bounds require no additional hypotheses

298

on Q, suggesting that Theorem 1may be sub-optimal. We also note that using the

299

same method (and one more derivative on Q) one could bound|Un+1Un|H1 and

300

|Un|H3, although we will not need these results here.

301

Following the approach of [20], these uniform bounds (along with the uniform

302

convergence results that follow from them) then give us the convergence of long-time

303

statistical properties of the discrete dynamical system (3.1) to those of the continuous

304

system (1.13).

305

(12)

Revised

Proof

Proof of Theorem1Central to our approach is the idea of G-stability for multistep

306

methods [9, §V.6]. First, for f , g∈L2(D)andνk∈ [0,1], we define the norm

307

|[f,g]|2νk= |f|2L2

2 +5+νk

2 |g|2L2−2(f,g)L2. (3.7)

308

Note that our notation is slightly different from that in [11,20]. Since both eigenvalues

309

of the quadratic form are finite and positive for allνk∈ [0,1], this norm is equivalent to

310

the L2norm, i.e. there exist positive constants c+and c, independent ofνk∈ [0,1],

311

such that

312

c

|f|2L2 + |g|2L2 ≤ |[f,g]|2νkc+

|f|2L2+ |g|2L2 (3.8)

313

for all f , gL2(D); computing explicitly, we find

314

c= 6−√ 32

4 and c+=7+√

41

4 . (3.9)

315

As in [20], an important tool for our estimates is an identity first introduced in [9]

316

forνk=0; the following form can be found in [11, proof of Lemma 6.1]: for f , g,

317

hL2(D)andνk∈ [0,1],

318

(3h4g+ f,h)L2+νk|h|2L2

= |[g,h]|2νk− 1

1+νk|[f,g]|2νk+|f2g+(1+νk)h|2L2

2(1+νk) . (3.10)

319

The fact that (3.1) forms a discrete dynamical system in H1×H1can be seen by

320

writing

321

(3−2k)Tn+1=4TnTn12k∂(2ψnψn1,2TnTn1) (3.11)

322

and inverting: given Un1and UnH1(D), the Jacobian is in H1, which, with

323

the Neumann BCzTn+1=QTH1/2(∂D), gives Tn+1H1. Similarly for Sn+1

324

and, since now Tn+1, Sn+1H1andωn+1=QuH1/2(∂D), forωn+1. Therefore

325

(Un1,Un)H1×H1maps to(Un,Un+1)H1×H1.

326

Let ωˆn := ωnΩ, Tˆn := TnTQ and Sˆn := SnSQ be defined as in

327

the continuous case, i.e. Ω, TQ, SQH2(D)satisfying the boundary conditions

328

Ω =Qu,zTQ =QT andzSQ =QS, and the constraint (3.29), which is essentially

329

(2.16). The scheme (3.1) then implies

330

(13)

Revised

Proof

3ωˆn+1−4ωˆn+ ˆωn1

2k +∂(2ψnψn1,2ωˆn− ˆωn1+Ω)

=p

ωˆn+1+Ω+xTn+1xSn+1 3Tˆn+1−4Tˆn+ ˆTn1

2k +∂(2ψnψn1,2Tˆn− ˆTn1+TQ)=Tˆn+1+TQ

3Sˆn+1−4Sˆn+ ˆSn1

2k +∂(2ψnψn1,2Sˆn− ˆSn1+SQ)=β(Sˆn+1+SQ) (3.12)

331

where we have kept some ψn, Tn and Sn for now. We start by deriving difference

332

inequalities for ωˆn,Tˆn and Sˆn. In order to bound terms of the form |∇ ˆψn|2L

333

c| ˆωn|2H1/2, we assume for now the uniform bound

334

| ˆωn|2H1/2k1/2Mω(· · ·) for all n=0,1,2, . . . (3.13)

335

where Mωwill be fixed in (3.31) below. We also assume for clarity that k≤1.

336

Multiplying (3.12a) by 2kωˆn+1in L2(D)and using (3.10), we find

337

|[ ˆωnˆn+1]|2νkνk| ˆωn+1|2+2pk|∇ ˆωn+1|2+|(1+νk)ωˆn+1−2ωˆn+ ˆωn1|2 2(1+νk)

338

=|[ ˆωn1ˆn]|2νk

1+νk −2k(∂(2ψnψn1ˆn+1), (1+νk)ωˆn+1−2ωˆn+ ˆωn1)

339

+2k(∂(2ψˆn− ˆψn1ˆn+1), Ω)+2k(∂(Ψ,ωˆn+1), Ω)

340

+2pk

(Ω,ωˆn+1)+ˆn+1, ∂xTn+1)ˆn+1, ∂xSn+1)

. (3.14)

341

whereν >0 will be set below. We bound the last terms as in the continuous case,

342

2|(Ω,ωˆn+1)≤ 1

8|∇ ˆωn+1|2+8|∇Ω|2

343

2|(∂xTn+1ˆn+1)| ≤ 1

8|∇ ˆωn+1|2+16c0|∇ ˆTn+1|2+16|TQ|2

344

2|(∂xSn+1ˆn+1)| ≤ 1

8|∇ ˆωn+1|2+16c0|∇ ˆSn+1|2+16|SQ|2

345

2(∂(Ψ,ωˆn+1), Ω)≤ p

8|∇ ˆωn+1|2+c

p|∇Ψ|2L|Ω|2,

346 347

and the previous one as

348

2|(∂(2ψˆn− ˆψn1ˆn+1), Ω)| ≤c|2∇ ˆψn− ∇ ˆψn1|L|∇ ˆωn+1|L2|Ω|L2 349

≤ p

8|∇ ˆωn+1|2+c

p(|∇ ˆωn1|2+ |∇ ˆωn|2)|Ω|2. (3.15)

350

351

Références

Documents relatifs

Using a theory based on Hamilton-Jacobi equations with constraint, we prove that, as the effect of the mutations vanishes, the solu- tion concentrates on a single Dirac mass, while

In Section 4, we prove first- and second-order necessary optimality conditions for the transformed (P), which are then translated into optimality conditions for the time

Our approach is based on duality arguments (see Proposition 2.7 which is adapted from a result due to Krein-Rutman [14]) and allows us to obtain the existence of eigenvectors in

This last result generalizes the exponential dissipation of the entropy along the heat flow [ BE85 ] and, in view of the stochastic control formulation of the Schr¨ odinger problem,

Such evidence suggests the existence of a replacement time value, t * , that represents the value under which the substitution of the plant modules is not profitable from an

Les travaux présentés dans ce chapitre nous ont d’abord permis d’établir la relation abstraite entre les Frappes de Processus et les approches de Thomas pour une

Theorem 1 below shows that with increasingly smooth functions ϕ satisfying appropriate zero boundary conditions, it is possible to improve the rate of convergence to 1/T k for

In this paper, we will investigate the propagation of a coherent quantum wave packet under the action of strong magnetic fields, for times of order t 1/ ~.. How is