• Aucun résultat trouvé

Effective temperatures of a heated Brownian particle

N/A
N/A
Protected

Academic year: 2021

Partager "Effective temperatures of a heated Brownian particle"

Copied!
6
0
0

Texte intégral

(1)

HAL Id: hal-00540388

https://hal.archives-ouvertes.fr/hal-00540388

Submitted on 26 Nov 2010

HAL is a multi-disciplinary open access

archive for the deposit and dissemination of

sci-entific research documents, whether they are

pub-lished or not. The documents may come from

teaching and research institutions in France or

abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est

destinée au dépôt et à la diffusion de documents

scientifiques de niveau recherche, publiés ou non,

émanant des établissements d’enseignement et de

recherche français ou étrangers, des laboratoires

publics ou privés.

Effective temperatures of a heated Brownian particle

Laurent Joly, Samy Merabia, Jean-Louis Barrat

To cite this version:

Laurent Joly, Samy Merabia, Jean-Louis Barrat. Effective temperatures of a heated Brownian particle.

EPL - Europhysics Letters, European Physical Society/EDP Sciences/Società Italiana di Fisica/IOP

Publishing, 2011, 94, pp.50007. �10.1209/0295-5075/94/50007�. �hal-00540388�

(2)

Laurent Joly,∗ Samy Merabia,and Jean-Louis Barrat

2

LPMCN, Universit´e de Lyon; UMR 5586 Universit´e Lyon 1 et CNRS, F-69622 Villeurbanne, France 3

(Dated: November 26, 2010) 4

We investigate various possible definitions of an effective temperature for a particularly simple 5

nonequilibrium stationary system, namely a heated Brownian particle suspended in a fluid. The 6

effective temperature based on the fluctuation dissipation ratio depends on the time scale under 7

consideration, so that a simple Langevin description of the heated particle is impossible. The short 8

and long time limits of this effective temperature are shown to be consistent with the temperatures 9

estimated from the kinetic energy and Einstein relation, respectively. The fluctuation theorem 10

provides still another definition of the temperature, which is shown to coincide with the short time 11

value of the fluctuation dissipation ratio. 12

PACS numbers: 05.70.Ln, 05.40.-a, 82.70.Dd, 47.11.Mn

13

In the recent years, so called “active colloids”, i.e.

col-14

loidal particles that exchange with their surroundings in

15

a non Brownian manner, have attracted considerable

at-16

tention from the statistical physics community [1]. These

17

systems are of interest as possible models for simple

liv-18

ing organisms, and the description of the corresponding

19

nonequilibrium states using the tools of standard

statis-20

tical physics raises a number of fundamental questions

21

[2, 3]. The most widely studied active colloids are those

22

that exchange momentum with the supporting solvent in

23

a non stochastic way, resulting into self propulsion. A less

24

studied possibility is that the colloid acts as a local heat

25

source and is constantly surrounded by a temperature

26

gradient. Experimentally [4], such a situation is achieved

27

when colloids are selectively heated by an external source

28

of radiation which is not absorbed by the solvent. If the

29

heat is removed far away from the particle, or, more

prac-30

tically, if the particle concentration is small enough that

31

the suspending fluid can be considered as a thermostat,

32

a simple nonequilibrium steady state is achieved. Each

33

colloidal particle is surrounded by a spherically

symmet-34

ric halo of hot fluid, and diffuses in an a priori Brownian

35

manner. The diffusion constant of such heated Brownian

36

particles was experimentally shown to be increased

com-37

pared to the one observed at equilibrium [4], and a semi

38

quantitative analysis of this enhancement was presented

39

in reference [5], based on an analysis of the temperature

40

dependence of the viscosity.

41

In this report, we use simulation to investigate in

de-42

tail the statistical physics of the simple non equilibrium

43

steady state (NESS) formed by a heated particle

sus-44

pended in a fluid. The most natural way of describing

45

such a system, in which the particles diffuse isotropically

46

in the surrounding fluid, is to make use of a Langevin

47

type equation for the center of mass velocity U ,

involv-48

ing in general a memory kernel ζ(t) and a random force

49 R(t): 50 MdU dt = − Z t −∞ ζ(t − s)U (s) ds + Fext+ R(t). (1)

In a system at thermal equilibrium at temperature T , the

51

correlations in the random force and the friction kernel

52

are related by the standard fluctuation dissipation

the-53

orem, hRα(t)Rβ(t′)i = δαβζ(|t − t′|)kBT [6]. Obviously

54

such a description is not expected to hold for a heated

55

particle, as the system is now out of equilibrium. A

gen-56

eralization of Eq. 1, involving a corrected fluctuation

57

dissipation relation with an effective temperature Teff

re-58

placing the equilibrium one, would however appear as a

59

natural hypothesis. In fact, such an approach was shown

60

to hold for sheared systems kept at a constant

tempera-61

ture by a uniform thermostat [7], or in the frame of the

62

particle for a particle driven at constant average speed [8].

63

The interpretation of recent experiments [3] also makes

64

implicitly use of such a description in describing the

sedi-65

mentation equilibrium of active particles, or in analyzing

66

the diffusion constant for hot Brownian motion [5].

67

The use of a Langevin equation with an effective

tem-68

perature has several direct consequences. The kinetic

69

energy associated with the center of mass, h12M U2i, is

70

necessarily equal to the effective temperature 32kBTeff.

71

The diffusion coefficient D and the mobility under the

72

influence of an external force µ = Ux/Fx are related by

73

an Einstein relation, D/µ = kBTeff [9]. More generally,

74

this relation can be seen as the steady state version of

75

the proportionality between the time dependent response

76

function to an external force, χ(t) = δUx(t)/δFx, and

77

the velocity autocorrelation in the nonequilibrium steady

78 state: 79 χ(t) = 1 kBTeff hUx(0)Ux(t)i. (2)

This relation was explored numerically for self propelled

80

particles in reference [2], and shown to be consistent with

81

the observed Einstein like relation. Independently of the

82

use of a specific Langevin model, this relation defines

83

an effective temperature trough a so called “fluctuation

84

dissipation ratio”. The applicability of an effective

tem-85

perature description is determined by the dependence of

86

this fluctuation dissipation ratio on time. We show in

87

the following that the time scale at which the fluctuation

(3)

2 5 10 15 r [σ] 1 2 3 T [ ε /kB ] 5 10 15 r [σ] 0 1 ρ [ σ -3 ] Tc

T

p

T

0

FIG. 1. Left– Snapshot of the simulated system for Tp =

3.5ε/kB(T0= 0.75ε/kB); Gray levels indicate the kinetic

en-ergy of atoms. Right– Steady radial temperature and density profiles for this system.

dissipation ratio of a heated particle is determined indeed

89

matters, so that a single temperature description, even

90

in such a seemingly simple system, is problematic.

91

Finally, the use of a Langevin description with an

ef-92

fective temperature entails the validity of several

“fluc-93

tuation relations” [10], which have been the object of

94

numerous recent experimental and numerical tests, both

95

in equilibrium and nonequilibrium systems. The study of

96

the fluctuation relation for the heated particle constitutes

97

the last part of this report.

98

Our work is based on a direct molecular simulation

99

(MD) approach of a crystalline nanoparticle diffusing

100

in a liquid. The simulation were carried out using the

101

LAMMPS package [11]. Details of the model can be

102

found in previous works [12, 13], where we used this

103

system to investigate heat transfer from nanoparticles.

104

The particle was made of 555 atoms with a fcc

struc-105

ture, tied together using FENE bonds. The liquid was

106

made of ∼ 23000 atoms (Fig. 1). All liquid and solid

107 108

atoms interacted via the same Lennard-Jones (LJ)

po-109

tential v = 4ε[(σ/r)12− (σ/r)6], at the exclusion of solid

110

atoms directly bonded to each others. In the following,

111

all results will be given in LJ units, namely σ, ε/kB and

112

τ =pmσ2/ε for length, temperature and time,

respec-113

tively. The atoms in the solid particle were held at

con-114

stant temperature Tp using a Nos´e-Hoover thermostat,

115

after subtracting the velocity of the center of mass. In

116

order to mimic the bulk liquid – far from the particle

117

– acting as a thermal bath, a rescaling thermostat was

118

applied only to liquid atoms lying beyond 15σ from the

119

center of the particle (this condition being evaluated each

120

time the thermostat was applied), to keep them at

con-121

stant temperature T0 = 0.75ε/kB. This amounts to an

122

assumption that the temperature profile around the

par-123

ticle follows the latter instantaneously. This is a

reason-124

able assumption, as heat diffusion is much faster than

125

mass diffusion in our system: Dheat∼ 1σ2/τ [13], while

126

Dmass ∈ [0.002; 0.02]σ2/τ (Fig. 2). Finally the whole

127

system was kept at fixed pressure p = 0.0015ε/σ3

us-128

ing a Nos´e-Hoover barostat. Simulations were run over

129

typically 107 timesteps in order to accumulate enough

130

statistics.

131

In previous work, we have shown that nanoparticles

132

are able to sustain extremely high heat fluxes, via two

133

mechanisms: Firstly, interfacial thermal resistance at the

134

nanoscale generates significant temperature jumps at the

135

interface, i.e. the contact temperature Tcof the liquid at

136

the nanoparticle surface is much lower than the particle

137

temperature Tp(Fig. 1). Secondly, the large

curvature-138

induced Laplace pressure prevents the formation of a

va-139

por layer at the interface; At the highest temperatures,

140

only a stable depleted region is observed (Fig. 1).

141

Two approaches were used to measure the effective

142

temperature of the particle. We started by measuring

143

the kinetic temperature TK, related to the center of mass

144

velocity of the particle. Due to the finite ratio between

145

solid and liquid masses, care has to be taken to measure

146

the relative velocity between the solid nanoparticle and

147

the liquid Ui = Usi−Uli(i = x, y, z), with Usiand Ulithe

148

velocities of the solid and liquid centers of mass along the

149

i direction. TK was then given by 12kBTK = 12meffhUi2i,

150

where meff= msml/(ms+ ml) [msand ml being the

to-151

tal mass of the solid and liquid components]. We checked

152

that this procedure behaved correctly for all mass ratios,

153

even when the mass of solid atoms is increased artificially

154

up to the point where ms= ml. All the velocity

measure-155

ments presented in the following were done consistently

156

using this procedure. TK was evaluated along the 3

de-157

grees of freedom of the particle in order to estimate the

158

uncertainties, which were below 1%.

159

We also measured the “Einstein” temperature TE,

de-160

fined as the ratio between the diffusion coefficient D and

161

the mobility µ of the particle [9]. The diffusion

coeffi-162

cient was computed as the plateau value of the integral of

163

the velocity autocorrelation function (VACF) CU U(t) =

164

hUi(t) Ui(0)i of the nanoparticle: D = limt→∞D(t), with

165

D(t) =Rt

0CU U(s)ds (Fig. 2a). The plateau is reached

166

after a correlation time typically around tc ∼ 30τ . The

167

mobility µ was computed by applying an external force

168

F = 10ε/σ to the particle, and measuring its steady

ve-169

locity U in the direction of the force: µ = U/F (linear

170

response in the applied force was carefully checked).

171

In Fig. 2.b, we have plotted both measures of the

par-172

ticle’s effective temperature as a function of Tp. One

173 174

can note that all temperature estimates collapse to T0at

175

equilibrium. A striking feature of Fig. 2.b is that the two

176

approaches to measure the effective temperature of the

177

particle provide different results. While this is expected

178

for active colloids with a ballistic motion at short times

179

[3], it is quite surprising in the case of a simple Brownian

180

particle, and cannot be understood in the framework of

181

a Langevin description. As discussed before [5], one can

182

finally note that neither TKnor TEidentify with the

con-183

tact temperature Tc, as could be naively expected [14]

184

(Fig. 2.b).

(4)

1 2 3 Tp [ε/kB] 0.8 1 1.2 1.4 1.6 T eff [ ε / k B ] Tc TK TE 0 50 100 t [τ] 0 0.01 0.02 D ( t ) [ σ 2 /τ ] (a) (b)

FIG. 2. (a) Integrated velocity autocorrelation func-tions of the particle (from bottom to top: kBTp/ε =

0.75, 1.5, 2, 2.5, 3, 3.5). (b) Einstein temperature TE and

ki-netic temperature TK as a function of the particle

tempera-ture Tp; the contact temperature Tc is also plotted for

com-parison. Lines are guides for the eye. When not indicated, uncertainties are below the symbol size.

To understand the existence of two temperatures in

186

the system, we have probed the fluctuation dissipation

187

theorem (FDT) for the Brownian system under study.

188

Generally speaking, considering a physical observable A,

189

the response of a system driven out of equilibrium at time

190

t = 0 by the action of a small external field F (t) is

char-191

acterized by the susceptibility χAC(t) = hδA(t)iδF (0) where in

192

the subscript of the susceptibility, C refers to the

vari-193

able conjugated to the field F : C = δHδF, H being the

194

Hamiltonian of the perturbed system. The FDT states

195

that the susceptibility χAC(t) is related to the

equilib-196

rium correlation function CAC(t) = hA(t)C(0)i through:

197

Rt

0χAC(s)ds = kB1TCAC(t) where T is the thermal bath 198

temperature, and the correlation function is estimated

199

at equilibrium. A sensitive way of probing the

devia-200

tion from this relation in nonequilibrium systems, which

201

has been extensively used for example in glassy systems

202

[15, 16] consists in determining separately the integrated

203

susceptibility function and the correlation function, and

204

in plotting them in a parametric plot with the time as

pa-205

rameter. The slope of the curve is then interpreted as the

206

inverse of an effective temperature, which may depend on

207

the time scale [15].

208

For the system under study, we obtain the integrated

209

response to an external force F by applying the force in a

210

stationary configuration at t = 0, and following the

evo-211

lution of the particle center of mass velocity U (t). The

212

parametric plot involves then the average velocity divided

213

by the applied force, µ(t) = hU (t)i/F = Rt

0χU X(s)ds,

214

versus the integrated velocity auto correlation function

215

CU X(t) = R0tCU U(s)ds = D(t). To obtain the response

216

function from the ensemble averaged particle velocity

217

hU (t)i, we have run simulations starting from 1000

in-218

dependent configurations of the system and tracked the

219 0 5 10 15 D(t) [10-3σ2/τ] 0 5 10 µ (t ) [10 -3 σ 2 /( τε )] 1 2 3 T p [ε / kB] 0.8 1 1.2 1.4 1.6 Teff [ ε / kB ] TK’ TE’ 1/kBTK’ 1/k BTE’

FIG. 3. Integrated response function as a function of the in-tegrated VACF of the nanoparticle, for kBTp/ε = 0.75

(equi-librium) and 3.5. Inset– Temperatures extracted from the fit of the main graph’s curves at small and large times, as a func-tion of the particle temperature. note that the lines are not merely guides to the eye, but correspond to the data deter-mined independently and already reported in Fig. 2 for the kinetic and Einstein temperature.

position of the Brownian particle before a steady state is

220

attained (corresponding to times smaller than tc). This

221

enabled us to obtain good statistics for the ensemble

av-222

eraged velocity of the particle, in particular during the

223

early stage of the transient t ≪ tc.

224

Figure 3 shows the resulting response/correlation

para-225

metric plot, for the different temperatures considered.

226

When Tp= T0, the nanoparticle is at equilibrium before

227

the external force is applied, and the fluctuation

dissi-228

pation theorem is obeyed. For higher values of the

par-229

ticle temperature Tp, the velocity hU (t)i depends non

230

linearly on the integrated VACF and the fluctuation

dis-231

sipation ratio is time dependent. This is particularly

vis-232

ible for the highest temperature considered in Fig. 3

233

Tp = 3.5ε/kB, where the two slopes dDdµ at small and

234

large D differ markedly. From these two slopes, it is

pos-235

sible to define two temperatures T′

K and TE′

characteriz-236

ing the response of the system respectively at short times

237

and long times. The inset of Fig. 3 compares these two

238

temperatures to the kinetic and Einstein temperatures

239

defined before. Strikingly the short time effective

tem-240

perature TK′ is very close to the kinetic temperature of

241

the nanoparticle TK, while the long time effective

temper-242

ature TE′ is close to the Einstein temperature TE.

There-243

fore our system, in spite of its simplicity, exhibits a “two

244

temperatures” behavior on the two different time scales

245

that are separated by the typical scale set by the loss of

246

memory in the initial velocity. The short time, fast

tem-247

perature sets the kinetic energy of the particles, while

248

the Einstein temperature which probes the steady state

249

response is determined by the long time behavior of the

250

integrated response.

251

For a system in contact with a thermal bath and driven

(5)

4 0 5 10 15 20 25 t [τ] 0.8 1.2 1.6 < δ W t 2 >/2<W t > [ ε / k B ] 1 2 3 Tp [ε/kB] 0.8 1 1.2 1.4 1.6 T eff [ ε / k B ] T K T E TTFT TTFT (a) (b)

FIG. 4. (a) Transient fluctuation temperature Tt =

hδW2

ti/2hWti as a function of the time t, for different

tem-peratures Tp of the nanoparticle. From bottom to top:

kBTp/ε = 0.75, 1, 1.5, 2, 2.5, 3, 3.5. (b) Transient fluctuation

temperature TTFTobtained with the long time limit of Ttas

a function of the particle temperature Tp. The lines

corre-spond to the data for the kinetic and Einstein temperature in Fig. 2.

out of equilibrium, the bath temperature plays also a key

253

role in quantifying the fluctuations of the work from an

254

external forcing [10]. Two situations have to be

distin-255

guished depending on the time window analyzed. If we

256

follow the evolution of a system in the transient regime

257

before a steady state is reached, starting from a system

258

at equilibrium, the transient fluctuation theorem (TFT)

259

predicts:

260

P (Wt)/P (−Wt) = exp(Wt/kBT ), (3)

where P (Wt) is the density probability of the work Wt.

261

In this equation Wt is the work from the external force

262

F , i.e. Wt = R0tU (s)F ds and T is the temperature

263

of the thermal bath. On the other hand, in a a

sta-264

tionary situation, the steady state fluctuation theorem

265

(SSFT) predicts P (Wt)/P (−Wt) → exp(Wt/kBT ) when

266

t ≫ tc where tc denotes a typical equilibrium

correla-267

tion time. In the SSFT, the work Wtis estimated along

268

a trajectory of length t: Wt =

Rti+t

ti U (s)F ds, where 269

an average on different values of the initial ti may be

270

performed. We have tested these fluctuation relations

271

for the heated Brownian particles, again applying an

272

external force F = 10ε/σ at t = 0 and recording the

273

statistics of the work using 1000 independent

configura-274

tions. It turned out however that the distribution of the

275

work Wtwas too noisy to determine accurately the ratio

276

P (Wt)/P (−Wt) and critically assess the validity of the

277

fluctuation theorems discussed above. To extract an

ef-278

fective temperature measuring the fluctuations of Wt, we

279

have used the observation that the statistics of the work

280

Wt is to a good approximation Gaussian. Under these

281

conditions, it is trivial to show that the distribution of

282

Wt obeys a law similar to Eq. 3 with an effective

tem-283

perature Tt= hδWt2i/2hWti. Note that strictly speaking

284

the TFT implies that Tt= T is independent of t. In Fig.

285

4.a we have shown the evolution of Tt as a function of

286

the time t for different temperatures Tp of the

nanopar-287

ticle. For all the temperatures considered, the initially

288

small values of hWti leads to a large uncertainty in the

289

value of Tt. For longer times t > 5τ , the temperature

290

Tt is approximately independent of the time t. We will

291

denote TTFT(Tp) the value of the effective temperature

292

Tt in this regime. Figure 4.b displays the evolution of

293

TTFT as a function of the temperature of the

nanoparti-294

cle Tp. It is clear that the resulting TTFT is very close

295

to the kinetic temperature TKcharacterizing the particle

296

dynamics on short time scales. While we are not aware

297

of a theoretical analysis of this situation, we believe the

298

reason for this proximity lies in the fact that the main

299

contribution to fluctuations in the work function

corre-300

sponds to the time regime in which the velocity is still

301

correlated to its value at t = 0, i.e. the same time regime

302

in which the fluctuation dissipation ratio corresponds to

303

the “fast” temperature.

304

Our work shows that, even in a conceptually rather

305

simple system, in a nonequilibrium steady state, a

de-306

scription in terms of a Langevin model involving a single

307

temperature is far from trivial. Further generalization

308

and interpretation of the behavior of interacting

parti-309

cles in terms of Langevin models and a single noise

tem-310

perature is expected to suffer similar difficulties, as can

311

already be inferred from the results of [2]. It would be

312

interesting to explore, if the recent extensions of

fluctu-313

ation dissipation theorems proposed in refs [17, 18] can

314

be applied to the present case, i.e. to identify

observ-315

ables for which a response-correlation proportionality

re-316

lation holds. Even so, the resulting observables are likely

317

to be different from those that are naturally measured

318

in experiments or simulations. We also note that, with

319

the present observables, experiments using optical

tweez-320

ers with a strongly absorbing particle could be used to

321

probe the different temperatures investigated here, with

322

the exception of the kinetic one. We expect that such

323

experiments will be able to detect a deviation from

equi-324

librium of the order of magnitude reported here.

325

We acknowledge useful exchanges with L. Bocquet, F.

326

Cichos, K. Kroy and D. Rings, and the support of ANR

327 project Opthermal. 328 ∗ laurent.joly@univ-lyon1.fr 329 † samy.merabia@univ-lyon1.fr 330 ‡ jean-louis.barrat@univ-lyon1.fr 331

[1] F. Schweizer, Brownian Agents and Active Particles 332

(Springer, Berlin, 2003). 333

[2] D. Loi, S. Mossa, and L. F. Cugliandolo, Physical Review 334

E 77, 051111 (2008). 335

[3] J. Palacci, C. Cottin-Bizonne, C. Ybert, and L. Bocquet, 336

Phys. Rev. Lett. 105, 088304 (2010). 337

(6)

[4] R. Radunz, D. Rings, K. Kroy, and F. Cichos, J. Phys. 338

Chem. A 113, 1674 (2009). 339

[5] D. Rings, R. Schachoff, M. Selmke, F. Cichos, and 340

K. Kroy, Phys. Rev. Lett. 105, 090604 (2010). 341

[6] R. Kubo, Statistical Physics II, nonequilibrium statistical 342

mechanics (Springer, Berlin, 1988). 343

[7] M. G. McPhie, P. J. Daivis, I. K. Snook, J. Ennis, and 344

D. J. Evans, Physica A 299, 412 (2001). 345

[8] T. Speck and U. Seifert, Europhys. Lett. 74, 391 (2006). 346

[9] A. Einstein, Annalen Der Physik 17, 549 (1905). 347

[10] R. van Zon and E. G. D. Cohen, Physical Review E 67, 348

046102 (2003). 349

[11] S. Plimpton, J. Comp. Phys. 117, 1 (1995), 350

http://lammps.sandia.gov/ . 351

[12] S. Merabia, P. Keblinski, L. Joly, L. J. Lewis, and J.-L. 352

Barrat, Physical Review E 79, 021404 (2009). 353

[13] S. Merabia, S. Shenogin, L. Joly, P. Keblinski, and J. L. 354

Barrat, PNAS 106, 15113 (2009). 355

[14] R. M. Mazo, J. Chem. Phys. 60, 2634 (1974). 356

[15] L. F. Cugliandolo, J. Kurchan, and L. Peliti, Physical 357

Review E 55, 3898 (1997). 358

[16] L. Berthier and J. L. Barrat, J. Chem. Phys. 116, 6228 359

(2002). 360

[17] U. Seifert and T. Speck, Europhys. Lett. 89, 10007 361

(2010). 362

[18] J. Prost, J. F. Joanny, and J. M. R. Parrondo, Phys. 363

Rev. Lett. 103, 090601 (2009). 364

Figure

FIG. 1. Left – Snapshot of the simulated system for T p = 3.5ε/k B (T 0 = 0.75ε/k B ); Gray levels indicate the kinetic  en-ergy of atoms
FIG. 2. (a) Integrated velocity autocorrelation func- func-tions of the particle (from bottom to top: k B T p /ε = 0.75, 1.5, 2, 2.5, 3, 3.5)
FIG. 4. (a) Transient fluctuation temperature T t = hδW t 2 i/2hW t i as a function of the time t, for different  tem-peratures T p of the nanoparticle

Références

Documents relatifs

The work done by the external force and the dissipated energy are measured close to the stochastic resonance where the injected power is maximum.. When the double well potential U

Once the bead is confined in this quasi-harmonic potential, acquisitions are performed and the sinusoidal force applied on the bead in the bistable configuration applied on the bead

It consists in the production of a virtual vector boson splitting into a real boson and a scalar boson. Its cross section is lower than the VBF one. In particular, in the search for

In contrast for large α &amp; 1, when the randomness of the system becomes domi- nated by the external stochastic forcing, we have shown that the symmetry relation given by the FT

During this interaction, the rel- evant points are the following: sid diameter increases and vortex stretching is still observed as the ring approaches the wall; siid the boundary

The availability of a two year long data set of these surface variables from SMOSREX (Surface Monitoring Of the Soil Reservoir EXperiment) makes it possible to study the

We have built a realisability framework and with a different but equivalent definition of the Ef- fective Topos, we have proved that the Effective Topos was a topos with an object

Key-words: Feynman–Kac flow, nonnegative selection function, binary selection function, interacting particle system, extinction time, random number of particles, central limit