• Aucun résultat trouvé

Critical behavior in porous media flow

N/A
N/A
Protected

Academic year: 2021

Partager "Critical behavior in porous media flow"

Copied!
8
0
0

Texte intégral

(1)

HAL Id: hal-01628931

https://hal.archives-ouvertes.fr/hal-01628931

Submitted on 5 Nov 2017

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.

Critical behavior in porous media flow

Marcel Moura, Knut Måløy, Renaud Toussaint

To cite this version:

Marcel Moura, Knut Måløy, Renaud Toussaint. Critical behavior in porous media flow. EPL -

Euro-physics Letters, European Physical Society/EDP Sciences/Società Italiana di Fisica/IOP Publishing,

2017, 118 (1), �10.1209/0295-5075/118/14004�. �hal-01628931�

(2)

Critical behavior in porous media flow

M. Moura1, K. J. M˚aløy1 and R. Toussaint2

1 PoreLab, Department of Physics, University of Oslo - PO Box 1048, Blindern, N-0316, Oslo, Norway.

2 Universit´e de Strasbourg, CNRS, IPGS UMR 7516 - 5 rue Ren´e Descartes, 67000, Strasbourg, France.

PACS 47.55.-t – Multiphase and stratified flows

PACS 47.15.gp – Hele-Shaw flows

PACS 64.60.Ht – Dynamic critical phenomena

Abstract –The intermittent burst dynamics during the slow drainage of a porous medium is studied experimentally. We have shown that this system satisfies a set of conditions known to be true for critical systems, such as intermittent activity with bursts extending over several time and length scales, self-similar macroscopic fractal structure and 1/fαpower spectrum. Additionally, we

have verified a theoretically predicted scaling for the burst size distribution, previously assessed via numerical simulations. The observation of 1/fαpower spectra is new for porous media flows and,

for specific boundary conditions, we notice the occurrence of a transition from 1/f to 1/f2scaling. An analytically integrable mathematical framework was employed to explain this behavior.

Introduction. – The topic of fluid motion inside a

1

porous network has deservedly been subjected to a

consid-2

erable number of studies over the past decades. Scientists

3

have studied the morphology and dynamics of the flow

4

[1–12] and proposed a set of numerical schemes able to

5

reproduce the observed macroscopic patterns [13–17] and

6

relevant pore-scale mechanisms [18–26]. The topic is also

7

of central importance for the study of groundwater flows

8

and soil contaminants treatment [27,28] and has direct

ap-9

plications in the energy sector, for example, in

hydrocar-10

bon recovery methods [29]. One particularly interesting

11

aspect of multiphase flow in porous media is its

intermit-12

tent dynamics [3, 4, 18], with long intervals of stagnation

13

followed by short intervals of strong activity. This kind of

14

general behavior [30–32] appears in many physical,

biolog-15

ical and economical systems, such as the stick-slip motion

16

of a block on an inclined plane [33], the propagation of a

17

fracture front in a disordered material [34–36], the number

18

of mutations in models of biological evolution [37], acoustic

19

emissions from fracturing [38, 39], variations in stock

mar-20

kets [40], and the rate of energy transfer between scales in

21

fully developed turbulence [41, 42]. Intermittent

phenom-22

ena arise irrespective of the (certainly different) specific

23

details of each system. In the particular case of porous

24

media flows, it is caused by the interplay between an

ex-25

ternal load (for example, an imposed pressure difference

26

across the system) and the internal random resistance due

27

to the broader or narrower pore-throats.

28

In the present work we show experimental results on

29

the burst dynamics during drainage in artificial porous 30

media and investigate the question of how the pressure 31

fluctuations (due to the burst activity) can encode useful 32

information about the system. The flows studied are slow 33

enough to be in the capillary regime, in which capillary 34

forces are typically much stronger than viscous ones [3,43]. 35

We have employed synthetic quasi-2D systems driven by 36

a controlled imposed pressure (CIP) boundary condition. 37

This boundary condition differs from the controlled with- 38

drawal rate (CWR), more commonly used [3, 9]. The dy- 39

namics is characterized both via direct imaging of the 40

flow and by local pressure measurements. We present 41

results related to the statistics of bursts, their morphol- 42

ogy and orientation within the medium, and the power 43

spectral density (PSD) associated with the fluctuations 44

in the measured pressure signal. In particular, we show 45

that for systems driven by the CIP boundary condition, 46

the PSD presents a 1/f scaling regime. The presence of 47

1/fαpower spectra is a widespread feature occurring in a 48

myriad of contexts [44–46], commonly signaling the collec- 49

tive dynamics of critical systems. Some examples are the 50

early measurements of flicker noise in vacuum tubes [47], 51

fluctuations in neuronal activity in the brain [48], quan- 52

tum dots fluorescence [49], loudness in music and speech 53

[50, 51] and fluctuations in the interplanetary magnetic 54

field [52]. Although 1/fα power spectra have also been

55

observed in some fluid systems, such as simulations and 56

experiments on hydrodynamic and magnetohydrodynamic 57

(3)

M. Moura et al.

netically forced flows [55], to the best of our knowledge the

59

results reported here provide the first experimental

obser-60

vations of 1/fαpower spectra in porous media flows.

61

Methodology. – Fig. 1 shows a schematic

representa-62

tion of the setup employed (additional details in Ref. [56]).

63

The quasi-2D porous network is formed by a modified

64

Hele-Shaw cell filled with a monolayer of glass beads

hav-65

ing diameters a in the range 1.0mm < a < 1.2mm. The

66

beads are kept in place by a pressurized cushion placed on

67

the bottom plate of the cell. A spongeous filter with pores

68

much smaller than those in the medium is placed between

69

the porous network and the outlet of the model. This filter

70

allows the dynamics to continue inside the medium even

71

after breakthrough [56]. Pressure measurements are taken

72

at the outlet with an electronic pressure sensor (Honeywell

73

26PCAFG6G) that records the difference between the air

74

pressure (non-wetting phase) and the liquid pressure

(wet-75

ting phase) at the outlet, i.e., pm= pnw− poutw . Since the

76

inlet is open to the atmosphere, pnw = p0 in all

experi-77

ments, where p0is the atmospheric pressure. The porous

78

matrix was initially filled with a mixture of glycerol (80%

79

in weight) and water (20% in weight) having kinematic

80

viscosity ν = 4.25 10−5m2/s, density ρ = 1.205 g/cm3

81

and surface tension γ = 0.064 N.m−1. We have

per-82

formed experiments on 4 different porous media with

di-83

mensions: (1) 27.3cm x 11.0cm, (2) 14.0cm x 11.5cm,

84

(3) 32.8cm x 14.6cm and (4) 32.0cm x 4.5cm, where the

85

first number corresponds to the length (inlet–outlet

di-86

rection) and the second to the width. The outlet of the

87

model is connected to an external reservoir. The height

88

difference h between the surface of the liquid in this

reser-89

voir and the model is used to control the imposed pressure

90

via an adaptive feedback mechanism (CIP boundary

con-91

dition). This mechanism guarantees that the pressure is

92

only increased when the system is in a quasi-equilibrium

93

situation (details in [56]). By slowly increasing the

im-94

posed pressure (via small steps in the height of the

reser-95

voir dh = 10µm =⇒ dp = ρgdh = 0.12 P a, where g

96

is the acceleration of gravity), new pore-throats may

be-97

come available to invasion. The value of dh was chosen to

98

satisfy the accuracy condition that the height would

typ-99

ically have to be increased several times before new pores

100

are invaded. As long as this condition is satisfied, the

101

results obtained should be independent of the particular

102

value of dh.

103

Burst size distribution. – We begin by analyzing

the size distribution of invasion bursts in a CIP experi-ment. A burst is understood as any connected set of pores

invaded in the interval Θ = t2−t1between two consecutive

time instants, t1and t2, at which the imposed pressure was

increased (i.e., the imposed pressure is constant during the

interval Θ, being changed only at its extremes t1 and t2).

Fig. 2 shows the individual bursts for experiment CIP-1 (the number identifies the model), colored according to their area (top) and randomly (bottom), the latter be-ing done to aid the visualization of separate bursts. Only

Fig. 1: (color online) Diagram of the experimental setup and boundary conditions (CIP or CWR). The numbers (1), (2) and (3) denote the porous medium, filter and external tubing.

bursts having their centroids in the mid 90% of the model’s length are considered, to avoid possible boundary effects [56]. A great deal of information can be obtained from

this image. Initially, one can observe the homogeneity

and isotropicality of the dynamics: the bursts don’t seem to follow a well defined size gradient (the top image does not seem to transition from blue to red following a spe-cific direction), nor have they a clear preferred orientation (they are not particularly elongated in any direction). It is hard, if not impossible, to say from this image in which di-rection the invasion takes place (it is from left to right). A

reflection (vertical or horizontal) or a 180◦rotation would

also not be clearly identified. The box counting fractal di-mension [57,58] of the invading cluster was measured to be D = 1.76 ± 0.05. Fig. 3 shows the burst size distribution N (n) for 3 separate experiments (the number of pores n being measured by normalizing the burst area by a

typi-cal pore area ≈ 0.3mm2). The system exhibits the scaling

N (n) ∝ n−τ, with τ = 1.37 ± 0.08, over at least two

decades. The burst dynamics is therefore spatially self-similar, a feature commonly associated with systems close to a critical transition [46, 57]. The exponent τ has been calculated via maximum likelihood estimation (MLE) [59] using the data from Fig. 2 for burst sizes in the interval 1 pore < n < 150 pores. MLE was used in order to avoid possible biases from data binning (MLE is a binning free method), see also [60]. The scaling is shown in Fig. 3 on top of the logarithmically binned histogram of the data for the sake of visualization. Experiment CIP-4 was left out of the analysis because boundary effects rendered the results unreliable (model 4 is too narrow). The measured expo-nent is consistent with the value τ = 1.30 ± 0.05 predicted by numerical simulations and percolation theory [21, 58]. Martys et al. [21] derived the analytical form

τ = 1 +De− 1/ν

0

D , (1)

where D and De are respectively the fractal dimensions 104

of the growing cluster and its external perimeter and ν0= 105

(4)

Burst size (number of por es) Bursts c ol ored r andomly

Fig. 2: (color online) Individual bursts for experiment CIP-1. The flow is from left to right, during ≈ 82h. Bursts color coded by their size normalized by a typical pore area (top) and randomly (bottom). The vast blue areas in the top image contain many smaller bursts (detail).

correlation length [57, 58]. Using the values D = 1.76

107

and De= 4/3 [14], we obtain τ = 1.33, very close to the

108

measured value τ = 1.37 ± 0.08 shown in Fig. 3. Our

109

measurements provide a direct experimental verification

110

of Eq. (1), proposed in Ref. [21].

111

Crandall et al. [61] performed measurements in a CWR

112

system finding the exponent τ = 1.53, which is compared

113

to the theoretical prediction of τ = 1.527 from Roux

114

and Guyon [62]. Nevertheless, Maslov [63] pointed out

115

an inconsistency in this theoretical prediction, the

cor-116

rect expression being given in Eq. (1). Modified invasion

117

percolation simulations and pressure measurements [3, 4]

118

have shown that, in a CWR, system very large bursts are

119

split into smaller ones. A burst size distribution was

ob-120

served, with exponent τ = 1.3 ± 0.05 for the simulations

121

and τ = 1.45 ± 0.10 for the experiments (consistent with

122

Eq. (1)), followed by an exponential cutoff [3, 4]. In the

123

CIP case large bursts can happen because the displaced

124

liquid can freely flow out of the model but in the CWR

125

case this is not possible since the available volume for the

126

displaced liquid is bounded by the outlet syringe volume.

127

Burst time distribution. – Let us now focus on the

128

distribution G(Θ) of time intervals Θ between two

succes-129

sive increments in the imposed pressure during which

inva-130

sion bursts have occurred. Fig. 4 shows this distribution,

131

produced for all bursts with Θ > 120s, a cutoff related

132

to the minimum time difference for proceeding the image

133

analysis used in the feedback mechanism [56]. It scales as

134

G(Θ) ∝ Θ−γwith γ = 2.04±0.15 (exponent was also

com-135

puted via MLE [59]). In the inset we show the distribution

136

of inverse intervals g(1/Θ), which is nearly uniform, since

137

it is related to G(Θ) by g(1/Θ) = G(Θ)Θ2 ∝ Θ2−γ. The

138

uniformity of g(1/Θ) will play an important role further

139

Fig. 3: (color online) Burst size distribution N (n). The line shows the scaling N (n) ∝ n−τ, with τ = 1.37 ± 0.08, which is consistent with the theoretical value τ = 1.30 ± 0.05 predicted by numerical simulations and percolation theory [21, 58]. The data has been shifted vertically to aid visualization.

on in the modeling of the pressure fluctuations PSD. 140

Connection between the burst size and time dis- 141

tributions. – We consider now the link between the 142

burst size distribution N (n) shown in Fig. 3 and the burst 143

time distribution G(Θ) in Fig. 4. Let ˙A = s/Θ denote 144

the average growth rate of a burst of area s during the 145

time interval Θ. This corresponds to an external perimeter 146

growth [57, 58], therefore ˙A ∝ ule where u is a character- 147

istic front speed (set by the Darcy law and the character- 148

istic capillary pressure) and le is the external perimeter, 149

related to the linear size across a cluster l as le ∝ lDe. 150

Since s ∝ lD, we have

151

s/Θ = ˙A ∝ ule∝ sDe/D =⇒ Θ ∝ sβ, (2)

with β = 1 − De/D. The distributions of s and Θ are 152

linked by |G(Θ)dΘ| = |p(s)ds| and since the area s of a 153

burst is proportional to its number of pores n (see Fig. 3), 154

it follows that p(s) ∝ s−τ. Therefore, 155

G(Θ) ∝ s−τds/dΘ =⇒ G(Θ) ∝ Θ−γ, (3)

with γ = (τ − 1 + β) /β = (τ − De/D) / (1 − De/D). Us- 156

ing the literature values τ = 1.3 [21, 58], De = 1.33 157

and D = 1.82 [58], we find γ = 2.11, quite close to the 158

measured value γ = 2.04 seen in Fig. 4. As an imme- 159

diate consequence of Eq. (3), the distribution of inverse 160

intervals scales as g(1/Θ) ∝ Θ−η, with η = γ − 2 = 161

(τ − 2 + De/D) / (1 − De/D). Using the literature values 162

above we find η = 0.11, which is in agreement with the 163

experimentally observed value η = γ − 2 = 0.04 ± 0.15 164

– i.e. these theoretical considerations explain the nearly 165

uniform distribution observed in the inset of Fig. 4. 166

Fluctuations in the measured pressure signal. – Next, we analyze the fluctuations in the pressure signal, following the pore invasion events. In Fig. 5, we show the typical pressure signature in a CIP experiment. The observed pressure pulses present a characteristic exponen-tial relaxation. We also observe that a pulse can trigger

(5)

M. Moura et al. 103 104 10-2 10-1 100 10-4 10-3 102 104

Fig. 4: (color online) Burst time distribution G(Θ). The scaling (red line) corresponds to G(Θ) ∝ Θ−γ with γ = 2.04 ± 0.15. In the inset we show the nearly uniform distribution g(1/Θ).

others and even give rise to large avalanches with the in-vasion of several pores. A pulse can be divided into two

phases: an initial fast drop in the capillary pressure pcand

a slower exponential relaxation back to the pressure level

ρgh set externally (see Fig. 1). The fast drop in pcoccurs

as the liquid is displaced (following the invasion of one or more pores) and subsequently redistributed to the sur-rounding menisci, causing a back-contraction of the inter-face [3,4,18]. The relaxation phase occurs as the liquid-air interface readjusts itself inside the available pore-throats and the liquid volume displaced from the pores flows out of the model. The fluid motion sets in viscous pressure drops which are reflected in the measured pressure, as seen in Fig. 5. These drops occur (see Fig. 1): 1) in the porous medium itself, 2) in the filter at the model’s outlet and 3) in the external tubing (the numbers are in correspondence with Fig. 1). The height difference h between the surface of the liquid in the reservoir and the model level accounts for a hydrostatic component ρgh. Adding these contribu-tions and assuming that the flow is governed by Darcy’s equation, we have pw− u µL1 k1 − uµL2 k2 − uS1µL3 S3k3 + ρgh = p0, (4)

where pwis the pressure in the wetting phase (liquid) just

167

after the liquid-air interface, u is the average Darcy

ve-168

locity of the flow in the porous network, µ = ρν is the

169

liquid’s dynamic viscosity, Li and ki with i = {1, 2, 3}

170

are the length and permeability respectively of the porous

171

network, filter and the tubing and S1 and S3 are the

172

respective cross sections of the model and the tubing.

173

Since the capillary pressure across the liquid-air interface

174 is pc= pnw− pw= p0− pw, Eq. (4) becomes 175 pc+ uR − ρgh = 0 , (5) where 176 R = R1+ R2+ R3 =⇒ R = µL1 k1 +µL2 k2 +S1µL3 S3k3 , (6) Pressure [Pa] 6.35 6.4 6.45 6.5 6.55 6.6 340 360 380 400 420

exponential signature of CIP pressure pulses

100s Time [s] ×104 ×104 Pressure [Pa] 100s 420 7 7.05 7.1 400 380 Time [s] multiple pulses superposition of pulses effectively increases

the decay time

Fig. 5: (color online) Typical exponential relaxation signature of pressure pulses. A pulse can trigger others and even give rise to a large avalanche (shown in the inset).

is equivalent to an effective resistance to the flow. The 177

volumetric flux in a pore is dV /dt = ua2/φ, where a is 178

a characteristic pore length scale (for example the bead 179

diameter) and φ is the porosity of the model. By intro- 180

ducing the concept of a capacitive volume κ = dV /dpc 181

(used first in Ref. [3]), where dV is the liquid volume dis- 182

placed from a pore throat in response to a change dpc in 183

capillary pressure, we have 184

dV dt = ua2 φ =⇒ u = κφ a2 dpc dt . (7)

Plugging this equation into Eq. (5), 185

κφR

a2

dpc

dt + pc− ρgh = 0 =⇒ pc(t) = ρgh + Ce

−t/tc, (8)

thus producing the exponential behavior seen in Fig. 5.

C = pc(0) − ρgh < 0 is a constant associated to how much

the capillary pressure decreases during the invasion of a set of pores before it starts to rise again. The characteristic time scale of the exponential decay is

tc=

κφR

a2 . (9)

The invasion of one pore quite frequently triggers the 186

invasion of others, in such a manner that before an ex- 187

ponential pulse decays completely, another one is seen 188

in the pressure signal, see Fig. 5. This mechanism de- 189

lays the complete relaxation of the pressure, effectively 190

increasing the decay time from tc to t∗ ≥ tc. Indeed, if 191

this relaxation-delaying mechanism was absent, the burst 192

time distribution G(Θ) shown in Fig. 4 should be peaked 193

around the value Θ = tc. Since we have shown that 194

G(Θ) ∝ Θ−γ we expect the effective exponential decay 195

time t∗ to follow the same distribution and, in particu- 196

lar, the effective decay rate λ = 1/t∗ should be uniformly

197

distributed in an interval [λmin, λmax] following the same 198

distribution as 1/Θ (see inset of Fig. 4). λmaxis related to 199

(6)

Fig. 6: (color online) Power spectral density comparison for CIP experiments (model’s numbers in the legend). Guide-to-eye lines are shown for the scaling S(f ) ∝ f−α, with α = 1 for lower frequencies and α = 2 for intermediate frequencies.

consider λmin= 0 for convenience. Later on we will show

201

that the distribution of decay rates has a crucial impact

202

on the power spectrum of the pressure signal.

203

Pressure signal PSD. – Next we analyze the power

204

spectral density (PSD) associated to the pressure signal

205

for the CIP experiments. The PSD S = S(f ) was

com-206

puted for all experiments using the Welch method [64].

207

We have noticed the existence of a 1/f scaling regime

208

(flicker/pink noise) for lower frequencies, followed by a

209

crossover and a 1/f2 scaling regime (brown noise) for

in-210

termediate frequencies. For higher frequencies, another

211

crossover follows and a region independent of f is seen

212

(white noise associated with fluctuations in the pressure

213

sensor and unimportant to our analysis). We see from

214

Fig. 6 that the scaling properties of the power spectrum,

215

in particular the occurrence of 1/f noise, seem to remain

216

unchanged despite the changes in both sample dimensions

217

and pore-size distribution (the samples were rebuilt before

218

each experiment, thus changing the pore-size distribution

219

[56]). The 1/f regime is associated with events having

fre-220

quency f < 10−2Hz, or alternatively, periods T > 100s.

221

From Fig. 5, we see that this corresponds to the

charac-222

teristic time intervals between the pressure pulses, thus

223

indicating that they are associated with the presence of

224

the 1/f scaling in the PSD.

225

Analytical modeling of the pressure signal and

226

PSD scaling explanation. – The non-trivial scaling

227

of the CIP power spectral density can be explained by the

228

following mathematical framework, which is an adaptation

229

of an argument proposed in [65] to explain a similar 1/f

230

to 1/f2transition in the very first reported observation of

231

1/f noise [47] (see also [66] and [67]). Apart from a nearly

232

constant offset, the pressure signal can be modeled as a

233

train of exponentially decaying pulses located at randomly

234

distributed discrete times tj,

235

pλ(t) =

X

j

AH(t − tj)e−λ(t−tj), (10)

where λ > 0 and A < 0 are initially taken to be constants 236

(the characteristic decay rate and amplitude of the pulses) 237

and H(t − tj) is the Heaviside step function, i.e., H(t − 238

tj) = 0 if t < tj and H(t − tj) = 1 if t ≥ tj. Let Pλ(f ) be 239

the Fourier transform of pλ(t). The PSD Sλ(f ) is 240

Sλ(f ) = lim T →∞ 1 T D |Pλ(f )| 2E = A 2r λ2+ 4π2f2, (11)

where r is the average rate of occurrence of pulses and the 241

brackets are the expected value operator (since in practice 242

one does not have access to an ensemble of measurements, 243

we have employed Welch’s method [64] to estimate the 244

PSD, which is based on the concept of a periodogram [68]). 245

The PSD shown in Eq. (11) is a Lorentzian curve which is 246

approximately constant for lower frequencies (f  λ/2π) 247

and decays as 1/f2 for higher frequencies (f  λ/2π).

248

A model with a single constant decay rate λ cannot 249

incorporate the 1/f region but, as previously argued, we 250

expect λ to follow the uniform distribution ξ(λ) = 1/λmax 251

in the interval [0, λmax]. Taking this distribution into ac- 252

count and writing λmax= 2 π ft, we have 253

S(f ) = Z λt 0 Sλ(f )ξ(λ)dλ = A2r 4π2f tf arctan ft f  . (12) Eq. (12) has the asymptotic behavior

S(f ) = (A2r 8πft 1 f if f  ft A2r 4π2 1 f2 if f  ft , (13)

thus presenting the 1/f to 1/f2transition observed in the

254

experiments. The transition frequency ft in experiment 255

CIP-1 is roughly ft = 1.5 10−2Hz (see Fig. 6). By us- 256

ing the constant A2r as a fitting parameter we can com- 257

pare the measured PSD with the theoretical prediction in 258

Eq. (12). Fig. 7 shows the resulting comparison produced 259

using A2r = 1.5 P a2/s. The dashed red vertical line marks 260

the transition frequency ft. The analytical result repro- 261

duces the experimental findings very well, scaling as 1/f 262

for f  ftand as 1/f2for f  ft. Indeed, this theory not 263

only captures the 1/f and 1/f2 domains but also fits the

264

data well for the crossover region between these domains. 265

The transition frequency ft can be estimated using 266

Eq. (9) and the resistance R from Eq. (6). As a first 267

order approximation, let us consider only the contribu- 268

tion to R from the term R1 relative to the resistance in 269

the porous medium itself. Using µ = ρν = 5.1 10−2P a.s, 270

L1 = 0.27m, a = 10−3m, κ = 1.1 10−12m3/P a (from 271

Ref. [3]), k1 = 1.6 10−9m2 and φ = 0.63 (both measured 272

in a similar model in Ref. [5]), we find 273

ft= 1 2πtc = k1a 2 2πκφµL1 =⇒ ft≈ 2.6 10−2Hz , (14)

(7)

M. Moura et al.

not far from the transition frequency ft = 1.5 10−2Hz

274

shown in Fig. 7. The overestimation comes from the terms

275

R2 and R3 in Eq. (6), ignored in the calculation above.

276

Finally, notice also the existence of a single isolated

277

point in the very low frequency part of the PSD, falling far

278

from the scaling region (extreme left for all experiments

279

in Fig. 6). This point is not an outlier in the data: its

280

existence signals the very slow positive drift of the

pres-281

sure signal, which occurs since the capillary pressure has

282

to increase to allow the invasion of narrower pores [3, 56].

283

Comparison with a system driven under a CWR

284

boundary condition. – In order to test the effect of

285

the boundary conditions in the PSD, we have run a

con-286

trolled withdrawal rate (CWR) experiment using model

287

(1). The resulting PSD is shown in the inset of Fig. 7.

288

The PSD still presents an interesting scaling, but with

dif-289

ferent scaling regimes: 1/f1.5, for lower frequencies, and

290

1/f3.5, for intermediate frequencies. The 1/f region is

291

only observed for systems driven under the CIP boundary

292

condition. The fact that the exponents for CWR differ

293

from CIP is not surprising, since the pressure relaxation

294

in that case no longer exponential, but linear, see Ref. [3].

295

Connection between the measured pressure and

296

the capillary pressure. – The pressure sensor

mea-297

sures the difference between the pressure in the air and the

298

liquid at the outlet, i.e., pm= pnw− poutw . The measured

299

signal is not exactly the capillary pressure pc = pnw− pw

300

across the liquid-air interface, since pw 6= poutw given that

301

viscous losses occur between the liquid-air interface and

302

the outlet, thus generally making pw> poutw . Those losses

303

occur in the porous medium itself and in the filter at the

304

outlet of the model (numbers 1 and 2 in Fig. 1). The

305

connection between pm and pc is pm = pc+ u (R1+ R2),

306

where R1and R2are the resistance terms from the porous

307

network and the filter. Using Eqs. (7) and (8), we have

308 pm= ρgh + C  1 − R1+ R2 R1+ R2+ R3  e−t/tc. (15)

Therefore, by comparing Eqs. (8) and (15), we see that

309

pmdiffers from pconly in the amplitude of the pulses, but

310

not in their characteristic exponential decay. Since our

311

analysis depended only on the distribution of the decay

312

rates, the differences between pmand pc are not crucial.

313

Further generalizations of the PSD analytical

314

framework. – One possible generalization of the model

315

would be to consider a system with a distribution of

am-316

plitudes A instead of a single value (as we might expect

317

from Fig. 5). In this case the scaling properties of the

318

PSD would still be left unchanged but the constant A2in

319

Eq. (12) and (13) would be replaced by the expected value

320

of A2. Another possibility would be to consider a

distri-321

bution for λ of the form ξ(λ) ∝ λ−δ. Here the 1/f2region

322

is still left unchanged but the 1/f scaling is changed to

323 10-3 10-2 10-1 100 102 104 CIP-1 (experiment) CIP-1 (theory) 10-3 10-2 10-1 100 105 CWR-1 (experiment) α= 1 α= 2 α= 1.5 α= 3.5

Fig. 7: (color online) Comparison between theoretical predic-tion and experiments. The analytical result (thin blue line) is given by Eq. (12), where ft= 1.5 10−2Hz (vertical dashed red

line) and A2r = 1.5P a2/s. The analytical prediction match the experimental measurements (green crosses, experiment CIP-1) well. On the inset we show the PSD for experiment CWR-1.

1/f(1+δ)[69]. As previously noted, the distribution of

de-324

caying rates λ is the crucial figure behind the 1/f scaling. 325

Conclusions. – We have analyzed the burst dynam- 326

ics from slow drainage experiments in porous media. We 327

showed that this dynamics presents many features com- 328

monly associated to critical systems. Intermittent bursts 329

of activity were observed over many time and length scales 330

and a theoretical expression for their size distribution scal- 331

ing, Eq. (1), was verified experimentally. The pressure 332

signal of the invasion presented an interesting PSD scal- 333

ing, with a 1/f scaling region which further transitions to 334

1/f2in the case of the CIP boundary condition. We have

335

employed an analytical framework [65] which satisfactorily 336

reproduces the scaling properties of the PSD. The deriva- 337

tion of closed expressions relating the pressure signal PSD 338

to properties of the porous medium and the fluids can lead 339

to new techniques for indirectly probing such systems. For 340

example, if one has access to the PSD only and not to the 341

full pressure signal, the transition frequency ftcan still be 342

measured and information on the ratio k1/φ between the 343

permeability and the porosity of the medium can be found 344

via Eq. (14). If the PSD and ftare known, Eq. (12) can be 345

fitted to measure the product A2r between the amplitudes 346

and rate of occurrence of bursts. 347

∗ ∗ ∗

We acknowledge the support from the University of 348

Oslo, University of Strasbourg, the Research Council of 349

Norway through its Centre of Excellence funding scheme 350

with project number 262644, the CNRS-INSU ALEAS 351

program and the EU Marie Curie ITN FLOWTRANS net- 352

(8)

REFERENCES

354

[1] M˚aløy K. J., Feder J. and Jøssang T., Phys. Rev.

355

Lett., 55 (1985) 2688.

356

[2] Lenormand R., Proc. R. Soc. Lond. A, 423 (1989) 159.

357

[3] M˚aløy K. J., Furuberg L., Feder J. and Jøssang T.,

358

Phys. Rev. Lett., 68 (1992) 2161.

359

[4] Furuberg L., M˚aløy K. J. and Feder J., Phys. Rev.

360

E, 53 (1996) 966.

361

[5] Løvoll G., M´eheust Y., Toussaint R., Schmittbuhl

362

J. and M˚aløy K. J., Phys. Rev. E, 70 (2004) 026301.

363

[6] Toussaint R., Løvoll G., M´eheust Y., M˚aløy K. J.

364

and Schmittbuhl J., Europhys. Lett., 71 (2005) 583.

365

[7] Or D., Adv. Water Resour., 31 (2008) 1129 .

366

[8] Sandnes B., Flekkøy E., Knudsen H., M˚aløy K. and

367

See H., Nat. Commun., 2 (2011) 288.

368

[9] Moebius F. and Or D., J. Colloid Interface Sci., 377

369

(2012) 406 .

370

[10] de Anna P. et al., Phys. Rev. Lett., 110 (2013) 184502.

371

[11] MacMinn C. W., Dufresne E. R. and Wettlaufer

372

J. S., Phys. Rev. X, 5 (2015) 011020.

373

[12] Bultreys T. et al., Water Resour. Res., 51 (2015) 8668.

374

[13] Wilkinson D. and Willemsen J. F., J. Phys. A: Math.

375

Gen., 16 (1983) 3365.

376

[14] Furuberg L., Feder J., Aharony A. and Jøssang T.,

377

Phys. Rev. Lett., 61 (1988) 2117.

378

[15] Rothman D. H., J. Geophys. Res., 95 (1990) 8663.

379

[16] Misztal M. et al., Frontiers in Physics, 3 (2015) .

380

[17] Ferrari A., Jimenez-Martinez J., Borgne T. L.,

381

M´eheust Y. and Lunati I., Water Resour. Res., 51

382

(2015) 1381.

383

[18] Haines W. B., The Journal of Agricultural Science, 20

384

(1930) 97.

385

[19] Morrow N. R., Industrial & Engineering Chemistry, 62

386

(1970) 32.

387

[20] Lenormand R., Zarcone C. and Sarr A., Journal of

388

Fluid Mechanics, 135 (1983) 337.

389

[21] Martys N., Robbins M. O. and Cieplak M., Phys.

390

Rev. B, 44 (1991) 12294.

391

[22] Løvoll G., M´eheust Y., M˚aløy K. J., Aker E. and

392

Schmittbuhl J., Energy, 30 (2005) 861.

393

[23] Berg S. et al., Proc. Natl. Acad. Sci. U.S.A., 110 (2013)

394

3755.

395

[24] Holtzman R. and Segre E., Phys. Rev. Lett., 115

396

(2015) 164501.

397

[25] Trojer M., Szulczewski M. L. and Juanes R., Phys.

398

Rev. Applied, 3 (2015) 054008.

399

[26] Schlter S. et al., Water Resour. Res., 52 (2016) 2194.

400

[27] Guymon G., Unsaturated zone hydrology (Prentice Hall,

401

Englewood Cliffs, N.J) 1994.

402

[28] Jellali S., Muntzer P., Razakarisoa O. and

403

Sch¨afer G., Transport in Porous Media, 44 (2001) 145.

404

[29] Tweheyo M., Holt T. and Torsæter O., J. Pet. Sci.

405

Eng., 24 (1999) 179 .

406

[30] Pomeau Y. and Manneville P., Commun. Math. Phys.,

407

74 (1980) 189.

408

[31] Manneville, P., J. Phys. France, 41 (1980) 1235.

409

[32] Hirsch J. E., Huberman B. A. and Scalapino D. J.,

410

Phys. Rev. A, 25 (1982) 519.

411

[33] Gomes M. A. F., Souza F. A. O. and Brito V. P., J.

412

Phys. D: Appl. Phys., 31 (1998) 3223.

413

[34] M˚aløy K. J., Santucci S., Schmittbuhl J. and

Tou-414

ssaint R., Phys. Rev. Lett., 96 (2006) 045501. 415

[35] Grob M. et al., Pure Appl. Geophys., 166 (2009) 777. 416

[36] Tallakstad K. T., Toussaint R., Santucci S., 417

Schmittbuhl J. and M˚aløy K. J., Phys. Rev. E, 83 418

(2011) 046108. 419

[37] Sneppen K., Bak P., Flyvbjerg H. and Jensen M. H., 420

Proc. Natl. Acad. Sci. U.S.A., 92 (1995) 5209. 421

[38] Sethna J. P., Dahmen K. A. and Myers C. R., Nature, 422

410 (2001) 242. 423

[39] Stojanova M., Santucci S., Vanel L. and Ramos O., 424

Phys. Rev. Lett., 112 (2014) 115502. 425

[40] Liu Y. et al., Phys. Rev. E, 60 (1999) 1390. 426

[41] Kolmogorov A. N., J. Fluid Mech., 13 (1962) 82. 427

[42] Salazar D. S. P. and Vasconcelos G. L., Phys. Rev. 428

E, 82 (2010) 047301. 429

[43] Lenormand R. and Zarcone C., Transport in Porous 430

Media, 4 (1989) 599. 431

[44] Press W. H., Comments on Astrophysics, 7 (1978) 103. 432

[45] Mandelbrot B., The fractal geometry of nature (W.H. 433

Freeman, San Francisco) 1982. 434

[46] Schroeder M. R., Fractals, chaos, power laws: min- 435

utes from an infinite paradise (W.H. Freeman, New York) 436

1991. 437

[47] Johnson J. B., Phys. Rev., 26 (1925) 71. 438

[48] Novikov E., Novikov A., Shannahoff-Khalsa D., 439

Schwartz B. and Wright J., Phys. Rev. E, 56 (1997) 440

R2387. 441

[49] Pelton M., Smith G., Scherer N. F. and Marcus 442

R. A., Proc. Natl. Acad. Sci. U.S.A., 104 (2007) 14249. 443

[50] Voss R. and Clarke J., Nature, 258 (1975) 317. 444

[51] Gardner M., Scientific American, 238 (1978) 16. 445

[52] Matthaeus W. H. and Goldstein M. L., Phys. Rev. 446

Lett., 57 (1986) 495. 447

[53] Bourgoin M. et al., Phys. Fluids, 14 (2002) 3046. 448

[54] Dmitruk P. and Matthaeus W. H., Phys. Rev. E, 76 449

(2007) 036305. 450

[55] Herault J., P´etr´elis F. and Fauve S., EPL, 111 451

(2015) 44002. 452

[56] Moura M., Fiorentino E.-A., M˚aløy K. J., Sch¨afer 453

G. and Toussaint R., Water Resour. Res., 51 (2015) 454

8900. 455

[57] Feder J., Fractals (Plenum Press, New York) 1988. 456

[58] Stauffer D., Introduction to percolation theory (Taylor 457

& Francis, London Bristol, PA) 1994. 458

[59] Clauset A., Shalizi C. R. and Newman M. E. J., 459

SIAM Review, 51 (2009) 661. 460

[60] Iglauer S. and W¨ulling W., Geophys. Res. Lett., 43 461

(2016) 11,253 2016GL071298. 462

[61] Crandall D., Ahmadi G., Ferer M. and Smith D. H., 463

PHYSICA A, 388 (2009) 574 . 464

[62] Roux S. and Guyon E., J. Phys. A, 22 (1989) 3693. 465

[63] Maslov S., Phys. Rev. Lett., 74 (1995) 562. 466

[64] Welch P., IEEE Transactions on Audio and Electroa- 467

coustics, 15 (1967) 70. 468

[65] Ziel A. V. D., Physica, 16 (1950) 359 . 469

[66] Bernamont J., Ann. Phys. (Leipzig), 7 (1937) 71. 470

[67] Milotti E., ArXiv Physics e-prints, (2002) . 471

[68] Press W., Numerical recipes in FORTRAN: the art of 472

scientific computing (Cambridge University Press, Cam- 473

bridge England New York, NY, USA) 1992. 474

Figure

Fig. 2 shows the individual bursts for experiment CIP-1 (the number identifies the model), colored according to their area (top) and randomly (bottom), the latter  be-ing done to aid the visualization of separate bursts
Fig. 3: (color online) Burst size distribution N (n). The line shows the scaling N(n) ∝ n −τ , with τ = 1.37 ± 0.08, which is consistent with the theoretical value τ = 1.30 ± 0.05 predicted by numerical simulations and percolation theory [21, 58]
Fig. 4: (color online) Burst time distribution G(Θ). The scaling (red line) corresponds to G(Θ) ∝ Θ −γ with γ = 2.04 ± 0.15.
Fig. 6: (color online) Power spectral density comparison for CIP experiments (model’s numbers in the legend)
+2

Références

Documents relatifs

Dans ce chapitre j’ai parlé sur la thérapie avec ces types, et j’ai fait un montage en Proteus qui représente le fonctionnement d’un électrostimulateur pulsé à partir

2 - Communication de la Commission au Parlement européen, au Conseil, au Comité économique et social européen et au Comité des régions, Responsabilité sociale des entreprises

• les fichiers de données (data files), qui peuvent être connus en interrogeant la vue V$DATAFILE du dictionnaire des données, contenant le dictionnaire des données et les

12 Apparent viscosity of foam versus total flow rate in 1 mm glass bead main

Using the number of preschool children for her parenting specification, Taniguchi (2002) could not confirm that women pursue self-employment as a means to mitigate

Generally speaking, in a discrete simulation environment, it is very hard to implement dynamic controls over simulation components; especially those passive objects generated and

Comme son grand frère EAnalysis est un logiciel d’aide à l’analyse musicale, mais il est spécialisé dans le domaine de la musique électroacoustique.. L’objectif

Chez Complicite, le discours mathématique et neurologique apporte ainsi de nouvelles formes à l’esthétique du fragment caractéristique de la compagnie, et la partition