+ All Categories
Home > Documents > 2 suppression of shear dilatancy by internal acoustic...

2 suppression of shear dilatancy by internal acoustic...

Date post: 19-Jul-2018
Category:
Upload: hoangdung
View: 217 times
Download: 0 times
Share this document with a friend
62
Auto-acoustic compaction in steady shear flows: Experimental evidence for 1 suppression of shear dilatancy by internal acoustic vibration 2 Nicholas J. van der Elst 1 , Emily E. Brodsky 1 , Pierre-Yves Le Bas 2 and Paul A. Johnson 2 3 1 Dept. of Earth and Planet. Science, 1156 High St., Univ. of California, Santa Cruz, California 95060 4 2 Geophysics Group, Los Alamos National Laboratory, P.O. Box 1663, Los Alamos, New Mexico 87545 5 6 Abstract: Granular shear flows are intrinsic to many geophysical processes, ranging from 7 landslides and debris flows to earthquake rupture on gouge-filled faults. The rheology of 8 a granular flow depends strongly on the boundary conditions and shear rate. Earthquake 9 rupture involves a transition from quasi-static to rapid shear rates. Understanding the 10 processes controlling the transitional rheology is crucial for understanding the rupture 11 process and the coseismic strength of faults. Here we explore the transition 12 experimentally using a commercial torsional rheometer. We measure the dilatation of a 13 steady shear flow at velocities between 10 -3 and 10 2 cm/s, and observe that dilatation is 14 suppressed at intermediate velocities (0.1 - 10 cm/s) for angular particles, but not for 15 smooth glass beads. The maximum reduction in thickness is on the order of 10% of the 16 active shear zone thickness, and scales with the amplitude of shear-generated acoustic 17 vibration. By examining the response to externally applied vibration, we show that the 18 suppression of dilatancy reflects a feedback between internally generated acoustic 19 vibration and granular rheology. We link this phenomenon to acoustic compaction of a 20 dilated granular medium, and formulate an empirical model for the steady-state thickness 21 of a shear-zone in which shear-induced dilatation is balanced by a newly-identified 22 mechanism we call auto-acoustic compaction. This mechanism is activated when the 23 acoustic pressure is on the order of the confining pressure, and results in a velocity- 24
Transcript

Auto-acoustic compaction in steady shear flows: Experimental evidence for 1

suppression of shear dilatancy by internal acoustic vibration 2

Nicholas J. van der Elst1, Emily E. Brodsky1, Pierre-Yves Le Bas2 and Paul A. Johnson2 3

1Dept. of Earth and Planet. Science, 1156 High St., Univ. of California, Santa Cruz, California 95060 4 2Geophysics Group, Los Alamos National Laboratory, P.O. Box 1663, Los Alamos, New Mexico 87545 5

6

Abstract: Granular shear flows are intrinsic to many geophysical processes, ranging from 7

landslides and debris flows to earthquake rupture on gouge-filled faults. The rheology of 8

a granular flow depends strongly on the boundary conditions and shear rate. Earthquake 9

rupture involves a transition from quasi-static to rapid shear rates. Understanding the 10

processes controlling the transitional rheology is crucial for understanding the rupture 11

process and the coseismic strength of faults. Here we explore the transition 12

experimentally using a commercial torsional rheometer. We measure the dilatation of a 13

steady shear flow at velocities between 10-3 and 102 cm/s, and observe that dilatation is 14

suppressed at intermediate velocities (0.1 - 10 cm/s) for angular particles, but not for 15

smooth glass beads. The maximum reduction in thickness is on the order of 10% of the 16

active shear zone thickness, and scales with the amplitude of shear-generated acoustic 17

vibration. By examining the response to externally applied vibration, we show that the 18

suppression of dilatancy reflects a feedback between internally generated acoustic 19

vibration and granular rheology. We link this phenomenon to acoustic compaction of a 20

dilated granular medium, and formulate an empirical model for the steady-state thickness 21

of a shear-zone in which shear-induced dilatation is balanced by a newly-identified 22

mechanism we call auto-acoustic compaction. This mechanism is activated when the 23

acoustic pressure is on the order of the confining pressure, and results in a velocity-24

weakening granular flow regime at shear rates four orders of magnitude below those 25

previously associated with the transition out of quasi-static granular flow. 26

27

1. Introduction 28

Frictional sliding processes in geophysics often involve granular shear flows at the 29

sliding interface. This is true for landslides and debris flows, as well as for earthquake 30

ruptures within granulated damage zones or gouge-filled faults. The frictional strength in 31

these contexts is controlled by the rheology of the granular flow, which has a strong 32

dependence on shear rate and boundary conditions [Campbell, 2006; Clement, 1999; 33

Iverson, 1997; Savage, 1984]. 34

35

For different shear rates, confining stresses, and packing densities, the description of a 36

granular flow can range from “solid-like” to “gas-like” [Jaeger et al., 1996], albeit with 37

complicated second-order behavior in each regime. In the solid-like or quasi-static 38

regime, forces are transmitted elastically through a network of grain contacts, called force 39

chains, and the shear and normal stresses at the boundaries are related to the stiffness and 40

orientations of these chains (Fig. 1a) [Majmudar and Behringer, 2005]. In this regime, 41

force chains are continually created and destroyed through shearing, but the rate of 42

buckling and destruction of old force chains is equal to the rate of creation of new ones, 43

and the shear resistance is, to first order, independent of the shear rate (Fig 1c). This 44

results in a solid-like frictional rheology. In the gas-like, inertial flow regime, stresses 45

are supported through grain-grain or grain-boundary collisions, analogous to a kinetic gas 46

model (Fig. 1b). The shear and normal stresses are related to the particle momentum 47

transfer rate, resulting in a power-law viscous-like rheology (Fig. 1c) [Bagnold, 1954; 48

Campbell, 2005]. 49

50

Earthquake rupture involves a transition between quasi-static and rapid inertial shear, in 51

which vibration and momentum become important for the rheology of the flow. A 52

description of granular rheology in this transitional regime is therefore required for a full 53

understanding of the process of rupture nucleation and propagation. However, our 54

understanding of the rheology of dense, rapid granular flows is far from complete. 55

56

Some of the dificulty in describing the rheology of a dense, rapid granular flow arises 57

from the athermal nature of the granular medium. Unlike a traditional gas or liquid, a 58

granular material does not explore particle configurations and approach an equilibrium 59

state in the absence of an external energy source [Jaeger et al., 1996]. Consequently, the 60

rheology of a granular medium depends strongly on the detailed configuration of grains. 61

Under the same pressure or volume conditions, some grain (packing) configurations can 62

be very stiff, while others may be soft or ‘fluidized.’ A striking example of this 63

dependency on grain configuration is the phenomenon of jamming/unjamming [Liu and 64

Nagel, 2010] where a granular medium transitions dramatically from a solid-like to a 65

fluid-like phase, or vice-versa, under small perturbations to particle configuration or 66

loading direction [Aharonov and Sparks, 1999; Cates et al., 1998; Liu and Nagel, 1998]. 67

68

A good way to un-jam a granular material is to add vibration. This provides an external 69

energy source that allows a granular system to explore packing configurations. Vibration 70

usually leads to a more compact configuration, depending on the initial packing density 71

and the amplitude of the vibration [Knight et al., 1995; Nowak et al., 1997]. Another 72

source of external energy is imposed shear. Under shear, a granular medium may dilate 73

or compact, depending on its starting configuration. Starting from a dense state, such as 74

random close packing, it must dilate in order to allow geometrically frustrated grains to 75

move past each other (i.e. un-jam) [Lu et al., 2007; Marone, 1991; Reynolds, 1885]. 76

Acoustic vibration and imposed shear are thus two sources of external energy that tend to 77

drive the packing density of a granular system in opposite directions. 78

79

For gouge-filled faults, both of these energy sources may be present, with acoustic energy 80

generated during rupture, or arriving in seismic waves from nearby earthquakes. In 81

granular shear experiments, bursts of acoustic vibration can trigger stick-slip events or 82

generate lasting changes in rheology [Johnson et al., 2008]. Acoustic fluidization, in 83

which scattered wave energy produces transient reductions in fault normal stress, has 84

been proposed as a mechanism for reducing shear resistance during earthquake slip 85

[Melosh, 1996]. 86

87

Here we perform experiments that explore the behavior of a sheared granular medium 88

when the shearing itself generates acoustic vibration. Our starting point is a recent 89

experiment that showed that the dilatation of a steady, shear-driven flow is non-90

monotonic with respect to shear velocity [Lu et al., 2007]. That is, at velocities 91

intermediate between quasi-static and grain-inertial flow (the transitional regime, Fig 1c), 92

dilatation of the shear zone is markedly reduced. This behavior is not anticipated by 93

theoretical treatments of granular flow, which predict a monotonic increase of dilatation 94

with shear rate [Bagnold, 1954; Campbell, 1990; 2006; Clement, 1999] nor by dilatation 95

experiments carried out at lower shear rates [Marone et al., 1990; Marone, 1991]. 96

Numerical experiments using spherical or circular particles also fail to anticipate this 97

transitional compaction [da Cruz et al., 2005; GDR MiDi, 2004]. This phenomenon is 98

only observed for highly angular particles, and is negligible or non-existent for smooth 99

particles. 100

101

We propose in this paper that this intermediate shear-rate thinning can be explained as 102

auto-acoustic compaction. At moderate shear rates, grain interactions generate acoustic 103

vibration that causes compaction of shear-induced dilation. The tradeoff between shear-104

induced dilation and shear-induced acoustic compaction results in a shear-rate dependent 105

steady-state thickness of the shear zone. 106

107

In this study, we reproduce the observations of Lu et al., [2007], and compare the shear-108

induced compaction (weakening) at intermediate shear rates to the compaction of a 109

sheared granular layer under externally applied acoustic vibration. We first measure the 110

steady-state thickness of a sheared granular material over a range of shear rates. We take 111

the additional step of recording the amplitude of acoustic emissions over the range of 112

experimental velocities. We then apply similar amplitudes of acoustic vibration using an 113

external vibration transducer and characterize the effect on the steady-state thickness of 114

the actively shearing layer. We find that acoustic vibration reduces the steady-state 115

thickness of the shear zone in the same way, regardless of whether it is externally or 116

internally generated. We also perform acoustic pulse and transient shear-rate step 117

experiments to combine the time-dependent evolution of shear dilation and acoustic 118

compaction into a quantitative model of steady-state layer thickness. The experiments 119

motivate a scaling analysis that allows prediction to be made about behavior at depth. 120

121

2. Method 122

2.1 Experimental apparatus 123

The experimental apparatus is a TA Instruments AR2000ex commercial torsional 124

rheometer with rotating parallel plate geometry (Fig. 2). This instrument is capable of 125

sensitive measurement and control of torque, angular velocity, normal stress, and layer 126

thickness, and covers a large range of velocities (~10-5 to 300 rad/s), but is limited in the 127

magnitude of the applied forces (50 N normal force, 0.2 N-m torque). This is nearly 128

identical to the instrument used in Lu et al. [2007]. 129

130

The granular sample is housed in a quartz glass cylindrical jacket with dimensions 19 mm 131

diameter by 15 mm height (Fig. 2). A layer of angular sand grains is epoxied to the base 132

of the chamber and to the upper rotor, to force internal shear rather than slip at the 133

boundaries. The quartz cylinder is filled to ~12 mm depth, so that the rotor plate is 134

jacketed as well. Friction between the rotor and the quartz jacket is minimized by 135

carefully centering the sample. Since the primary observation of interest in these 136

experiments is the sample thickness, the friction between the rotor and jacket is relatively 137

unimportant. 138

139

2.2 Granular Media 140

We test two different granular materials with different grain shapes (Fig. 3), but 141

equivalent sizes and densities: spherical glass beads (mean diameter: 350 µm) and 142

angular beach sand (from Cowell’s Beach next to the Santa Cruz wharf) (range: 250-143

500 µm, mean: 350 µm). The beach sand is chosen for its high angularity and high 144

fracture resistance. The composition of the sand is roughly 44% quartz, 37% lithics, and 145

19% feldspars [Paull et al., 2005]. The composition of the heavy minerals and lithic 146

fragments is detailed by Hutton [1959]. 147

148

2.3 Acoustic Vibration 149

External acoustic vibration is produced by a ceramic transducer (PZT-5) affixed to the 150

base plate of the rheometer, adjacent to the sample chamber. An accelerometer (Bruel 151

and Kjaer 4373 charge accelerometer) attached directly to the cylindrical jacket allows 152

measurement of the acoustic acceleration within the sample, from which the strain 153

amplitude can be calculated. The accelerometer has a flat response up to 35 kHz. In 154

order to achieve high strain amplitude with a small source transducer, we vibrate at the 155

resonant frequency of the mechanical system (found to be 40.2 kHz). Note that this is the 156

natural frequency of the entire apparatus, not the sample chamber itself. Detailed studies 157

of the shear-generated acoustic spectrum and the frequency dependence of the sample 158

response are beyond the scope of this study. 159

160

2.4 Experimental Procedure 1: Velocity ramps 161

The primary experiment consists of a suite of velocity steps between an angular velocity 162

of 10-3 to 100 rad/s (9.5×10-4 to 95 cm/s at the outer rim of the rotor) under controlled 163

normal stress, and we report sample thickness. The velocity is incremented gradually 164

from slow to fast and fast to slow, multiple times per sample. The repeated velocity 165

ramps extend the work of Lu et al. [2007], who focused mostly on a single velocity ramp 166

per sample. Each velocity step lasts 20 seconds, and the reported thickness is averaged 167

over the last 10 seconds of each step. The duration of the steps is chosen to allow the 168

shear zone to evolve to a new steady-state thickness value after the small step in velocity. 169

Using a longer step duration does not change the results. Throughout this paper, we refer 170

to this experimental procedure as a velocity ramp, although in reality it consists of small 171

discrete velocity steps, run to steady-state. 172

173

We first run the velocity ramp in the absence of acoustic vibration to establish the 174

baseline steady-state thickness of the shear zone as a function of shear rate. During these 175

experiments, we record the amplitude of acoustic vibration produced internally by grain 176

interactions during shearing. We then perform velocity ramps in the presence of constant 177

amplitude external vibration of similar amplitude and examine the effect on the steady-178

state thickness of the shearing sample. 179

180

2.5 Experimental Procedure 2: Acoustic pulses and shear-rate steps 181

In the second experiment we subject the sample to acoustic pulses and sudden velocity 182

steps. We fit the resulting compaction as a function of time, and use this equation to 183

match the steady-state thickness observations from the velocity ramp experiments. We 184

also compare the magnitude of compaction produced by similar amplitude external and 185

shear-induced acoustic vibration. 186

187

In the acoustic pulse experiment we shear the sample at a slow, constant rate of 0.01 rad/s 188

to minimize internally generated vibration and shear dilatation, and then introduce 60-189

second pulses of acoustic vibration at various amplitudes. In the shear-rate step 190

experiment, the sample is sheared at a constant rate of 0.1 rad/s, and then subjected to 60-191

second jumps in shear rate. The amplitude of the shear-induced acoustic vibration is 192

captured by the accelerometer, allowing us to compare the magnitude of compaction 193

under equivalent-amplitude external and shear-induced acoustic vibration. 194

195

The velocity ramp experiments are summarized in Table 1, and the pulse experiments are 196

summarized in Table 2. 197

198

2.6 Boundary conditions and relation to controlled volume experiments 199

All experiments are carried out under controlled normal stress conditions. Normal stress 200

is maintained by the rheometer controller software through adjustment of the sample 201

thickness. These experiments explore the non-fracture deformation regime (in which 202

strain is accommodated primarily through grain rolling and sliding, rather than 203

cataclasis), over a range of velocities that includes both quasi-static behavior and inertial 204

behavior where collisions between grains become important. To reach the inertial 205

granular flow regime with this apparatus, the normal stress must be limited to < 10 kPa. 206

207

We employ constant normal stress rather than constant volume boundary conditions 208

because the large variations in shear zone thickness would otherwise result in either 209

decoupling of the sample during compaction phases, or locking up of the mechanical 210

drive during dilatation. However, qualitatively similar results are obtained when 211

measuring the evolution of normal stress under constant volume conditions [Lu et al., 212

2007]. 213

214

The correspondence between normal stress and layer thickness for alternate boundary 215

conditions is a consequence of the cyclic rule, which states the relationship between the 216

partial derivatives of a three variable system 217

∂V∂ γ

⎛⎝⎜

⎞⎠⎟ σ

= −∂V∂σ

⎛⎝⎜

⎞⎠⎟ γ

dσd γ

⎛⎝⎜

⎞⎠⎟V

. (1) 218

The change in volume or layer thickness is related to the change in stress with shear rate 219

through the compressibility term ∂V ∂σ( ) γ . Layer compaction under constant load thus 220

implies normal stress reduction under constant volume [Lu et al., 2007]. 221

222

3. Results 223

3.1. Velocity ramps – shear induced compaction 224

In the absence of acoustic vibration, the sample thickness is a well-defined, reproducible 225

function of shear rate (Fig. 4), regardless of whether the velocity ramp is increasing or 226

decreasing. Both angular sand and smooth glass beads deform with a thickness that is 227

independent of shear-rate at very low shear rates (10-3 - 10-1 cm/s), and show strong 228

dilation at very high shear rates (>10 cm/s). However, the behavior of the two granular 229

media differs greatly at intermediate shear rates (0.1-7 cm/s), where the thickness of the 230

angular sand decreases markedly with shear rate. The glass beads show only a hint of 231

this thinning at intermediate shear-rates. 232

233

The reversible behavior shown in Figure 4 follows an initial run-in phase that consists of 234

a relatively gradual compaction trend that diminishes with total displacement, and 235

essentially vanishes over a few velocity ramp cycles. Most of the experiments are 236

reported after this run-in phase is complete, but in one velocity ramp experiment we 237

subtract a linear trend with displacement to remove the tail end of the run-in phase. The 238

velocity ramp experiments are summarized in Table 1. In the following, we focus 239

exclusively on the reversible and repeatable component of compaction, which we 240

interpret as reflecting thickness changes in the actively shearing layer (Appendix A). 241

242

The dilatational behavior of the granular media at either end of the velocity range can be 243

understood in terms of the end-members of granular flow (Fig. 1). In granular physics, 244

the flow regime is thought to be determined by the dimensionless inertial number, which 245

compares the magnitude of the grain inertial stress to the confining stress [Bagnold, 1954; 246

Bocquet et al., 2001; Campbell, 2006; Iverson, 1997; Savage, 1984]. As discussed 247

above, at low shear rates grain inertial stresses are very small, and stresses are carried 248

frictionally or elastically along force chains/networks. In this regime, shear and normal 249

stresses and sample thickness are independent of shear rate [Thompson and Grest, 1991]. 250

At high shear rates grain inertial stresses are of similar order as the confining stress, 251

supporting dilation of the shear zone. The thickness of the sample is then strongly 252

dependent on the shear rate. 253

254

The glass beads show a relatively monotonic transition between the regimes, where the 255

thickness of the sample is constant in the quasi-static regime, and then increases as the 256

sample enters the grain inertial regime (Fig. 4). For the angular sand, complex behavior 257

is observed that is not anticipated by either end-member granular flow regime. At 258

intermediate shear rates there is a robust reduction in steady-state layer thickness, 259

reproducing the findings of Lu et al., [2007]. This transitional compaction, on the order 260

of 50 µm at maximum, occurs regardless of the direction of the velocity ramp. The 261

active shear zone thickness is on the order of a few grains, with a scale depth of ~600 µm 262

(Appendix A). The maximum compaction is therefore on the order of 10% of the shear 263

zone thickness. 264

265

3.2 Velocity ramps with acoustic vibration 266

We next perform velocity ramps in the presence of external vibration (Fig. 5). The 267

introduction of acoustic vibration has two effects. First, it causes additional irreversible 268

compaction that decays linearly with log time from the start of the vibration, similar to 269

the run-in phase in the absence of vibration (Fig. 6). Second, it induces a strong 270

dependence between steady state thickness and shear rate at low shear rates that was not 271

seen in the previous experiment. This low shear rate behavior is seen for the glass beads 272

as well as the angular sand (Fig. 5b). At low shear rates, the layer thickness is 273

significantly reduced compared to the non-vibrated case. This reduced thickness is again 274

a steady-state value, and is reproducible regardless of whether the velocity ramp is 275

increasing or decreasing. The layer thickness increases with shear rate, up to the 276

transitional regime for angular sand (Fig. 5a). At higher shear rates, there is negligible 277

difference between the experiments with and without external acoustic vibration. The 278

vibrational compaction of the sample at low shear rates is reminiscent of acoustic 279

vibration experiments performed in the absence of shear, in which vibration is observed 280

to induce rapid compaction that is recoverable over long time scales [Johnson and Jia, 281

2005]. 282

283

A representative acoustic experiment (VRS3, Table 1) is shown as a function of 284

experimental time in Figure 6, to give a better sense of the total evolution of sample 285

thickness over the course of multiple velocity ramps. This illustrates the irreversible 286

compaction during the acoustic vibration, and shows that the additional dip at low shear 287

rates in the vibration experiment is not an artifact of removing the irreversible trend. For 288

the example in Figure 6, the sample is first sheared through several velocity ramps in the 289

absence of acoustic vibration to establish the baseline shear-rate dependent behavior. 290

The pattern of thickness vs. shear rate (c.f. Figure 5) is traced and then retraced in reverse 291

as velocity is varied over experimental time. Acoustic vibration is then switched on and 292

the sample is run through several more velocity ramp cycles. Finally, vibration is 293

switched off and the sample is allowed to recover back to its original thickness vs. shear-294

rate behavior. 295

296

For the constant normal load boundary conditions used here, this compaction is not 297

accompanied by a shear stress decrease, i.e. the coefficient of friction remains the same. 298

If the experiment were carried out at constant volume, this compaction would correspond 299

to a reduction in both normal and shear stress. 300

301

3.3 Linking compaction and acoustic vibration 302

The acoustic vibration recorded by the accelerometer (Fig. 7) reveals a qualitative link 303

between the compaction of the angular sand in the transitional regime and the compaction 304

due to external acoustic vibration. Vibration of the sample chamber is recorded both 305

during external excitation and for shearing without external vibration. Peak acoustic 306

amplitude is measured for 1-second intervals, and then averaged over the duration of each 307

20-second velocity step. Peak acoustic strain amplitude ε is estimated for the externally 308

induced vibration by dividing the peak acceleration a by the frequency of the vibration (f 309

= 40.2 kHz) to get particle velocity, and then dividing by the acoustic wave speed c 310

ε ≈ vc≈afc

. (2) 311

The acoustic wavespeed for a solid is given by 312

c =2G 1−ν( )ρ 1− 2ν( ) , (3) 313

where ρ is density, G is the shear modulus, and ν is Poisson’s ratio. The shear modulus 314

of our sample is 1.1 × 108, as measured by an oscillatory strain test with frequency 100 315

Hz. The sample density is 1500 kg/m3. Assuming a Poisson’s ratio between 0 and 0.25, 316

this gives an acoustic wavespeed c between 390 and 480 m/s. This agrees well with other 317

experiments on acoustic travel time in granular media, extrapolated to slightly lower 318

pressures [Coghill and Giang, 2011; Jia et al., 1999], as well as with theoretical 319

predictions using effective medium theory [Makse et al., 2004; Walton, 1987]. 320

321

Figure 7 shows several suggestive features linking steady-state compaction to acoustic 322

amplitude. First, the beginning of the transitional compaction for angular sand begins 323

when internally produced acoustic vibration becomes detectable. The amplitude for glass 324

beads is smaller at this shear rate, consistent with the much smaller transitional thinning. 325

Second, at high shear rates, the internally generated vibration overwhelms the external 326

vibration such that there is no difference in the recorded vibration amplitude with or 327

without external vibration. This corresponds to the joining of the thickness vs. velocity 328

curves from the two experiments (Fig. 5). These observations suggest that acoustic 329

vibration produces compaction in a similar way regardless of whether the vibration is 330

externally or internally generated. 331

332

The rollover in amplitude at higher shear rates (Fig. 7) is due to the fact that the recorded 333

peak amplitude begins to clip at higher shear rates. To get a more robust estimate of the 334

scaling of vibration amplitude with shear rate, we also compute the signal power of the 335

shear-induced vibration signal for angular sand (Fig. 8). Power is computed from 336

periodograms in a pass band between 1 and 35 kHz. This pass band avoids instrument 337

and electronic noise at low frequencies and resonance peaks at higher frequencies (where 338

the amplitude clips). The signal power is defined as the integral of the spectrum within 339

the pass band [Stearns, 2003]. Shear-induced acoustic power increases linearly with 340

shear rate (Fig. 8). We will see in a subsequent section that the compaction magnitude 341

also increases linearly with shear rate, indicating a correlation between acoustic power 342

and compaction. 343

344

3.4 A qualitative model for shear zone thickness 345

These observations suggest a qualitative model for the steady state thickness of a sheared 346

layer in the presence of acoustic vibration. Both acoustic vibration and shear 347

displacement provide external energy and drive the system to explore packing densities, 348

but in opposite directions and with different timescales. Acoustic vibration produces 349

compaction in the shear zone at a rate proportional to the dilation beyond random close 350

packing, while shear deformation continually renews contacts and re-dilates the sample at 351

a rate proportional to the compaction below the critical dilatancy required for shearing. 352

At some thickness, these two mechanisms are balanced, and a steady state is reached. 353

354

These ingredients qualitatively explain the shear-rate dependent behavior seen in Figure 355

5. At low shear rates, dilation is slow, and acoustic compaction has a relatively long time 356

period over which to operate. At somewhat higher shear rates, dilation becomes more 357

rapid, and steady-state compaction is reduced. At even higher shear rates, additional 358

acoustic energy is produced by the shearing itself (Fig. 7), driving renewed acoustic 359

compaction of the shearing layer. This is most pronounced for angular sand grains, 360

which generate high amplitude acoustic vibrations during shear, but is also noticeable for 361

the quieter smooth glass beads. Eventually, grain inertial effects dominate the dilatation, 362

and shear dilatancy and acoustic compaction become irrelevant. 363

364

To develop a more quantitative model of the steady-state shear zone thickness, we now 365

investigate the time dependent evolution of thickness in the presence of externally 366

applied acoustic vibration and internally generated vibration arising from shear-rate 367

pulses. This investigation will constrain the rate of acoustic compaction. Combined with 368

a measurement of the rate of shear dilatation, a prediction can be made for the steady-369

state thickness as a function of shear rate. 370

371

3.5 Acoustic pulse experiments 372

In the externally applied vibration pulse experiments, the angular sand is sheared at a 373

constant 0.01 rad/s and subjected to 60-second bursts of acoustic vibration, after which it 374

is allowed to recover for up to 10 minutes (Fig. 9). The bursts are repeated 5-10 times for 375

each acoustic amplitude in order to characterize precision. The experiments are 376

summarized in Table 2. The pulses of acoustic vibration produce compaction in the 377

angular sand that increases with the amplitude of the pulse. After cessation of the 378

acoustic burst, the sample recovers a significant component of the thickness that was lost 379

during the acoustic burst, over ~100 seconds. 380

381

The evolution of thickness h during compaction over the acoustic burst is fit with a 382

logarithmic function of time t 383

h t( ) = h0 − b logtτ − +1⎛

⎝⎜⎞⎠⎟ , (4) 384

where h0 is the starting thickness, and b and τ − are empirical constants that describe the 385

magnitude and timescale of compaction, respectively. This function is motivated by 386

other work where the density (i.e. inverse of volume) of a vibrated granular medium is 387

observed to evolve as ~1/log(t) [Ben-Naim et al., 1998; Caglioti et al., 1997; Knight et 388

al., 1995], however other functions (e.g. stretched exponential, double exponential) may 389

fit the data just as well. In this study, we are primarily interested in whether some 390

function that describes the evolution of thickness with time can also predict the steady-391

state observations as a function of shear rate, and a theoretical treatment of the 392

micromechanics of acoustic compaction are beyond the scope of this experimental work. 393

394

We start by fitting the curves for each vibration amplitude individually, treating both b 395

and τ − as unknowns. We then fix τ − to equal the average of the values for all the curves 396

(10 s), and refit the constant b (Table 2). This allows us to map all of the variation with 397

amplitude into a single constant. While it is possible to fit the curves with a single value 398

of τ − and varying b, it is not possible to fit the curves with a single value of b and 399

varying τ − . This implies that the magnitude of compaction is strongly dependent on 400

vibration amplitude, but the timescale is not. 401

402

For the highest amplitude vibration, the acoustic compaction produces a very small rapid 403

drop in the normal stress, before the rheometer recovers by lowering the upper plate (Fig. 404

9). This reflects the fundamental relationship between sample thickness and stress in 405

these experiments (Eq. 1). Since the rheometer is unable to keep up with the compaction 406

rate (Appendix B), the time scale τ − that we measure with the rheometer must be treated 407

as an upper bound. We will eventually solve for this timescale more accurately by fitting 408

the steady-state velocity ramp experiments. This effect is relatively negligible for the 409

small-amplitude external vibration experiments, but will be considerably larger for the 410

shear-rate pulses examined in the next section. 411

412

413

3.6 Shear rate pulses 414

We also measure compaction during transient increases in shear rate (Fig. 10). The 415

response to a shear-rate step is more difficult to interpret than the response to an acoustic 416

burst, because there are several competing effects that offset vibrational compaction, 417

including the direct dilational response to a velocity step [Marone et al., 1990; Marone, 418

1991] and grain inertial effects [Bagnold, 1954]. These effects were not an issue for the 419

acoustic vibration pulses, because the shear rate was small and identical for all pulses. 420

We shear the sample at a starting rate of 0.1 rad/s – the maximum shear rate before the 421

transitional weakening – to minimize the direct effect of the velocity jump. We then 422

apply high shear-rate pulses of up to 5 rad/s – the maximum shear rate before inertial 423

dilation effects become apparent in Figure 5. Each curve in Figure 10 represents an 424

average over 5 - 10 runs (see Experiment SPS1, Table 2). The compaction magnitude is 425

observed to increase with the magnitude of the shear rate. When shear rate is stepped 426

back down to 0.1 rad/s, the sample recovers a significant fraction of the shear-induced 427

compaction over a timescale of a hundred seconds, just as it did after the acoustic pulses. 428

429

We fit the shear-induced compaction curves with the same function as for the acoustic 430

pulses (Eq. 4), again with τ − = 10 seconds, and again with the caveat that this timescale 431

is limited by the instrument response (Appendix B). The compaction magnitude 432

parameter b is listed in Table 2, along with the amplitude of internal, shear-generated 433

vibration as measured by the accelerometer. 434

435

We also fit the dilatational evolution in the recovery phase with an exponential function, 436

h t( ) = h0 − Δhexp −tτ +

⎛⎝⎜

⎞⎠⎟ , (5) 437

Where h0 is the steady-state layer thickness after complete recovery (i.e. the critical 438

thickness for shearing in the absence of vibration), Δh is the total drop during the 439

compaction phase, and τ + is the timescale for re-dilation. We find an average recovery 440

timescale τ + of 55 seconds, corresponding to a displacement scale of 5.5 rad, given a 441

shear rate of 0.1 rad/s. 442

443

There is strong rapid normal stress drop during the shear-rate jumps due to the slow 444

response of the rheometer (Appendix B), and for the largest amplitudes, the rheometer 445

does not completely catch up to restore the normal force even after 60 seconds (Fig. 10). 446

The underlying compaction timescale, as suggested by the rapid normal stress drop, is 447

therefore likely to be less than the 10 seconds estimated from the thickness curves, and 448

may be as little as 1 second or less. 449

450

Plotting the values of compaction magnitude vs. acoustic strain amplitude for both shear 451

and acoustic pulses (from Table 2), we find that, for both pulse types, the compaction 452

magnitude is comparable for the same amplitude vibration (Fig. 11). This indicates that 453

the acoustic compaction mechanism is activated by vibration in the same way regardless 454

of whether the acoustic energy is generated by grain interactions during shear, or injected 455

from an external source. The recovery of dilatation after the cessation of the pulse shows 456

that the acoustic vibration is causing compaction within the actively shearing layer, 457

essentially suppressing shear dilatation and allowing the grains to shear in a more 458

compact configuration at the same stress. 459

460

We infer that the compaction observed in the transitional regime during the velocity ramp 461

experiments (Fig. 5) is also the result of internally-generated acoustic vibration that feeds 462

back on the rheology of the medium. The empirical compaction and dilatation functions 463

that we have measured (Eqs. 4,5) now allow us to formulate a quantitative model for the 464

steady-state thickness of a sheared layer that reflects the competition between acoustic 465

compaction and shear dilation. 466

467

4. Discussion 468

4.1 Key components of the steady-state thickness model 469

Any model that explains the observations presented in Figure 5 must meet at least the 470

following fundamental criteria: 1) At low shear rates, the dilation rate must scale with 471

shear rate in order for external acoustic vibration to produce more steady-state 472

compaction at lower velocities. 2) For angular sand, the shear-induced compaction rate 473

must increase faster with shear rate than the shear-induced dilation rate, in order to 474

produce the transitional compaction regime. 3) The external and internally generated 475

acoustic compaction should behave similarly for equivalent amplitude vibrations. 476

477

4.2 Competition between shear-dilatancy and auto-acoustic compaction 478

We propose a basic model for steady-state shear zone thickness that represents a balance 479

between shear-induced dilation and vibration-induced compaction. The evolution of 480

dilation follows an exponential function of time (Eq. 5), as observed in the recovery stage 481

of Figure 10 after the shear-rate pulse. This implies a dilation rate 482

dhdt

+

=h0 − h( )τ + , (6) 483

where τ + is the timescale for dilation, and h0 is the steady-state thickness in the absence 484

of vibration. We further assume that τ + is inversely proportional to the shear rate ω, 485

τ + =dω

, (7) 486

where d is an angular displacement length scale for renewal of force chains. 487

488

The evolution of compaction follows a logarithmic function of time (Eq. 4). This implies 489

a compaction rate 490

dhdt

= −bτ − exp −

h0 − hb

⎛⎝⎜

⎞⎠⎟ , (8) 491

where τ − is a constant that does not depend on vibration amplitude, and b is some 492

function of the amplitude of the acoustic vibration. The steady state thickness of the 493

shear zone is the thickness h at which the dilation rate (Eq. 6) and compaction rate (Eq. 8) 494

are balanced, i.e. 495

dhdt

+

= −dhdt

, (9) 496

Substituting Equation 7 into Equation 6, Equation 9 gives 497

h − h0d

ω =bτ − exp

h − h0b

⎛⎝⎜

⎞⎠⎟ , (10) 498

where, again the constant b depends on the internally or externally generated vibration 499

amplitude. 500

501

4.3 Experimental fit 502

We now return to the steady-state thickness (velocity ramp) experiments (Fig. 5), and use 503

Equation 10 to solve for b as a function of shear rate. We use the same equation to solve 504

independently for b in the case with and without external vibration, and refer to the fit 505

parameters as bacoustic and bshear, respectively. 506

507

In order to constrain the compaction magnitude parameter b (Eq. 10) we use the 508

dilatational recovery length scale d = 5.5 radians, estimated from the shear-rate pulse 509

experiments, and the compaction timescale τ − , from the acoustic pulse experiments 510

(Table 3). As previously discussed, the timescale τ − is poorly constrained by the pulse 511

experiments because of the finite response time of the rheometer (Appendix B). We 512

instead choose a value of τ − that gives a constant compaction magnitude bacoustic over the 513

range of shear rates where the contribution of shear-induced vibration is negligible (Fig. 514

12). This requires a compaction timescale τ − = 1 second. 515

516

The steady-state layer thickness in the presence of vibration can be explained quite well 517

by a constant compaction magnitude bacoustic at low shear rates (Fig. 12). The systematic 518

relationship between steady-state thickness and shear rate simply reflects the longer 519

timescale for shear dilatation at low shear rates τ + = d ω( ) . At low shear rates, the 520

acoustic compaction has a relatively long time to operate. As the shear rate increases, 521

less and less compaction can occur within the timescale for dilatation. 522

523

In the absence of external vibration (shear-induced vibration only) the compaction 524

magnitude bshear increases linearly with shear-rate (Fig. 12), consistent with the linear 525

increase in shear-induced acoustic power observed earlier (Fig. 8). A possible 526

interpretation of the linear increase with shear rate would be that each discrete acoustic 527

emission event has a constant amplitude, independent of shear rate, and the rate of 528

acoustic emission events increases proportional to shear rate. Compaction magnitude 529

continues to increase until the inertial granular flow regime is reached and inertial 530

dilation overwhelms vibration-induced compaction. Note that this transition is reached at 531

much lower shear rates for the quieter glass beads. 532

533

We plot the observed and modeled (Equation 10) steady-state thickness vs. shear-rate in 534

Figure 13. The shear-induced compaction magnitude (bshear = 18 µm (rad/s)-1) is found by 535

fitting a straight line to the shear-induced compaction curve in Figure 12. The external 536

acoustic compaction magnitude (bacoustic = 10 µm) is found by fitting a horizontal line to 537

the low shear rate, acoustic compaction curve in Figure 12. The steady-state compaction 538

in the presence of both shear-induced and external acoustic compaction is very nearly 539

predicted by a linear sum between the acoustic and shear-induced compaction 540

components (Fig. 13), i.e. 541

btotal = bacoustic + bshear . (11) 542

We conclude that the mechanism for acoustic compaction of the actively shearing layer is 543

identical, regardless of whether the vibration is generated internally or externally. 544

545

4.4 The quasi-static – acoustic transition: a new granular flow regime 546

In shear flows, the transition to dispersive inertial flow occurs when the dispersive 547

pressure ρv2 equals the confining pressure p [Bocquet et al., 2001; Clement, 1999; 548

Iverson, 1997; Lu et al., 2007; Savage, 1984] 549

p = ρv2 , (12) 550

where ρ is density, and v is the shear rate. From Equation 12, we can define a 551

dimensionless number I such that 552

I ≡ ρv2

p. (13) 553

The dimensionless number I (or its square root) is the commonly used inertial number for 554

granular flow [Jop et al., 2006; GDR MiDi, 2004; Savage, 1984]. For comparison with 555

other references, note that the inertial number is often defined as I ≡ ρd 2 γ 2 p . In our 556

study, we find that the velocity profile is exponential with a scale depth on the order of a 557

grain diameter (Appendix A), such that v ~ γ d , leading to Equation 13. An inertial 558

number I << 1 implies quasi-static flow, and I >> 1 implies dispersive inertial flow. The 559

transition from a static regime in which stresses are supported elastically to a dispersive 560

regime where stresses are supported by inertial collisions is inherently dilatational 561

[Bagnold, 1954] 562

563

The maximum shear rate in our experiments is 100 cm/s, corresponding to an inertial 564

number of 0.3. Consequently, we never reach the fully inertial regime where stresses and 565

dilatation should scale as the square of shear rate [Bagnold, 1954]. Nevertheless, we 566

observe dilatation that exceeds quasi-static values at shear rates above ~30 cm/s, 567

corresponding to a dispersive pressure ~200 Pa (ρ = 2x103 kg/m3), which is only about 568

3% of the normal pressure p = 7 kPa. 569

570

The transition to the acoustic regime is observed at shear rates as low as 1 mm/s. The 571

dispersive pressure at this velocity is < 10-6 times the confining pressure, and the 572

compaction cannot be attributed to inertial behavior. Instead, this low-velocity transition 573

represents the transition to acoustic fluidization. The acoustic pressure pa equals the bulk 574

modulus (K=ρc2) times the acoustic strain ε, giving 575

pa = ρc2ε , (14) 576

where ρ is the density of the medium, c is the acoustic wave speed [Thompson, 1971, 577

chap. 4]. We experimentally identify acoustic compaction at a threshold strain ε on the 578

order of 10-5 (Fig. 9). Taking the acoustic velocity c ~ 500 m/s, and ρ = 2000 kg/m3, the 579

threshold acoustic pressure (Eq. 13) is 5 kPa, on the order of the confining stress (7 kPa). 580

This suggests another non-dimensional number 581

J ≡ ρc2εp

, (15) 582

The transition from quasi-static to acoustic granular flow occurs when confining pressure 583

is balanced by acoustic pressure, i.e. J = 1. Note that the acoustic approximation 584

becomes invalid as the acoustic pressure fluctuations approach the absolute pressure, i.e. 585

J ~ 1, and the acoustic wave speed should decrease dramatically in the fluidized regime, 586

limiting further increases in the acoustic pressure. 587

588

Melosh [1979; 1996] suggested that acoustic fluidization could allow granular materials 589

to flow at shear stresses far below the frictional strength suggested by the overburden 590

pressure. The energy density to “fluidize” a rock mass by acoustic vibration is orders of 591

magnitude smaller than the energy density required to fluidize the same rock mass by 592

kinetic particle motion [Melosh, 1979]. In this theory, acoustic pressure fluctuations 593

produce transient reductions of the normal stress, allowing stress to locally exceed the 594

coulomb frictional threshold. The mechanism is activated when the peaks in acoustic 595

pressure reach the order of the overburden. 596

597

The mechanism we propose here is very similar to the acoustic fluidization of Melosh, 598

but differs in a fundamental way: instead of normal force being reduced directly by 599

acoustic stress fluctuations, it is reduced indirectly through the phenomenon of acoustic 600

compaction. The general conceptual model for acoustic compaction comes out of 601

granular physics [Liu and Nagel, 2001; Liu and Nagel, 2010; Mehta, 2007]. In this 602

framework, shear stress in a granular packing is supported by a framework of force 603

chains, supported by a network of buttressing grains [Majmudar and Behringer, 2005]. 604

Increased loading in the shear direction primarily compresses and rotates the strong force 605

chains, which leads to bulk dilatation [Tordesillas et al., 2011]. Particle reconfiguration 606

(compaction) can only occur when strongly loaded force chains buckle catastrophically. 607

When static equilibrium is restored by the formation of new force chains, the packing is 608

again in a jammed state and unable to explore configurations. Acoustic vibration, on the 609

other hand, accesses both strong force chain grains and buttressing grains directly, and 610

may disrupt grain contacts without catastrophic buckling of force chains. This promotes 611

incremental compaction into a lower energy configuration. 612

613

Auto-acoustic fluidization, as observed in these experiments, is related to the amplitude 614

of acoustic vibration produced through grain interactions during shear. As such, it is 615

strongly dependent on the characteristics of the grains. Angular grains generate sufficient 616

acoustic energy to strongly affect the rheology of the flow, but smooth glass beads do 617

not. This implies that the physical characteristics of gouge particles observed in fault 618

zones can tell us something about the rheology of the flow during rupture. 619

620

For this mechanism to be active at seismogenic depths of ~10 km, with overburden 621

pressure on the order of 3×108 Pa, assuming c = 3x103 m/s, and ρ = 3x103 kg/m3, 622

Equation 14 requires that the acoustic strain amplitude ε be on the order of 10-2. It is not 623

known how internally generated acoustic strain amplitude should scale with slip rate at 624

these conditions, and this value is at the upper limit of plausible elastic strain in rock. 625

However, if the effective confining pressure is reduced by fluid pressure or some other 626

mechanism, the required acoustic strain amplitude will be reduced accordingly. There is 627

considerable evidence supporting the idea that effective pressures may indeed be low in 628

many faults [Hickman et al., 1995; Sleep and Blanpied, 1992]. 629

630

Future experiments must 1) establish what range of grain charateristics (angularity, aspect 631

ratio, etc.) are capable of generating sufficient acoustic vibration to feed back on the 632

rheology of the shear flow, and 2) quantify how the amplitude of acoustic vibration scales 633

with shear rate at seismically relevant confining stresses. 634

635

4.5 The acoustic – inertial transition 636

The acoustic transition defines the point at which the active shear zone compacts relative 637

to the quasi-static thickness. This compaction grows larger until the inertial regime is 638

encountered. We define the acoustic-inertial transition as the shear rate at which the 639

dispersive pressure of granular collisions exceeds the acoustic pressure, that is, when the 640

non-dimensional number K > 1, where 641

K ≡ρv2

ρc2ε=IJ

. (16) 642

Assuming the maximum acoustic strain ε = 1 (where the wave particle velocity equals the 643

wave speed and the acoustic approximation breaks down), Equation 16 reduces to a ratio 644

of velocities that can be thought of as the timescale for the transmission of stresses over 645

the dimensions of a grain: the acoustic timescale ∝c-1 and the shear time scale ∝v-1 646

K * =v2

c2. (17) 647

Equation 17 is the square of the Mach number. Substituting the approximate wave speed 648

equation for a fluid c2 = p/ρ [Thompson, 1971], Equation 17 becomes 649

K * =ρv2

p, (18) 650

which is just the inertial number (Eq. 13). The transition to the inertial regime therefore 651

occurs at the same shear rate, regardless of whether the transition is quasi-static to 652

inertial, or acoustic to inertial. The acoustic regime therefore substitutes for a portion of 653

the quasi-static regime, but does not inhibit the transition to grain-inertial flow. 654

655

The apparent equivalence between Equations 17 and 18 emphasizes the fundamental 656

distinction between the quasi-static regime and the acoustic regime. The same elastic 657

time scale governs both the acoustic regime and the quasi-static regime. However, in the 658

quasi-static regime, the elastic stresses are too small to have a dynamic effect on the 659

particles, and c-1 controls the time scale for equilibration of static forces. In the acoustic 660

regime, transient elastic stresses are large enough to induce inter-particle slip, and c-1 661

controls the timescale over which perturbations in the packing configuration are 662

communicated throughout the medium. In the inertial regime, on the other hand, neither 663

stress equilibration nor perturbations to the particle packing can be achieved before the 664

configuration is renewed by shear. 665

666

5. Conclusion 667

We have experimentally measured the dilatational behavior of angular granular media 668

over a range of geophysically relevant shear rates. We have quantified the physical 669

conditions for a new granular flow regime at shear rates transitional between quasi-static 670

and inertial granular flow, that we term the acoustic regime. In this regime, internally 671

generated vibrations induce auto-acoustic compaction of the actively dilating shear zone. 672

The steady-state thickness of the shear zone is described by a quasi-empirical model that 673

balances auto-acoustic compaction and shear dilatation. 674

675

Experiments using external acoustic vibration reveal the fundamental link between the 676

intermediate shear-rate compaction and acoustic compaction, showing that the same 677

magnitude of steady-state compaction is achieved for the same magnitude of acoustic 678

vibration, regardless of whether it is applied externally or generated internally by 679

shearing and grain collisions. The magnitude of internally generated acoustic vibration is 680

dependent on the characteristics of the grains, and thus the phenomenon of auto-acoustic 681

fludization occurs for angular grains, but not for quieter smooth grains. 682

683

The acoustic regime is activated when peak acoustic pressure from shear-induced 684

vibration exceeds the confining stress. At 10 kPa confining stress, this transition occurs 685

at a shear rate of only 1 mm/s in angular sand – four orders of magnitude below shear 686

rates typical of the transition to inertial, dispersive granular glow. Scaling the process up 687

to seismogenic conditions suggests that this mechanism may be an important velocity 688

weakening mechanism as long as confining stresses during earthquake rupture are sub-689

lithostatic. 690

691

The velocity weakening acoustic rheology demonstrated here has significant implications 692

for coseismic stress reduction on earthquake faults, which generate high frequency 693

shaking both in the wake and in advance of a propagating rupture front, as well as for the 694

susceptibility of gouge-filled faults to triggering by high-frequency transient stresses. 695

696

Appendix A: Shear zone thickness 697

The shear zone in each experiment is restricted to a thin layer near the upper rotor, with 698

displacement and shear rate decreasing exponentially with distance from the rotor (Fig. 699

A1). The exponential displacement profile is constant regardless of the shear rate, that is, 700

the velocity profile changes only through a pre-factor proportional to the rotor velocity. 701

702

We image the active shear zone by cross-correlating video snapshots of the outer 703

boundary using a high-definition webcam. Snapshots are taken at rim displacement 704

intervals of 10 cm for three representative velocities: 0.1, 1, and 10 cm/s. We then cross-705

correlate successive images to estimate horizontal lag as a function of depth. This 706

involves taking a horizontal row of pixels from each pair of images and calculating the 707

cross correlation function between the two. The lag corresponding to the maximum in 708

the correlation function is taken as the horizontal displacement. We repeat this process 709

for each row of pixels to get displacement as a function of depth away from the rotor. 710

We then average the cross correlation function over 100 image pairs to produce the 711

images in Figure A1. 712

713

The shear rate profile with depth is well fit by an exponential function with a decay 714

length scale of 0.63 mm, or approximately 2 grain diameters (Fig. A1). We fit an 715

exponential function to the peaks in the cross-correlation image for the shear rate of 0.1 716

cm/s, and then superpose this fit to the second two cross correlation images. No 717

adjustment in the fit parameters is required to fit the profile at 1 cm/s and 10 cm/s. This 718

is consistent with the results of Lu et al., [2007]. 719

720

Appendix B: Rheometer response to a sudden pressure drop. 721

The AR2000ex rheometer maintains normal pressure through a firmware feedback loop, 722

adjusting the height of the rotor (gap) until the measured force falls within the bounds set 723

by the controller software. The algorithm for adjusting the gap in response to a sudden 724

drop in normal stress is not documented in the rheometer software, and thus the 725

instrument response time after a sudden compaction event is unknown. We perform a 726

simple experiment in which we observe the rheometer response to a sudden compaction 727

event (drop in pressure). 728

729

The experiment proceeds as follows: 1) We place an air bladder (1 qt. Ziploc-style bag) 730

on the rheometer platform (Figure B1). 2) A corner of the bag is pinched to reduce the 731

internal volume and increase the pressure. 3) The rotor is lowered until it compresses the 732

bag and a normal force of 1 N is registered by the rheometer. 4) The constriction on the 733

bag is then suddenly released, resulting in a rapid drop in pressure. We then observe the 734

change in gap (rotor height) with time (Figure B2, inset). The experiment is repeated for 735

pressure drops of ~0.5 and ~1 N, as well as for continuous pressure fluctuations induced 736

by intermittently squeezing and releasing the bag. 737

738

We find that the rheometer has a significantly delayed response to a sudden pressure 739

drop. The gap decreases gradually at first, before accelerating over ~10 seconds, up to a 740

maximum adjustment rate that is dependent on the deviation of the normal force from the 741

nominal value (25 µm/s/N) (Figure B2). This response explains the prolonged drop in 742

normal force in experiment SPS1 (Fig. 10), and means that the measured timescale for 743

compaction in Section 3.5 (Experiments SPS1 and APS1) must be treated as an upper 744

bound on the intrinsic compaction timescale. 745

746

Acknowledgements 747

This work was supported by a grant from Institutional Support at Los Alamos National 748

Lab via the Institute of Geophysics and Planetary Physics. 749

750

References 751

Aharonov, E., and D. Sparks (1999), Rigidity phase transition in granular packings, 752

Physical Review E, 60(6), 6890-6896. 753

Bagnold, R. A. (1954), Experiments on a Gravity-Free Dispersion of Large Solid Spheres 754

in a Newtonian Fluid under Shear, Proceedings of the Royal Society of London 755

Series a-Mathematical and Physical Sciences, 225(1160), 49-63. 756

Ben-Naim, E., J. B. Knight, E. R. Nowak, H. M. Jaeger, and S. R. Nagel (1998), Slow 757

relaxation in granular compaction, Physica D, 123(1-4), 380-385. 758

Bocquet, L., W. Losert, D. Schalk, T. C. Lubensky, and J. P. Gollub (2001), Granular 759

shear flow dynamics and forces: Experiment and continuum theory, Physical 760

Review E, 65(1) 761

Caglioti, E., V. Loreto, H. J. Herrmann, and M. Nicodemi (1997), A ''tetris-like'' model 762

for the compaction of dry granular media, Physical Review Letters, 79(8), 1575-763

1578. 764

Campbell, C. S. (1990), Rapid Granular Flows, Annual Review of Fluid Mechanics, 22, 765

57-92. 766

Campbell, C. S. (2005), Stress-controlled elastic granular shear flows, Journal of Fluid 767

Mechanics, 539, 273-297. 768

Campbell, C. S. (2006), Granular material flows - An overview, Powder Technology, 769

162(3), 208-229. 770

Cates, M. E., J. P. Wittmer, J. P. Bouchaud, and P. Claudin (1998), Jamming, force 771

chains, and fragile matter, Physical Review Letters, 81(9), 1841-1844. 772

Clement, E. (1999), Rheology of granular media, Current Opinion in Colloid & Interface 773

Science, 4, 294-299. 774

Coghill, P. C., P. J., and P. Giang (2011), Ultrasonic velocity measurements in powders 775

and their relationship to strength in particles formed by agglomeration, Powder 776

Technology, 208(3), 694-701. 777

da Cruz, F., S. Emam, M. Prochnow, J. N. Roux, and F. Chevoir (2005), Rheophysics of 778

dense granular materials: Discrete simulation of plane shear flows, Physical 779

Review E, 72(2) 780

Hickman, S., R. Sibson, and R. Bruhn (1995), Introduction to Special Section - 781

Mechanical Involvement of Fluids in Faulting, Journal of Geophysical Research-782

Solid Earth, 100(B7), 12831-12840. 783

Hutton, C. O. (1959), Mineralogy of beach sands between halfmoon and monterey bays, 784

California, California Div. Mines Spec. Rept., 59 785

Iverson, R. M. (1997), The physics of debris flows, Reviews of Geophysics, 35(3), 245-786

296. 787

Jaeger, H. M., S. R. Nagel, and R. P. Behringer (1996), Granular solids, liquids, and 788

gases, Reviews of Modern Physics, 68(4), 1259-1273. 789

Jia, X., C. Caroli, and B. Velicky (1999), Ultrasound propagation in externally stressed 790

granular media, Physical Review Letters, 82(9), 1863-1866. 791

Johnson, P. A., and X. Jia (2005), Nonlinear dynamics, granular media and dynamic 792

earthquake triggering, Nature, 437(7060), 871-874. 793

Johnson, P. A., H. Savage, M. Knuth, J. Gomberg, and C. Marone (2008), Effects of 794

acoustic waves on stick-slip in granular media and implications for earthquakes, 795

Nature, 451(7174), 57-U55. 796

Jop, P., Y. Forterre, and O. Pouliquen (2006), A constitutive law for dense granular 797

flows, Nature, 441, 727-730. 798

Knight, J. B., C. G. Fandrich, C. N. Lau, H. M. Jaeger, and S. R. Nagel (1995), Density 799

Relaxation in a Vibrated Granular Material, Physical Review E, 51(5), 3957-3963. 800

Liu, A., and S. Nagel (2001), Jamming and Rheology: Constrained dynamics on 801

microscopic and macroscopic scales, Taylor & Francis, London. 802

Liu, A. J., and S. R. Nagel (1998), Nonlinear dynamics - Jamming is not just cool any 803

more, Nature, 396(6706), 21-22. 804

Liu, A. J., and S. R. Nagel (2010), The Jamming Transition and the Marginally Jammed 805

Solid, Annual Review of Condensed Matter Physics, Vol 1, 1, 347-369. 806

Lu, K., E. E. Brodsky, and H. P. Kavehpour (2007), Shear-weakening of the transitional 807

regime for granular flow, Journal of Fluid Mechanics, 587, 347-372. 808

Majmudar, T. S., and R. P. Behringer (2005), Contact force measurements and stress-809

induced anisotropy in granular materials, Nature, 435(7045), 1079-1082. 810

Makse, H. A., N. Gland, D. L. Johnson, and L. Schwartz (2004), Granular packings: 811

Nonlinear elasticity, sound propagation, and collective relaxation dynamics, 812

Physical Review E, 70(6) 813

Marone, C., C. B. Raleigh, and C. H. Scholz (1990), Frictional Behavior and Constitutive 814

Modeling of Simulated Fault Gouge, Journal of Geophysical Research-Solid 815

Earth and Planets, 95(B5), 7007-7025. 816

Marone, C. (1991), A Note on the Stress-Dilatancy Relation for Simulated Fault Gouge, 817

Pure and Applied Geophysics, 137(4), 409-419. 818

Mehta, A. (Ed.) (2007), Granular Physics, Cambridge Univ. Press, Cambridge. 819

Melosh, H. J. (1979), Acoustic Fluidization - New Geologic Process, Journal of 820

Geophysical Research, 84(Nb13), 7513-7520. 821

Melosh, H. J. (1996), Dynamical weakening of faults by acoustic fluidization, Nature, 822

379(6566), 601-606. 823

MiDi, G. (2004), On dense granular flows, Eur. Phys. J E, 14, 341-365. 824

Nowak, E. R., J. B. Knight, M. L. Povinelli, H. M. Jaeger, and S. R. Nagel (1997), 825

Reversibility and irreversibility in the packing of vibrated granular material, 826

Powder Technology, 94(1), 79-83. 827

Paull, C. K., P. Mitts, W. Ussler, R. Keaten, and H. G. Greene (2005), Trail of sand in 828

upper Monterey Canyon: Offshore California, Geological Society of America 829

Bulletin, 117(9-10), 1134-1145. 830

Reynolds, O. (1885), On the dilatancy of media composed of rigid particles in contact, 831

Philos. Mag., 20(469) 832

Savage, S. B. (1984), The mechanics of rapid granular flows, Adv. Appl. Mech, 24, 289-833

366. 834

Sleep, N. H., and M. L. Blanpied (1992), Creep, Compaction and the Weak Rheology of 835

Major Faults, Nature, 359(6397), 687-692. 836

Stearns, S. D. (2003), Digital Signal Processing with Examples in MATLAB, Boca Raton, 837

Florida. 838

Thompson, P. A. (1971), Compressible-fluid dynamics, McGraw-Hill, New York. 839

Thompson, P. A., and G. S. Grest (1991), Granular Flow - Friction and the Dilatancy 840

Transition, Physical Review Letters, 67(13), 1751-1754. 841

Tordesillas, A., J. Y. Shi, and T. Tshaikiwsky (2011), Stress-dilatancy and force chain 842

evolution, International Journal for Numerical and Analytical Methods in 843

Geomechanics, 35(2), 264-292. 844

Walton, K. (1987), The Effective Elastic-Moduli of a Random Packing of Spheres, 845

Journal of the Mechanics and Physics of Solids, 35(2), 213-226. 846

847

Figure Captions 848

Figure 1. Cartoon depiction of end-member granular flow regimes. (a) In the low shear 849

rate quasi-static regime, boundary stresses are supported elastically through force chains. 850

Thin lines represent grain-grain contacts and arrows represent forces acting on the 851

boundaries. (b) In the high shear-rate grain-inertial regime, stresses are supported through 852

momentum transfer in collisions. (c) Cartoon of experimental behavior in end member 853

regimes. The y-axis represents stress under constant volume conditions or shear zone 854

thickness under constant stress conditions. In the quasi-static regime, dilatation and 855

stress are independent of shear rate, resulting in a friction-like rheology. In the grain 856

inertial regime, dilatation and stress are proportional to the momentum transfer rate, 857

resulting in a power-law viscous-like rheology. This study focuses on the intermediate 858

shear rate transitional regime, where stresses are supported elastically, but vibration 859

becomes important for force chain stability. 860

861

Figure 2. Experimental apparatus: TA Instruments AR2000ex torsional rheometer, with 862

parallel plate geometry. (a) Photograph of test chamber in mounting bracket. The 2 cm 863

quartz jacket is here filled with glass beads. (b) Schematic of the experimental geometry. 864

Shear rate ω and normal stress σ are controlled through the upper rotor. The normal 865

stress σ is held constant by adjusting the rotor height h. We also measure, but do not 866

report on, the torque τ (i.e. shear stress) required to shear at a given velocity. 867

868

Figure 3. Granular media used in velocity stepping experiments: a) smooth glass beads, 869

b) angular Santa Cruz beach sand 870

871

Figure 4. Steady-state thickness vs. shear rate, for sheared angular sand and glass beads. 872

Individual curves represent multiple up-going and down-going velocity ramps (Table 1), 873

and thick error bars show means and standard deviations between runs. Thickness is 874

independent of shear rate at low shear rates, and strongly dependent on shear rate for 875

intermediate and high shear rates. Compaction is observed at intermediate shear rates for 876

angular sand, but not for smooth glass beads. Thickness is reported relative to starting 877

thickness, offset by 100 microns for the glass beads. 878

879

Figure 5. Steady state layer thickness as a function of shear rate for (a) angular sand and 880

(b) smooth glass beads. Layer thickness is given relative to the minimum layer thickness 881

under acoustic vibration. Thin lines are individual runs (Table 1), thick lines are averages 882

over all runs. In the absence of vibration, the layer thickness is largely independent of 883

shear rate at low shear rates, and strongly dependent at higher shear rates. In the presence 884

of acoustic vibration, the layer thickness is reduced, and is dependent on shear rate over 885

the entire range. For angular sand, a reduction in steady state thickness occurs at 886

intermediate velocities, but not for smooth glass beads. 887

888

Figure 6. Sample thickness and shear rate as a function of experimental time, for angular 889

sand, showing evolution of thickness during external vibration (Experiment VRS3, Table 890

1). Shear rate is ramped up and down (b), producing repeatable and reversible changes in 891

sample thickness (a). Boxes highlight individual slow-to-fast velocity ramps (c.f. Fig. 892

5a). External acoustic vibration is applied at ~4000 seconds, resulting in logarithmic 893

irreversible compaction, and a qualitative change in the reversible thickness at low 894

velocity (labeled ‘LV dip’). Vibration ceases at ~10,000 seconds, after which the low 895

shear-rate compaction returns to the pre-vibration behavior (‘no LV dip’). 896

897

Figure 7. Acoustic vibration amplitude, recorded by an accelerometer attached to the 898

sample jacket. Transitions in steady-state thickness (see Figure 5) correspond to 899

transitions in the amplitude of shear-induced vibration. The blue curve is vibration 900

produced by shearing angular sand; the lower green curve is the vibration produced by 901

glass beads. The red curve shows the superposition of external acoustic vibration on 902

shear-induced vibration produced by angular sand. The accelerometer begins to clip at a 903

shear rate of ~7 rad/s (grey patch), so the rollover is not entirely physical. At 1 rad/s the 904

shear-induced vibration is equivalent to the external vibration amplitude (dashed red 905

line). 906

907

Figure 8. Relative power of the shear-induced acoustic vibration signal, as recorded by 908

the accelerometer. Power is the integral of the instrumental velocity spectrum between 0 909

and 35 Hz, normalized by the peak value. Power increases linearly with shear rate 910

(straight line for reference). The error bars show the standard deviation between 911

experimental runs. The majority of the variation at higher shear rates is due to fluctuation 912

in normal force due to the broad normal force tolerance (+/- 10 %) allowed by the 913

rheometer. 914

915

Figure 9. The response of a steadily sheared granular medium to transient acoustic 916

pulses. The sample is sheared at 0.01 rad/s, and subjected to 60-second acoustic pulses 917

(marked by vertical dashed black lines). (a) shows sample thickness relative to the start 918

of the acoustic vibration, and (b) shows normal stress. There is a small normal stress dip 919

for the highest amplitude vibration (red curve), reflecting a lag in the rheometer normal 920

force control (Appendix B). Black lines show logarithmic fits to the data (Eq. 4). The 921

legend gives the strain amplitude of the acoustic vibration as measured by the 922

accelerometer. 923

924

Figure 10. (a) The response of a steadily sheared granular medium to a velocity step. 925

The sample is sheared at 0.1 rad/s, and subjected to 60-second intervals of higher shear 926

rate (bounded by vertical dashed black lines). The legend gives the shear rate in rad/s 927

(equivalent to cm/s at the rotor rim). The compaction and recovery are fit to Equations 4 928

and 5, respectively (black curves). (b) Normal stress drops as compaction begins, due to 929

the time lag in the rheometer response (Appendix B). 930

931

Figure 11. Compaction magnitude b (Eq. 4) is a consistent function of acoustic strain 932

amplitude for external (Table 2: Aps1) and internally generated (Table 2: Sps1) acoustic 933

vibration. Compaction magnitude increases linearly above a threshold strain amplitude 934

of about 10-5. Triangles show compaction for external vibration, squares correspond 935

internal, shear-generated vibration, for shear rates of 1, 2, 3.2, and 5 rad/s, from left to 936

right. Each point is an average of 5-10 runs (Table 2), and error bars show standard 937

deviations between runs. 938

939

Figure 12. Compaction magnitude parameter b, fit through Eq. 10 to the steady-state 940

thickness vs. shear rate (Fig. 5). The red curve, showing compaction rate due to external 941

vibration, is flat at low shear rates, consistent with the constant external acoustic strain 942

amplitude. Shear-induced compaction (blue curve) increases linearly with shear-rate, 943

consistent with the linear increase in the power of the internally generated acoustic 944

vibration (Fig. 8). Compaction rate for the glass beads (green curve) is considerably 945

smaller. The black dashed line shows a linear relationship for reference. The rollover at 946

high shear-rates reflects the entry into the inertial granular flow regime, where sample 947

dilation is dominated by grain collisions. Compare to acoustic vibration amplitude as a 948

function of shear rate (Fig. 7). 949

950

Figure 13. Steady-state thickness of the shear zone as a function of shear rate (Eq. 10). 951

Observations (curves with error bars) are the same as in Figure 5: the upper curve is 952

angular sand without vibration, the lower curve is with external vibration. The green 953

curve is the contribution from external vibration through Equation 10 with a constant 954

compaction magnitude bacoustic; the blue curve is the contribution from shear-induced 955

vibration with compaction magnitude bshear a linear function of shear rate; the black curve 956

shows the linear sum of the two compaction components. The increase in thickness for 957

both curves at high shear rates is caused by inertial dilation and is not addressed by the 958

model. 959

960

Figure A1. Displacement profiles for different velocities. The left panel shows a 961

snapshot of the angular sand, which is being sheared to the right. The following panels 962

show horizontal cross-correlation of snapshots taken at a constant displacement (10 cm) 963

but for different shear rates (0.1, 1, and rad/s from left to right). The darker colors reflect 964

higher cross correlation coefficients. Grain rotation and non-horizontal particle motion 965

prevent correlation beyond a few mm displacement. An exponential function (black 966

curve) is fit to the first cross-correlation image (0.1 cm/s) and superposed on the adjacent 967

panels. The shear-rate profile does not change except through a multiplicative pre-factor. 968

969 Figure B1. Cartoon of experimental setup. The rheometer rotor is lowered into contact 970

with an inflated plastic bag, which is pinched on the edge to increase the pressure. At the 971

start of the experiment, the bag is released, resulting in a sudden relaxation in pressure 972

(dashed outline). 973

974 Figure B2. Gap adjustment rate as a function of the reduction in normal force. Black 975

circles show the response to continuous fluctuations in the resisting force. The maximum 976

adjustment rate depends linearly on the deviation of normal force from the nominal value 977

(25 µm/s/N). The two colored paths show sudden force drop experiments, where the 978

normal force is reduced from 1 N to 0.5 N (green) or 0 N (red). For these sudden drops, 979

there is a delayed rheometer response, during which the gap adjustment rate accelerates 980

from zero (large circles at start of path) to the maximum rate over ~10 seconds. Inset: 981

paths for sudden force drop experiments as a function of time, showing the ~10 second 982

delay. The maximum rates are shown by straight lines for reference. Points A and B on 983

the inset corresponds to points A and B along the path in the main figure. 984

985

Tables 986

Table 1. Velocity ramp experiments 987

Name Material Number w/o vibrationa

Number w/ vibrationa

correctionb

VRS1 Sand 4 (2,2) 8 (4,4) yes VRS2 Sand 2 (1,1) 8 (4,4) No VRS3* Sand 10 (5,5) 6 (3,3) No VRS4 Sand 0 8 (4,4) No VRG1 Glass beads 17 (9,8) 6 (3,3) No a (N,M) signifies N slow-to-fast and M fast-to-slow velocity ramps, respectively 988 b linear correction for run-in phase applied to runs without vibration 989 * shown in Figure 6 990

991

Table 2. Pulse experiments (sand) 992

Name Pulse type Starting shear rate

Pulse shear rate (rad/s)

Acoustic strain

b (µm)a Runs

SPS1 Shear 0.1 1.0 1.3 × 10-5 1.9 ± 1.9 10 0.1 2.0 2.2 × 10-5 5.6 ± 2.7 5 0.1 3.16 2.9 × 10-5 10.5 ± 0.6 5 0.1 5.0 3.5 × 10-5 14.2 ± 1.2 5 APS1 Acoustic 0.01 n/a 1.3 × 10-6 0.9 ± 4.6 13 0.01 3.8 × 10-6 2.7 ± 3.0 10 0.01 1.4 × 10-5 0.1 ± 3.3 5 0.01 2.4 × 10-5 8.6 ± 3.6 5 0.01 3.0 × 10-5 9.1 ± 3.3 6 a Compaction magnitude (Eq. 4) 993

994

Table 3: Fit parameters for steady-state thickness model (Eq. 10) 995

bacoustic (µm)a bshear (µm/rad/s)a d (rad)b τ- (s)a 10 18 5.5 1 a fit to combined velocity ramp experiments VRS1-4 (Fig. 12) 996 b fit to acoustic pulse experiment SPS1997

998

Figure    1  

(a)  Low  shear  rate   (b)  High  shear  rate  

Dilataion  or  stress  

Shear  rate  

Quasi-­‐sta5c  

Grain-­‐iner5al  

Transi5onal    ?   ?   ?   ?   ?  

(c)  

2  cm  

2  cm  

h  

ω  

σ  

ε  

Accelerometer  

Vibra5on  Transducer  

(a)   (b))  

Figure    2  

(a)   (b)  

1  mm  

Figure    3  

Figure    4  

100  

50  

0  

-­‐50  

-­‐100  

-­‐150  

Thickness  (μm)  

Shear  rate  (rad/s)  10-­‐3   10-­‐2   10-­‐1   100   101   102  

Angular  sand  Glass  beads  

Figure  5  

(b))  

(a))   160  

120  

80  

40  

0  

-­‐40  

Thickness  (μm)  

Shear  rate  (cm/s)  10-­‐3   10-­‐2   10-­‐1   100   101   102  

Vibra5on  No  vibra5on  

Thickness  (μm)  

Shear  rate  (cm/s)  10-­‐3   10-­‐2   10-­‐1   100   101   102  

Vibra5on  No  vibra5on  

200  

150  

100  

50  

0  

-­‐50  

(b))  

LV  dip   No  LV  dip  

(a)  

(b)  

Thickness  (m

m)  

Experimental  5me  (s)  0  

12.6  

12.5  

2000   4000   6000   8000   10000   12000   14000   16000  

12.7  

12.8  

0   2000   4000   6000   8000   10000   12000   14000   16000  

102  

100  

10-­‐2  

10-­‐4  

Shear  rate    (rad/s)  

(a))  

(b))  

Figure  6  

Figure    7  

0  

Acous5c  strain  am

plitu

de  

Shear  rate  (rad/s)  10-­‐3   10-­‐2   10-­‐1   100   101   102  

Angular  sand  Angular  sand  +  Vibra5on  Glass  beads  

1  

2  

3  

4  

5  

6  ×  10-­‐5  

Figure    8  

0  

Acous5c  strain  am

plitu

de  

Shear  rate  (rad/s)  100  

0.2  

0.4  

0.6  

0.8  

1  

90  80  70  60  50  40  30  20  10  0  

(a)  

(b)  

1.3×10-­‐6  

3.8×10-­‐6  

1.4×10-­‐5  

2.4×10-­‐5  

3.0×10-­‐5  

Thickness  (m

m)  

-­‐20  

-­‐30  

-­‐10  

10  

-­‐60   -­‐40   -­‐20   0   20   40   60   80   120  

0  

100  Time  (s)  

Normal  Stress  (Pa)  

3000  

-­‐60   -­‐40   -­‐20   0   20   40   60   80   120  

2000  

100  

2000  

0  

-­‐3000  

-­‐1000  

-­‐2000  

Time  (s)  

Figure    8  

(a)  

(b)  

1  2  3.2  5  

Thickness  (m

m)  

-­‐20  

-­‐30  

-­‐10  

10  

-­‐50   0   50  

0  

Time  (s)  

Normal  Stress  (Pa)  

1  

0.5  

0  

-­‐1  

-­‐0.5  

Figure    10  

100   150   200   250   300   350  

-­‐50   0   50  Time  (s)  

100   150   200   250   300   350  

-­‐40  

×  104  

6  

Compac5on

 magnitude

 (μm)  

Acous5c  Strain  Amplitude  4  

8  

14  

3.5  3  2.5  2  1.5  1  0.5  0  

10  

12  

16  

-­‐4  

-­‐2  

4  

0  

2  

×  10-­‐5  

Shear-­‐induced  (Internal)  vibra5on  External  vibra5on  

Figure    11  

Iner5al  regime  Internal  vibra5on  External  vibra5on  Glass  beads  

Compac5on

 magnitude

 (μm)  

Shear  rate  (rad/s)  10-­‐3   10-­‐2   10-­‐1   100   101   102  

103  

102  

101  

100  

10-­‐1  

Figure    12  

-­‐60  

-­‐80  

-­‐100  

Thickness  (μm)  

Shear  rate  (rad/s)  10-­‐3   10-­‐2   10-­‐1   100   101   102  

-­‐0  

-­‐20  

-­‐40  

20  

80  

60  

40  

Figure    13  

No  vibra5on  Acous5c  vibra5on  Fit  (shear-­‐induced)  Fit  (external  acous5c)  Fit  (sum)  

0.1  cm/s   1  cm/s   10  cm/s  

Displacement  (cm)  -­‐0.5   0   0.5   -­‐0.5   0   0.5   -­‐0.5   0   0.5  

Dep

th  (cm)  

0  

0.1  

0.2  

0.3  

0.4  

0.5  

0.6  

0.7  

0.8  

0.9  

0  

0.1  

0.2  

0.3  

0.4  

0.5  

0.6  

0.7  

0.8  

0.9  

(cm)  -­‐0.5   0   0.5  

Figure    A1  

h  

σ  

pinched  

relaxed  

Figure    B1  

A  

B  

B  A  

5  

0  

Adjustm

ent  rate  (μm/s)  

Normal  force  drop  (N)  0  

15  

10  

20  

25  

Figure    B2  

0.2   0.4   0.6   0.8   1  

0  

-­‐20  

-­‐40  

-­‐60  

-­‐80  

-­‐100  

Gap  (μ

m)  

Time  (s)  0   5   10   15   20  


Recommended