roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (candela...

69
1 Roughness of fault surfaces over nine decades of length scales 1 Thibault Candela 1 , François Renard 1,2 , Yann Klinger 3 , Karen Mair 2 , Jean Schmittbuhl 4 , and Emily 2 E. Brodsky 5 3 4 1 ISTerre, University of Grenoble I, CNRS, OSUG, BP 53, 38041 Grenoble, France. 5 E-mail: [email protected]; [email protected] 6 2 Physics of Geological Processes, University of Oslo, Oslo, Norway. E-mail: [email protected] 7 3 Institut de Physique du Globe de Paris, Sorbonne Paris cité, Univ. Paris Diderot, UMR 7154 CNRS, 8 1 rue Jussieu, F-75005 Paris, France. E-mail: [email protected] 9 4 UMR 7516, Institut de Physique du Globe de Strasbourg, Strasbourg, France. E-mail: 10 [email protected] 11 5 Department of Earth and Planetary Sciences, University of CaliforniaSanta Cruz, Santa Cruz, 12 California 95064, USA. E-mail: [email protected] 13 14

Upload: others

Post on 19-Aug-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

1

Roughness of fault surfaces over nine decades of length scales 1

Thibault Candela 1, François Renard

1,2, Yann Klinger

3, Karen Mair

2, Jean Schmittbuhl

4, and Emily 2

E. Brodsky 5 3

4

1ISTerre, University of Grenoble I, CNRS, OSUG, BP 53, 38041 Grenoble, France. 5

E-mail: [email protected]; [email protected] 6

2Physics of Geological Processes, University of Oslo, Oslo, Norway. E-mail: [email protected] 7

3Institut de Physique du Globe de Paris, Sorbonne Paris cité, Univ. Paris Diderot, UMR 7154 CNRS, 8

1 rue Jussieu, F-75005 Paris, France. E-mail: [email protected] 9

4UMR 7516, Institut de Physique du Globe de Strasbourg, Strasbourg, France. E-mail: 10

[email protected] 11

5Department of Earth and Planetary Sciences, University of California–Santa Cruz, Santa Cruz, 12

California 95064, USA. E-mail: [email protected] 13

14

Page 2: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

2

Abstract 15

We report on the roughness measurements of five exhumed faults and eight surface ruptures over a 16

large range of scales: from 50 micrometers to 50 km. We used three scanner devices (LiDAR, laser 17

profilometer, white light interferometer), spanning complementary scale ranges from 50 micrometers 18

to 10 m, to measure the 3-D topography of the same objects, i.e. five exhumed slip surfaces (Vuache-19

Sillingy, Bolu, Corona Heights, Dixie Valley, Magnola). A consistent geometrical property emerges 20

as the morphology of the slip surfaces shows a straight line covering five decades of length-scales in a 21

log-log plot where axes are fault roughness and spatial length scale. The observed fault roughness is 22

scale dependent, with a common self-affine behavior described by four parameters: two power-law 23

exponents H, constant among all the faults studied and slightly anisotropic ( 07.058.0// H in the 24

slip direction and 04.081.0 H perpendicular to it), and two pre-factors showing a quite large 25

variability. For larger scales between 200 m and 50 km, we have analyzed the 2-D roughness of the 26

surface rupture of eight major continental earthquakes. These ruptures also show self-affine behavior 27

( 1.08.0 RH ), which is consistent with the slip-perpendicular behavior of the smaller-scale 28

measurements. We show that small degrees of non-alignment between the slip orientation and the 29

exposed trace result in sampling the slip-perpendicular geometry. Although a data gap exists between 30

the scanned fault scarps and rupture traces, the measurements are consistent within the error bars with 31

a single self-affine scaling exponent in the slip-perpendicular direction, i.e. consistent dimensionality, 32

over nine decades of length scales. 33

34

35

36

37

38

39

Page 3: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

3

1. Introduction 40

Faults appear most commonly as more or less continuous linear breaks at the surface of the 41

Earth, but their traces show wavy irregularities at all scales (Brown and Scholz, 1985). These 42

irregularities, which we call roughness, control the geometry (Power et al., 1987), mechanics, and 43

transport properties of fault zones (Power and Durham, 1997) and contribute to their 3D architecture 44

(Bistacchi et al., 2010; Faulkner et al., 2010). Fault roughness may control several faulting processes 45

and parameters such as the total resistance to slip, the aseismic versus seismic behavior, the alteration 46

of shear resistance during sliding, the magnitude of stress concentration and heterogeneity in the fault 47

zone (Chester and Fletcher, 1997; Chester and Chester, 2000), and the deformation and damage of the 48

rock on either side of the fault (Arvid and Fletcher, 1994; Dieterich and Smith, 2008; Griffith et al., 49

2010). 50

The morphology of fault planes controls the dynamics of faulting and slip. Based on the results 51

of kinematic source inversion models that reconstitute the spatio-temporal evolution of slip during an 52

earthquake, several studies show that both coseismic slip and stress appear to be very heterogeneous 53

along the fault plane (Bouchon, 1997; Mai and Beroza, 2002). A possible explanation is that the fault 54

plane is rough and asperities concentrate stress and slip heterogeneities at various spatial scales 55

(Schmittbuhl et al., 2006; Candela et al., 2011a, 2011b). Earthquake activity and aseismic creep can 56

create and destroy this roughness that causes the heterogeneity of the stress field in the fault zones. 57

Other studies have also shown the importance of non-planar structures in the rupture propagation 58

(e.g. Aochi and Madariaga, 2003) and the close relationship between the rupture geometry and its 59

propagation velocity (Vallée et al., 2008; Bouchon et al., 2010). Numerical models of earthquake 60

rupture and strong motion need accurate 3-D morphological models of fault surfaces to improve 61

simulations of rupture scenarios. 62

As direct observations are not possible at the depths of earthquake nucleation, data of exhumed 63

fault scarps (Power et al., 1987; Renard et al., 2006; Sagy et al., 2007; Candela et al., 2009; Brodsky 64

et al., 2011 and references therein) or earthquake surface ruptures (Wesnousky, 2006, 2008; Klinger, 65

2010) provide the means to characterize fault plane morphology over a wide range of spatial scales. 66

Page 4: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

4

Pioneering fault roughness measurements (Brown and Scholz, 1985; Power et al., 1987; Power 67

and Tullis, 1991; Schmittbuhl et al., 1993) performed on exhumed scarps found that their roughness 68

cannot be described by a single number such as the standard deviation of the roughness amplitude. 69

Rather, fault surface topography was measured as non-stationary and more particularly self-affine 70

fractal, the amplitude of the topography increasing with the wavelength under consideration. 71

A 2-D rough profile (Figure 1) is self-affine if it remains statistically invariant under the scaling 72

transformation zzxx H , (Feder, 1988; Meakin, 1998), where x is the coordinate 73

along the 2-D profile, z the roughness amplitude and H the Hurt exponent (or roughness 74

exponent). If the power law scaling exponent lies in the range 10 H , different magnification 75

factors will be needed in the directions parallel and perpendicular to the profile for a small portion of 76

the profile to appear statistically similar to the entire profile (Figure 1). As a consequence, in this case, 77

the slope at large scales along a self-affine profile scales as 1

/

H

xxzs , and tends to flatten 78

for long wavelengths, suggesting a significant role of the small spatial scales (Schmittbuhl et al., 79

1995a). Since the exponent H does not contain any information on the amplitude of the signal, and is 80

related only to the progression in space, a second parameter is needed to describe fully the power law 81

and the amplitude of the scaling behavior. In this study, we call this parameter the pre-factor of the 82

power law. 83

Pioneering studies measured 2-D profiles on exhumed fault planes using mechanical 84

profilometers (Brown and Scholz, 1985; Power et al., 1987; Power and Tullis, 1991). Recently, with 85

the development of a new generation of 3-D laser scanners, fast and accurate acquisitions of 86

topographic data are now available, allowing more fault surfaces to be characterized (Renard et al., 87

2006; Sagy et al., 2007; Candela et al., 2009; Brodsky et al., 2011). These first results confirmed the 88

roughness of several faults over a wide range of spatial scales and their morphologically anisotropic 89

property with a Hurst exponent smaller in the slip direction than perpendicular to it. 90

In the present study, we investigate roughness properties of five fault surfaces in different 91

geological contexts (with varying lithology, accumulated displacement, tectonic regime) using three 92

independent scanner devices (a Light Detection And Ranging apparatus – also called LiDAR, a laser 93

Page 5: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

5

profilometer, and a white light interferometer), spanning a range of scales from 5.10-5

m to 10 m. In 94

addition, we analyzed the geometry of eight map-scale high resolution rupture traces of large 95

continental earthquakes, giving access to a range of scales from 200 m to 50 kilometers. 96

The availability of LiDAR has greatly facilitated the measurement of topography data from 97

natural fault surfaces. However, the level of noise inherent in the measurements of these instruments, 98

which is estimated by scanning flat reference surfaces (see Section 3), requires caution when 99

interpreting the results of LiDAR topography measurements. In contrast to the noise level of the 100

LiDAR instruments, the noise level inherent in the laboratory laser profilometer and the white light 101

interferometer (WLI) profilometer is much lower, well below the magnitude of the surface 102

topography. For this reason the laser profilometer and WLI data can be considered to be essentially 103

noise free for the measurements presented here. 104

We characterize the self-affine scaling of both data sets (scanned fault surfaces and rupture 105

traces) using the same statistical tool, i.e. the Fourier power spectrum analysis. This is a robust 106

technique well-suited for characterizing self-affine roughness of fault zones (Candela et al., 2009). A 107

self-affine model implies that the Fourier spectrum as a function of either the spatial frequency or 108

wavelength plots as a linear trend in log-log coordinates. Two parameters describe such a self-affine 109

model in the spectral domain: the slope of the power spectrum and its pre-factor (i.e. the intercept at a 110

given length scale) on a log-log plot of the Fourier spectrum. The slope (directly proportional to H ) 111

describes how the roughness changes with scale, while the pre-factor determines the magnitude of the 112

surface roughness at a given scale (Mandelbrot, 1983, p. 350, Power and Tullis, 1991). Both 113

parameters are necessary and sufficient to describe a self-affine geometry. We compute these two 114

parameters for the fault surfaces and rupture traces in order to decipher if a global tendency emerges 115

and/or to characterize the fluctuations. 116

In Section 2, the characteristics of both the scanned fault surfaces and rupture traces are described. 117

Section 3 is devoted to the presentation and the application of the roughness analysis method. In 118

Section 4, we present results that indicate an anisotropic self-affine model with two different Hurst 119

exponents in perpendicular directions and varying pre-factors, could fit each fault scarp data set, 120

Page 6: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

6

despite different geological contexts. Secondly, we highlight a similar self-affine regime that fits the 121

eight earthquake rupture traces. In Section 5 we discuss the possibility that a unique anisotropic self-122

affine geometrical model is maintained from the micrometric scale to the map-scale earthquake 123

surface rupture. The variability of the pre-factors for each data set (fault surfaces and ruptures traces) 124

is discussed with respect to the different geological contexts sampled and more particularly the 125

accumulated slip. Physical processes at work in the generation of the observed roughness are also 126

debated. 127

128

2. Fault roughness data 129

2.1. Exhumed fault scarps 130

We have analyzed five natural fault surfaces which were selected because of their particularly 131

well preserved slip surfaces, large exposures and few pits or weathering patterns. Existing data sets on 132

the Vuache-Sillingy (Renard et al., 2006; Candela et al., 2009, Angheluta et al., 2011) and Magnola 133

(Candela et al., 2009) faults have been updated and extended with three new faults (Corona Heights, 134

Dixie Valley, and Bolu). A complete list of our fault data, including GPS locations, nature of the rock, 135

direction of motion, and estimated finite geological offset, is given in Table 1. The finite geological 136

offset is always the most difficult of these parameters to estimate. In fact, if the total offset of a fault 137

zone can be estimated from the displacement of geological markers, the total displacement could have 138

occurred on several parallel slip surfaces, making it difficult to estimate the total slip on a given slip 139

surface. For each of our scanned surfaces, we will consider a range of slips bracketed by the extreme 140

values: the total slip of the fault zone and a minimum slip we could estimate from field observations. 141

The Corona Heights and Dixie Valley faults cut through silicate rocks. The Dixie Valley (Basin 142

and Range province in Nevada) fault has a mainly normal slip component and cross-cuts through 143

rhyolite. Chemical changes during faulting at depth have altered the mineralogy and chemical 144

composition of the rock in the fault zone (Power et al., 1987; Power and Tullis, 1989, 1992). The 145

material that forms the fault consists almost entirely of secondary quartz, a mineral which is 146

extremely resistant to weathering and allows extremely good preservation of the slip surfaces. Using 147

Page 7: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

7

gravity studies combined with reflection seismology studies (Okaya and Thompson, 1985), Power and 148

Tullis (1989) estimate that the total normal slip is probably between 3 and 6 km. Geological and 149

mineralogical constraints indicate that the slickenside surface formed at depths of less than 2 km and 150

temperatures less than Co270 (Power and Tullis, 1989). Major historical earthquakes and 151

microearthquakes occurred on the region of the studied fault scarp (Wallace and Whitney, 1984; 152

Dozer, 1986). Additionally, Power and Tullis (1989) have argued that seismic faulting played a role in 153

the development of the slickenside surfaces, based on textural features they described in the fault 154

surface materials. 155

The Corona Heights strike-slip fault (Figure 2), located in the Castro district of San Francisco, 156

cross-cuts brown Franciscan cherts and was exposed by post-1906 earthquake anthropogenic 157

quarrying. The relatively recent exposure of the fault and the high resistance of cherts to weathering 158

allows for excellent preservation of the slip surface (Figure 2). In most places on the outcrop, an 159

anastomosing set of slip surfaces is present (Figure 2). Individual patches of the fault surfaces may 160

have been activated at different times and different depths. Although the total slip for the fault zone as 161

a whole could be large (> 1000 m), individual surfaces have recorded smaller (~1 m) displacements 162

which are difficult to precisely estimate due to the absence of well-defined structural markers. 163

The other three faults offset limestone rocks (Vuache-Sillingy in the French Alps, Magnola in the 164

Appenines, and Bolu in Turkey). The Vuache-Sillingy fault is an active strike-slip fault system in the 165

western part of the French Alps and has accumulated a total displacement in the kilometer range 166

(Thouvenot, 1998). The fault surface we analyzed (Renard et al., 2006; Candela et al., 2009) lies on a 167

short segment of this fault system, where the accumulated slip was small, in the range of 3010 168

meters, as estimated on aerial photographs. The fault plane was exhumed in the 1990s by quarrying 169

and, as a consequence, the LiDAR measurements were performed on fresh, vegetation free surfaces, 170

where weathering is minimal. The Magnola fault (Candela et al., 2009), in the Fuccineo area, is part 171

of the extensive fault system in central Apennines, Italy. This 15 km long normal fault shows 172

microseismic activity and presents an average vertical displacement larger than 500 meters. The site 173

we study has been recently exhumed (Palumbo et al., 2004; Carcaillet et al., 2008) with less alteration 174

Page 8: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

8

by weathering than older exhumed portions of the fault. The Bolu fault is part of the North Anatolian 175

strike-slip fault system. The study area (Figure 3) is a part of the section that ruptured during the 1944 176

earthquake (Kondo et al., 2005; Kondo et al., 2010; Barka, 1996). The small vertical component of 177

the motion (~1 m), compared to the dominant horizontal strike-slip motion (~3.5 m), was responsible 178

for the partial exhumation of the fault plane (Figure 3) during the 1944 earthquake (Barka, 1996). 179

More recently, anthropogenic activity (excavation for a garbage dump) also contributed to the 180

exhumation of the outcrop. The total geological offset of the North Anatolian fault can be of about 181

2585 km (Hubert-Ferrari et al., 2002), but it is not easy to define the slip accommodated 182

specifically on each individual sub-parallel slip surface constituting the fault zone of the Bolu 183

segment. Paleo-seismological investigations on the Bolu segment give a lower bound of 184

approximately 20 m (Kondo et al., 2005; Kondo et al., 2010). 185

The five faults studied show slip activity during the Quaternary. The Bolu fault records both the 186

propagation and termination of the 1944 earthquake. The Dixie Valley outcrop lies north of a segment 187

that broke in 1954; it is the same outcrop studied by Power and Tullis (1987). For three of these 188

faults, the slip surfaces were exhumed recently by anthropogenic activity during the 20th century 189

(Vuache-Sillingy, Bolu, and Corona Heights). These three surfaces were therefore exposed to 190

atmospheric alteration for only a short period of time and therefore their roughness reflects only 191

faulting processes. The Dixie Valley fault was exhumed by normal faulting activity combined with 192

local quarrying. In this region, desert weather conditions and the silicate rocks result in slow chemical 193

alteration. For this fault, we have chosen surfaces free of mechanical erosion, where the mirror-like 194

polishing due to the latest slips activity was still present. Finally, the Magnola slip surfaces were those 195

for which the alteration was the most important. For this fault, we selected slip surfaces that were 196

exhumed by the quaternary vertical activity of this normal fault, and for which the erosion was 197

minimal (i.e. surface away from a stream that could have increased the erosion rate); however we 198

cannot discount that weathering has altered the roughness properties of these surfaces. We will show 199

later that the roughness properties of the Magnola fault do not deviate significantly from those of the 200

Page 9: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

9

other faults and interpret this observation as evidence that weathering processes did not influence our 201

roughness analyses at the spatial scales we considered. 202

Even if it is difficult to accurately determine under which conditions at depth (confining 203

pressure, temperature, strain rate and chemical environment) fault surfaces were built, the five fault 204

surfaces studied here sample a range of different control parameters (total fault zone displacement 205

(10’s of meters to ~10’s of kilometers), lithology (rhyolite, chert and limestone), tectonic regime 206

(strike-slip, oblique and normal) which possibly could have controlled fault surface roughness. 207

208

2.2. Scanner devices and digital elevation models of fault roughness 209

Fault surface topography was scanned in the field using five different 3-D portable LiDAR laser 210

scanners that use the time of flight of a light beam to accurately measure distances. The laser scanner 211

records the topography of each exposed fault surface by collecting a cloud of points whose three 212

dimensional coordinates correspond to points on the fault surface (Renard et al., 2006; Sagy et al., 213

2007; Candela et al., 2009; Resor and Meer, 2009; Zhanyu et al., 2010). 214

The actual point spacing depends on the distance between the target and the scanner and a 215

chosen angular spacing. For each fault outcrops, fresh sub-surfaces were selected and scanned for 216

higher resolution acquisition (Figure 2, 3). 217

Our data sets cover surface scales from 21 m to

2800 m at a spatial length scale resolution x 218

from mm1 to mm30 . This spatial length scale resolution x corresponds to the point spacing after 219

the data processing (see Section 3), and is systematically taken to be twice as large as the average 220

irregular spacing during the acquisitions, that is from mm5.0 to mm15 . The actual precision in the 221

spatial positioning is estimated to be at most half the original average spacing, that is mm25.0 to 222

mm5.7 . The height precision achievable depends on scanning conditions and is closely related to 223

the spatial length scale resolution and to the roughness amplitude of the surface. In Table 1, z 224

represents the estimated amplitude of the instrumental noise. 225

Page 10: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

10

The scans were combined with digital photographs to distinguish clear slip surfaces from eroded 226

areas. This manual cleaning of the extremely large datasets (several tens of millions of points) was 227

completed using 3-D Reshaper software, a point cloud editor and visualization tool. Once all non-fault 228

features such as trees, grass, or anthropogenic structures were removed, the whole fault scarp or 229

selected smaller patches were analyzed (Figure 2, 3). Typically, in our data sets, less than 5% of 230

spurious points were removed from the raw scanner data. As pointed out by Candela et al. (2009), the 231

estimation of the fault surface properties was not significantly biased by the presence of randomly 232

distributed holes and missing data in the cloud of points (see Appendix A.1 for a quantitative analysis 233

of the bias inherent to data acquisition in the estimation of the geometrical properties of fault 234

surfaces). 235

In the laboratory, we used a home-made laser profilometer (Méheust, 2002), to measure samples 236

of the fault surfaces (between 2200 mm and

21000 mm , Figure 2 and 3), set on a 2-axis moving 237

table. Each surface is scanned by geometric triangulation, measuring the distance between the sample 238

and a laser head (Schmittbuhl et al., 2008; Candela et al., 2009). One difference between the LiDAR 239

and the profilometers is that with the laser profilometer, the data are regularly spaced. The spatial 240

length scale resolution x is equal here to the horizontal step, i.e. 20 microns. The actual precision in 241

the spatial positioning is 1 micron and the vertical resolution ( z ) is better than 1 micron. The 242

reliability and accuracy of the cloud of points obtained with this laser profilometer required that only 243

few spurious points were removed (less than 0.01%). 244

At the millimeter scale, the topography of several slip surfaces (between25.0 mm and

240 mm , 245

Figure 2 and 3) was measured using white light interferometry microphotography (Dysthe et al., 246

2002). This is done with a microscope that uses a broad-band white light source and that is coupled to 247

a Michelson interferometer. A reference arm creates interference fringes with maximum intensity at 248

equal optical path lengths of the imaging beam and reference beam. By vertical movement of the 249

sample and simultaneous image capturing, the interference, intensity envelope, and thereby the 250

relative height of the imaged surface at each pixel is determined with a resolution of nmz 3 . The 251

Page 11: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

11

horizontal resolution depends on the lens used; the highest magnification it is at the diffraction limit 252

of white light, of about 5.0 micron. In the present study, we have selected horizontal steps ( x ) 253

between 1 and 2 microns (Table 1). The actual precision in the spatial positioning is estimated to be 254

025.0 micron. As for data acquired with the laser profilometer, the cloud of points obtained are 255

regularly spaced and only some spurious points have been manually removed. The whole suite of 256

characteristics of the scanners devices and digital elevation models (spatial precision, resolution x , 257

noise on the data z ) used in this study are shown in Table 1. 258

As examples, two fault surfaces (Corona Heights fault, Figure 2 and Bolu fault, Figure 3) have been 259

selected to illustrate the topographic data acquired at all scales with the three techniques presented. 260

These fault surfaces are composed of many discrete slip surfaces delimiting bumpy lenses elongated 261

in the direction of slip (Figure 2). These multi-scale bumpy lenses give the wavy aspect of the fault 262

surfaces that are overprinted by fine linear polished striations and coarser corrugations generated by 263

abrasions (Figures 2 and 3). 264

265

2.3. Earthquake surface rupture data 266

In contrast to the exhumed fault scarps presented in Section 2.1, where the 3-D roughness is 267

characterized, surface rupture data at larger scales provide only 2-D measurements (Figure 4), since 268

full fault surfaces are not accessible for direct measurement. Fault trace roughness has been 269

previously analyzed (Scholz and Aviles, 1986; Wechsler et al., 2010). In the present study, high 270

resolution geological maps of large continental strike-slip earthquake surface ruptures, in various 271

geological settings, were analyzed using the dataset of Klinger (2010); see for an example Figure 5, in 272

Klinger (2010). For each event, the surface ruptures length, earthquake magnitude, and total 273

geological offset are provided in Table 2. 274

Of these three parameters, we note that the total slip of a fault is the least constrained one 275

because of the difficulty in finding markers of the displacement over long distances on strike-slip 276

continental faults. For Owens Valley fault (Beanland and Clark, 1994), the total offset was estimated 277

in the range 20-30 km. For the Haiyuan fault (Zhang et al., 1987) it is in the range from 10-15 km 278

Page 12: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

12

(Burchfiel et al. 1991) to 95±15 km (Gaudemer et al. 1995). For the Gobi-Altay fault, the total 279

displacement lies in the range of 2-20 km (Kurushin et al., 1997). For the Superstition Hills fault, the 280

total slip is still debated and, by comparison with the closeby San Jacinto Fault, the total displacement 281

is estimated to be less than 24 km (Sharp et al. 1967). For the Luzon earthquake, the cumulated 282

displacement of the fault is in the range of 50-200 km (Karig, 1983; Mitchell et al. 1986), and 283

Rigenbach et al. (1993) noticed that because of the activity of this fault since the beginning of the 284

Pleistocene, the total offset should be smaller than 200 km. The Hector Mine and Landers earthquakes 285

occurred on faults that are part of the East California Shear Zone that has a total geological offset 286

close to 65 km (Jachens et al., 2002). The individual geological total slip for Landers (3.1 to 4.6 km) 287

and Hector Mine (3.4 km) faults were estimated from the offset of magnetic anomalies (Jachens et al., 288

2002). Finally, Korizan (1979) and Zirkuh (1997) earthquakes occurred on the Abiz fault that 289

accommodates 60 km of right lateral motion along the Sistan suture between Iran and Afghanistan 290

(Berberian et al., 1999). 291

These rupture traces have been acquired using field cartography that allows mapping of the 292

geomorphologic traces of the rupture, combined with slip distributions and high resolution satellite 293

images (Klinger et al., 2005, 2006; Klinger, 2010). The actual point spacing is irregular and its 294

average is between 200 m and 1300 m. However, the precision of each point based on the combined 295

field observations and high resolution satellite images is close to the meter-scale (Klinger et al., 296

2005). This last detail is crucial for interpreting a multi-meter fine description of the roughness of the 297

rupture trace. 298

To avoid any bias due to local wiggles of the digitized rupture trace, the data set is re-sampled to 299

ensure consistent spatial sampling, independent of the length of each rupture (Klinger, 2010). This re-300

sampling procedure does not affect the Fourier transform computation and makes it possible to keep 301

the scaling information of the rupture traces (see Appendix A.2). 302

Geometric discontinuities that are commonly observed on the eight high resolution maps of large 303

continental strike-slip earthquake surface ruptures are fault azimuth changes or bends, and relay 304

zones, which are also referred to as jogs. These discontinuities reflect the multi-scale en-echelon 305

Page 13: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

13

pattern of the fault system, and range from a few hundred meters to several kilometers in size. Due to 306

the presence of abrupt steps associated with relay zones, which influence the Fourier transform and 307

therefore bias the roughness analysis, the ruptures traces are divided into individual segments (Figure 308

4). A reconstruction of the entire rupture trace can still be made when removing steps but the 309

information contained in the first order large scale segmentation has been removed. Note that for 310

some earthquake surface traces (Luzon, Superstition Hills, Hector Mine), no abrupt steps were 311

detected and the whole rupture trace can be directly analyzed. 312

313

3. Analysis of scaling properties of roughness data 314

In this section we detail the procedure to characterize the scaling properties of the scanned fault 315

surface topography. The same approach is followed for the digitized earthquake surface ruptures since 316

they are considered as rough profiles extracted along one direction of a fault surface. 317

First the original cloud of points (Figure 3) with irregularly spaced points was rotated, so that the 318

mean rupture surface was horizontal and parallel to major axis (X, Y). The 3-D set of points was 319

transformed in to 2-D ( YX , ) + 1-D ( Z ) data set where Z is the vertical direction, perpendicular to 320

the mean fault plane (Renard et al., 2006; Sagy et al., 2007; Candela et al., 2009; Zhanyu et al., 321

2010). A set of parallel cuts was taken through the cloud of points to obtain a series of thin bands of 322

points striking at an angle from the X axis. Then, each band of points was projected to obtain a 323

series of profiles with irregularly spaced points. The thickness of the bands of points is closely related 324

to the average spacing (X, Y) of the original raw data. Linear interpolation on a regular spacing is 325

performed independently on all profiles to yield a set of heights iXh , function of the coordinate 326

iX along the cut. The regular spacing taken for the linear interpolation has been systematically 327

chosen to be twice as large as the average irregular spacing of the original profiles. 328

To describe the scaling properties of these rough profiles (Figure 1), we search for possible spatial 329

correlations of the height fluctuations. Along each profile, we computed auto-correlation functions. If 330

the auto-correlation function of a rough profile is a power law and scales as 331

Page 14: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

14

Hxxxhxh 2)(),( , then the rough profile is self-affine with H the Hurst exponent, if 332

multi-affinity is excluded. One way to estimate the Hurst exponent is to compute the Fourier 333

transform. This method is well-suited and robust for recognizing and characterizing self-affine 334

roughness, as demonstrated by Candela et al. (2009). The Hurst exponent H can be estimated from the 335

Fourier power spectrum, which has a power law form for a 2-D self-affine profile (Barabàsi and 336

Stanley, 1995; Meakin, 1998). The steps in the procedure to compute the Fourier power spectrum of 337

each profile are as follows: 1) First, removing of the residual drift in the signal is performed in order 338

to avoid any ramp artifact for the Fourier analysis (see Schmittbuhl et al., 1995b for a quantitative 339

analysis of this trend artifact). Indeed, adjusting the reference mean fault plane is always hard to 340

determine exactly a priori. In addition, even if corrections are possible a posteriori, trend suppression 341

is non-trivial and complex. All length scales are involved in the faulting process, even very large 342

ones, and therefore suppressing macroscopic information may influence the scale invariance analysis 343

at large scales. For the following analyses, and as suggested by Schmittbuhl et al. (1995a), the trend 344

has been defined simply as the line fit through the first and last point. 2) In order to ensure that there 345

are no step functions at the end of the finite window, we apply a 3% cosine taper. Note that a taper 346

function of 5 and 10% has been tested and the results appear robust. 3) The Fourier power spectrum 347

)(kP , i.e. the square of the modulus of the Fourier transform, is calculated as a function of 348

wavenumber k . 4) The Fourier power spectrum is normalized by dividing the power at each 349

wavenumber by the length of the profile. 5) Each cloud of points is computed as a whole by stacking 350

and averaging all 2-D Fourier transforms to reduce the noise associated with individual profiles. In 351

other words, for each fault patch, power spectral estimates with regularly spaced wavenumbers is 352

obtained by averaging the power spectra of the individual profiles in a geometric progression (Figure 353

5). 354

As an example, the computed Fourier power spectra along the slip direction and perpendicular to 355

it from patches of the Corona Heights fault surface acquired with each device (LiDAR, laser 356

profilometer, WLI) are displayed in Figure 6. The uncertainty in the average spectral values obtained 357

for each fault patch can be estimated following the method of Bendat and Piersol (1986). A one sigma 358

Page 15: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

15

confidence interval for the spectral power is given by: 1

)(ˆ)(

1

)(ˆ

s

s

s

s

n

kPnkP

n

kPn with 359

1)( kynkn ys , where )(kP and )(ˆ kP are the actual and calculated spectral power, 360

respectively; yn and sn are the total number of profiles spaced a distance y apart perpendicularly 361

to the profile direction, and the number of independent profiles used to calculate )(ˆ kP , respectively. 362

Note that sn depends on scale. For the largest wavenumbers, there are many more independent 363

estimates of the total spectral power, and hence the error estimate is smaller than at the smallest 364

wavenumbers (see Figure 6). 365

Note that for wavelengths below 50 mm, the fault surfaces we scanned with the LiDAR are so 366

smooth that at this small scale, their spectral power falls within the range of those of the flat plate we 367

use as a planar reference surface (Figure 6). That means that even if our LiDAR surface 368

measurements were acquired at a spatial length scale resolution of mm1 to m10 (cf. Section 2.2), 369

total power estimates of the surfaces are accurate only between 0.05-10 m scales. In contrast, the laser 370

profilometer data and WLI data can be considered to be essentially noise free since their inherent 371

noise level falls well below the magnitude of the fault surface topography (Figure 6). 372

When plotting the average power spectrum as a function of wavenumber in a log-log space, a 373

self-affine function reveals a linear slope, which is itself a function of H through HkCkP 21)( 374

(with C the pre-factor). Taking into account the possible uncertainties in the spectral power as 375

previously described for our entire data set of fault surface patches, the upper limit of the error bar on 376

the Hurst exponent estimated using a least square method is equal to 05.0 and does not vary 377

significantly with the wavenumber. Due to the fact that only one rough profile for each rupture trace 378

is analyzed (Figure 4), the noise in the spectrum is higher compared to the fault surface patch 379

(constituted of an average of a multitude of profiles). For a single power spectral estimate, the 380

standard deviation is equal to the mean (Press et al., 2007). So that the upper limit of the error bar in 381

the estimated Hurst exponent of the eight rupture traces analyzed is equal to 1.0 . 382

Page 16: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

16

383

4. Fault roughness results 384

4.1. Two universal Hurst exponents 385

In order to extract the Hurst exponents characterizing the scaling behavior of the faults 386

roughness, we have fit the linear part of each averaged spectrum obtained in the slip direction and 387

perpendicular to it for each individual fault patch (Figure 6). Because we are ultimately interested in 388

the variability of pre-factors, as discussed in the next section, we have worked separately on each fault 389

patch with each type of instrument instead of calculating average spectra for the full surface. We 390

have focused our analysis of fault surface roughness between the largest scale accessible with the 391

LiDAR, i.e. approximately m10 , and to the scale of mm05.0 accessible by the WLI. However, the 392

WLI gives access to scales down to 1 micron. 393

Figure 7 compiles all Hurst exponents derived from each individual fault patch (41 sub-surfaces 394

in total) in the direction of slip and perpendicular to it. The five faults scanned can be characterized by 395

two different global Hurst exponents between mm05.0 and m10 (see Table 1): 07.058.0// H 396

in the direction of slip and 04.081.0 H perpendicular to the slip. We will refer to 6.0// H 397

and 8.0H afterwards. This result highlights the fact that, for one direction (parallel or 398

perpendicular to slip), the relative amplitude of short and large wavelengths remains identical for the 399

whole data set. Fault surfaces with a Hurst exponent smaller or larger than one standard deviation of 400

their distribution ( 07.0 in slip direction and 04.0 normal to slip) exist, but all Hurst exponents 401

fall within two standard deviations (Figure 7). Consequently, our global Hurst exponents appear to 402

appropriately characterize the scaling behavior of the five faults studied when taking into account the 403

fluctuations indicated in Table 1. 404

A systematic bending of the Fourier power spectra (along both the parallel and the perpendicular 405

slip directions) at the length scale mm05.0 is observed on the WLI data (Figure 6, 9). This change of 406

regime occurs at larger length scales relative to the expected WLI resolution, and we suspect that it 407

could be the hallmark of the transition between two physical processes appearing at the grain scale. 408

Page 17: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

17

Chen and Spetzler (1993) suggest that a characteristic length scale appearing at the grain scale is due 409

to a change in the dominant mode of deformation from small scale intergranular cracking to 410

intraganular cracking at large scale. The same observation and interpretation was made by Meheust 411

(2002) on tensile cracks. Sagy et al. (2007) also suggested a change in behavior at the sub-centimeter 412

scale for large-slip faults. However, it is beyond the scope of this work to quantitatively characterize 413

this possible characteristic length scale and we limit our scaling analysis to scales larger than 414

mm05.0 . 415

416

4.2. Pre-factor variability 417

In Section 4.1, we focused the analysis on the slope of the power-law, which appears to be 418

constant along slip and perpendicular to it for each fault surface. In addition, for each scanned fault 419

patch, the along slip direction is smoother than perpendicular to it. This is seen by the lower position 420

of power spectrum amplitudes at each scale (Figure 6). More precisely, the power-law relationships 421

between )(kP and k have both lower power-law exponents and pre-factors. However, our results 422

point out that the pre-factor derived from each fault patch is variable along the fault scarps. Here, we 423

illustrate this variability on the data acquired with the LiDAR on the Corona Heights fault, along the 424

slip direction. A significant vertical shift is observed on the Fourier power spectra corresponding to 425

the six different patches of the Corona Heights fault (Table 1, Figure 8). The slopes, however, are 426

similar (Table 1, Figure 7). We emphasize that, even if for some fault patches it is not the case, for 427

other patches this spatial variability on the pre-factor is clearly larger than the size of the error bars in 428

the average spectral values (see Figure 8, for example between Corona-A and Corona-E). 429

As proposed by Power et al. (1987) and Power and Tullis, (1989), who have observed a similar 430

vertical shift in their Fourier power spectra, a tempting explanation of the variability of the pre-factor 431

within a same fault surface is that the finite displacement accumulated by rougher fault patches (or 432

segments with larger pre-factors) is smaller than for the smoother fault patches. In other words, it is 433

possible that the spatial heterogeneity on the pre-factor illustrated in Figure 8 highlights variable 434

accumulated displacement (probably much less important than the total offset recorded by the fault 435

Page 18: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

18

zone) on each individual segment (or fault patch) constituting the whole fault zone. However, as 436

emphasized by Power and Tullis (1989), this possibility is difficult to quantify because the total 437

displacement for sub-parallel individual surfaces (which represent a part of the total offset of the fault 438

zone) cannot be observed with certainty in the field. Alternative explanations for the difference in 439

surface characteristics would include formation of the individual fault patches at different conditions 440

(depth, confining pressure, temperature, strain rate) or the presence of different initial heterogeneities 441

in the rock before deformation. It is also possible that the variability of roughness reflects the natural 442

variability of an underlying stationary pre-factor distribution. 443

In order to highlight the specific trend of each fault in the spectral domain, we have averaged the 444

similar spectra (that means spectra with an approximately identical slope but with a slightly different 445

vertical position) obtained from each scanner device. Each curve on Figure 9 therefore represents an 446

average of similar spectra obtained for multiple fault patches. In this way, this technique gives a 447

smoother spectrum that represents the global self-affine character of the entire fault surface at the 448

specific scales accessible by each device, while preserving good wavelength resolution. However, as a 449

direct consequence of the previously described spatial variability of the pre-factor along the fault 450

surface (Figure 8), the global spectra obtained for each device are vertically shifted in Figure 9. 451

Consequently, even if each global spectrum presents a similar slope, in most cases it is difficult to 452

connect them all together with a unique linear trend (Figure 9). It is worthwhile to point out here that 453

when we compute the scaling from the LiDAR measurements, we average all the 2-D power spectra 454

of all individual rough profiles extracted from different fault patches (Figure 5, 8). Each of these 455

profiles has a different pre-factor but the average gives one specific pre-factor characterizing the 456

global 3-D geometrical property of the fault surface at the LiDAR scale. At smaller wavelengths, with 457

another device (i.e. laser profilometer or WLI), we select several sub-regions with one given average 458

pre-factor among the whole population that we explored at larger scale (i.e. with the LiDAR) but not 459

necessarily equal to the average pre-factor of the large scale measurements (see Figure 8, 9). In other 460

words, with the laser profilometer or the WLI, we expect to sample rougher or/and smoother sub-461

regions of the fault (i.e. with a pre-factor magnitude that is, respectively larger or smaller) than the 462

Page 19: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

19

average behavior recorded by the LiDAR by combining data over the full surface. As a consequence, 463

it is clear that for the same fault scarp, the specific global spectra calculated for each device by 464

stacking similar spectra obtained for several fault patches, might be vertically shifted in some cases as 465

observed in Figure 9. Finally, this variability in the pre-factor: (i) explains why the Hurst exponents 466

have to be necessarily calculated separately for each device (cf. Section 4.1), and (ii) implies that it is 467

not necessary to invoke a change in slope to connect the vertically shifted global spectra in Figure 9. 468

To summarize, our analysis highlights that the roughness of the five fault scarps studied can be 469

characterized over more than 5 decades of length scales (between mm05.0 and m10 ) by two 470

universal Hurst exponents in the direction of slip and perpendicular to it even if it is difficult to point 471

out a single pre-factor in both directions. This description of the fault surfaces is independent of the 472

geological context, i.e. lithology (rhyolite, chert and limestone) and tectonic regime (strike-slip, 473

oblique and normal). Moreover, no clear relationship is observed between the range of pre-factor 474

magnitude extracted from the global spectra shown in Figure 9 for each fault (see Table 3 for a 475

complete list of the maximum and minimum pre-factors of each faults), and the slip estimated for the 476

fault zone (see Figure 10). 477

478

4.3. Roughness of large continental earthquakes surface ruptures 479

In Figure 11, a compilation of the roughness results is provided for the eight surface ruptures. 480

Results of the Fourier power spectrum analysis are shown for the individual segments (bounded by 481

abrupt steps), and the whole rupture trace. For each surface rupture trace, each individual segment has 482

a roughly identical self-affine exponent (see also Table 2) at large scales (i.e. above the regime 483

controlled by data re-sampling – see Section 2.3). Moreover, the profiles obtained for the entire 484

“reconstructed” rupture traces represent an average of the roughness over all the individual segments 485

and keep an approximately identical self-affine exponent (Figure 11). 486

The variability in the power spectra amplitude (i.e. the pre-factor) of the profiles extracted from 487

the exhumed fault scarps (see Section 4.2) is also observed between each individual segment of the 488

rupture traces. More precisely, along the rupture traces, individual segments display variability in the 489

Page 20: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

20

roughness amplitude but keep the same self-affine scaling properties of amplitude versus length scale 490

(Figure 11). Since the 2-D roughness scaling properties of the reconstructed rupture traces correspond 491

to a sum of those of the individual segments, it is to be expected that their spectra fall between the 492

smoothest and the roughest short segments (Figure 11). In this way, the range of pre-factors inferred 493

by the individual segments can be interpreted as a typical fluctuation or error in the estimated pre-494

factor of the “reconstructed” rupture traces. 495

On the same log-log graph (Figure 12a), a stack of all the spectra calculated on the whole rupture 496

traces complements Figure 11 and emphasizes that all of the surface rupture data can be described by 497

a unique self-affine exponent of 1.08.0 . Another interesting result highlighted in Figure 12a is 498

that, even though the eight rupture traces analyzed clearly sample variable geological settings, the 499

same self-affine exponent fits all of the data best. In particular, even though the surface rupture traces 500

are related to fault zones which have accumulated a large range of finite geological offsets (see Table 501

2), no trend is revealed between this parameter and the 2-D roughness scaling of the rupture traces. 502

Indeed, taken into account both (i) the average pre-factor of each whole rupture traces and (ii) their 503

respective typical fluctuations extracted from the corresponding individual segments, no correlation 504

appears with the finite geological offset (Figure 12b), admitting however that the standard deviation 505

in the data is quite large and could hide the physical correlation, if any. 506

507

5. Discussion 508

5.1. Can surface measurements of faults be correlated with earthquake processes at depth? 509

Firstly we discuss how our measurements, performed on surface outcrops, could be relevant 510

to faulting processes that occur at crustal depths up to 15 km. On the five faults we have studied, two 511

of them (Magnola and Dixie Valley) have a mainly normal component and their roughness has 512

therefore recorded processes at work in the first kilometers of the crust. For the three other strike-slip 513

faults, the roughness has recorded only shallow depth faulting processes. Dixie Valley and Bolu 514

outcrops have recorded the surface termination of major earthquakes in the last century. For all faults, 515

we find roughness exponents in the range 0.6 – 0.8. This can be compared to a recent study of the 516

Page 21: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

21

roughness of exhumed faults in the Sierra Nevada and Italian Alps (Griffith et al. 2010, Bistacchi et 517

al. 2011). In these studies, the roughness of fault traces was characterized on outcrops for which the 518

deep origin (c.a. 10 km depth) of the faulting process could be identified because of the presence of 519

pseudotachylites and specific mineral assemblages. Griffith et al. (2010) measured Hurst exponents 520

with a quite large variability, in the range 0.4 – 1. Interestingly, Bistacchi et al. (2011) report Hurst 521

exponents in the range 0.6 – 0.8 for spatial scales in the range 0.5 mm – 500 m, similar to those 522

observed in our data and suggesting that the roughness property of faults is also maintained at depth. 523

524

5.2. Implications of the self-affine scaling of the fault roughness on the machinery of 525

earthquakes 526

One of the main contributions of the present study is to provide a robust and realistic fault 527

geometry model, which is still currently missing. Here, we demonstrate that one way to describe fault 528

surfaces with few parameters is to consider them as self-affine. Using our realistic geometry of rough 529

fault surfaces in seismic rupture dynamic models will provide new keys to move forward towards a 530

better understanding of the fundamentals of earthquake behavior. In this sense, based on self-affine 531

geometrical model of fault surfaces as highlighted in our present work, recent numerical and 532

theoretical studies have shown results in agreement with seismological observations. Indeed, 533

quantification of roughness has been used to infer the scaling of the stress and slip variations along 534

the fault plane after an earthquake (Schmittbuhl et al., 2006; Candela et al., 2011a, 2011b). Fault 535

roughness is also useful for calculating the expected seismic radiation (Dunham, 2011). 536

537

5.3. Reconciling the Hurst exponent of the scanned fault surfaces and rupture traces 538

The rupture trace geometry of strike-slip earthquakes should correspond to the extrapolation of 539

the fault surface roughness sampled on fault scarps along the slip direction and the Hurst exponent 540

should be close to 0.6. However, our results highlight a small difference in the Hurst exponents of the 541

eight rupture traces compared to the average trend in slip direction of the five fault surfaces scanned. 542

The spatial correlations of the rupture trace irregularities are characterized by 1.08.0 RH and 543

Page 22: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

22

pre-factors magnitudes, close to those in the direction perpendicular to slip, collected on the five fault 544

surfaces that were scanned. We propose that the roughness increase at the rupture trace scale is due to 545

the fact that the slip direction is not strictly sampled as it is the case for fault surfaces scanned. Slip 546

distributions of strike-slip earthquakes show that at some locations a vertical component could be 547

significant (Florensov and Solenko, 1965; Kurushin et al., 1997). Therefore, at the surface rupture 548

scale, a slight vertical slip component along the whole rupture trace of the strike-slip earthquake 549

would explain how we could sample a roughness oblique to the slip direction. 550

This interpretation is investigated in Figure 13. The Hurst exponents and the pre-factors of a 551

synthetic anisotropic self-affine surface (see Appendix A.1), computed using a Fourier based method 552

(Candela et al., 2009), and with a roughness root-mean-square standard deviation that scales as 553

6.0005.0 LRMS in slip direction and 8.0015.0 LRMS normal to it, were calculated on series of 554

2-D profiles extracted at an angle from the slip direction. The same procedure was performed on a 555

scanned patch of the Corona Heights fault (Corona-A in Table 1). A clear similarity is observed in the 556

angular variability of the roughness exponent computed for the synthetic surface and the natural fault 557

patch (considered as representative of our set of fault surfaces). Both surfaces expose the same non-558

linear dependence of the Hurst exponent with the azimuthal direction of profile extraction ( ), as 559

pointed by Renard et al. (2006) and Candela et al. (2009) using different statistical tools. When 560

departing a few degrees from the direction of the smallest exponent, i.e. the slip direction, the Hurst 561

exponent sampled is already very close to the largest exponent, i.e. the direction normal to slip 562

(Figure 13). This effect could explain the slight difference in the roughness exponents observed 563

between the surfaces ruptures and fault surface scanned along slip. In the same way, this non-564

alignment of the ruptures traces with the slip direction, could also explain that their pre-factors 565

magnitudes are close to the maximum sampled for the fault surface scanned. 566

Moreover, the traces of the ruptures show a structural complexity underlined by relay zones 567

(compressive and extensive jogs), and bends, which could locally accommodate a significant vertical 568

component at many locations along the fault (Klinger et al., 2006). In this case, the rupture trace 569

morphology corresponds to a combination of topography along slip direction and normal to it, which 570

Page 23: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

23

might partially explain the increase of the measured Hurst exponent and pre-factor compared to those 571

of the along-slip fault surfaces that were scanned. 572

Taking into account the previous arguments, we propose that the self-affine regime observed at the 573

outcrop scale is also present at the map-scale of earthquake rupture traces. In other words, even 574

though there is a gap of data between the scanned fault surfaces and the rupture traces that remains to 575

be investigated and the traces do not directly constrain the slip-parallel geometry, a unique self-affine 576

geometrical model could fit our measurements from the WLI scale to the map-scale of rupture traces. 577

Indeed, considering the variability of the pre-factor observed both on the scanned fault scarps (see 578

Section 4.2) and the rupture traces (Section 4.3), it seems reasonable to directly connect both data sets 579

by a unique line covering 9 decades of length-scales (Figure 14). As already mentioned, analyzing the 580

roughness of fault traces in Italian Alps, Bistacchi et al. (2011) have recently reported Hurst 581

exponents in the range 0.6 – 0.8 for spatial scales in the range 0.5 mm – 500m and therefore bridging 582

our gap of data. 583

584

5.4. Pre-factor variability independent of displacement 585

In Section 4.2, by combining the measurements acquired with various scanner devices and 586

covering complementary scales, we have shown that no clear relationship exists between the range of 587

roughness amplitude (i.e. the range of pre-factor magnitude) of each exhumed fault surface and the 588

slip estimated for the scanned fault surface (Figure 10). Note here that, following Sagy et al., (2007), 589

even if we focus on the smoothest sub-surfaces of each fault (i.e. the lower bound of the range of pre-590

factors), no trend is revealed with the estimated offset accommodated (see Figure 10). Other recent 591

studies have noted changes in fault roughness and damage parameters as a function of slip for faults 592

spanning a range of offsets shifted from the dataset here (Sagy et al., 2007; Mitchell and Faulkner, 593

2009; Savage and Brodsky, 2011; Brodsky et al., 2011). Sagy et al. (2007) noticed a difference for 594

faults that have slip more than 10 m versus those that slip less than 1 m. The fault roughness data 595

presented in the present study only spans offsets that are within the large-slip faults population by 596

these criteria. The lack of a discernible evolution signal is consistent with the suite of data of Brodsky 597

Page 24: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

24

et al. (2011) that also shows that smoothing is a very weak process once >10 m of offset is 598

accumulated. 599

Another important detail of our roughness scaling analysis is that no correlation was observed 600

between the finite displacement by the fault zone hosting the rupture trace and its roughness (see 601

Section 4.3 and Figure 12b). This observation is consistent with the study of Klinger (2010), who 602

showed that the correlation between the characteristic fault segment length and the thickness of the 603

seismogenic crust is maintained, independently of the slip accumulated. In addition, we suggest in this 604

study that a consistent spatial organization persists over the entire range of length scales accessible 605

(i.e. from several hundred meters to ~50 km), independent of the total geological offset. Note here 606

that in following the reasoning of Klinger (2010), a specific length scale should appear at 607

approximately 20 km, i.e. the thickness of the seismogenic crust. However, because of the lack of 608

sufficient frequency content between 20 km and 50 km, this probable characteristic length scale is not 609

clearly revealed by our analysis. 610

Smoothing of fault geometry has been suggested by other studies (Wesnousky, 1988; Manighetti 611

et al., 2007), and is thought to result from surface roughness of faults being inversely related to their 612

total displacement. Such smoothing of a fault trace has primarily been studied by examining large 613

scale geometrical asperities, such as step-overs of several kilometers. In our study, given that we 614

remove these first-order geometric discontinuities to perform our Fourier transform analysis, we lose 615

these hallmarks of the fault maturity. 616

The notion of "geometric regularization" or “maturity” refers to the intuitive idea that the fault 617

zone geometry becomes more and more flat during the successive slips by abandoning or smoothing 618

the complexity of the initial structures (segments). Ben-Zion and Sammis (2003) recommended that it 619

is necessary to separate the abandoned structural units from those which actively participate in the 620

accommodation of the slip to reveal a possible regularization of the geometry of the fault zone with 621

slip. It is therefore important to emphasize that both the fault scarps scanned on the field and the 622

rupture traces are markers of the morphology of the active structures of fault zones. Our results thus 623

demonstrate that the active portions of faults from the spatial scale of micrometer up to at least the 624

Page 25: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

25

thickness of the seismogenic crust preserve a complex geometry during successive displacements. In 625

other words, the complexity of an active fault surface is spatially organized following a self-affine 626

geometrical model, independently of the lithology and the tectonic regime. Our observations, as those 627

of Brodsky et al. (2011) and Klinger (2010), suggest that a re-roughening mechanism is active in the 628

fault zone to maintain the geometrical complexity during successive slips. For example, as proposed 629

by Klinger (2010), rupture might naturally branch on preexisting secondary faults as a result of the 630

dynamic stress build-up ahead of the rupture (Poliakov et al., 2002; Bhat et al., 2004). This effect 631

could explain the persistence of some level of complexity and prevent the complete smoothing of the 632

fault geometry. 633

634

5.5. Scale-free process at the origin of fault roughness 635

Previous studies on natural fault surfaces have shown that the slope of the Fourier spectrum 636

could change from large to small scales (e.g. Lee and Bruhn, 1996), implying that different processes 637

may be involved during the generation of surface fault textures at different spatial scales. In fact, any 638

inflexion of the spectrum would correspond to a characteristic scale. Such a crossover scale could be 639

interpreted as the transition between physical processes, as is the case for dissolution surfaces such as 640

stylolites (Renard et al., 2004). In the present study, five fault surfaces in different geological settings 641

have been scanned with three different scanner instruments spanning complementary length scales. 642

Considering the variability of the pre-factor (see Section 4.2), an anisotropic mono-affine geometric 643

model characterized by two different global Hurst exponents 6.0// H and 8.0H best fits the 644

entire data set between m10 and mm05.0 . In our data set, because the slopes follow a consistent 645

trend, and considering the level of noise, no variation of the Hurst exponent could be detected 646

between m10 and mm05.0 . This observation could be evidence that a unique scale-free process 647

could be at the origin of the scanned fault scarps. 648

In contrast to mode I fracture roughness for which a lot has been done in physics and mechanics 649

communities, few studies have focused on the processes at work in the generation of the shear surface 650

roughness. Since the 1990s, increasing experimental evidence showed that the roughness exponent for 651

Page 26: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

26

fractures surfaces had a universal value of about 0.80 (Bouchaud et al., 1990; Bouchaud, 1997). 652

Simultaneously with these experiments, theoretical and numerical works have been produced at a 653

steady rate with the aim of modeling the fracture propagation in order to identify the origin of the self-654

affine scaling of fracture roughness (Bouchaud, 1997, Alava et al., 2006; Hansen & Schmittbuhl, 655

2003; Bonamy and Bouchaud, 2011). The common denominator of all this work is to model fracture 656

propagation as a network of elastic beams, bonds, or electrical fuses with random failure thresholds 657

and subject to an increasing external load. In other words, in order to reproduce the self-affine 658

roughness exponent experimentally observed, the fracture propagation is assimilated as a damage 659

coalescence process in a heterogeneous material. The damage becomes localized by long-range elastic 660

interaction between multi-scale cracks. In this sense, we propose that elastic interactions related to 661

linkage of many discrete slip surfaces, controlling the generation of the multi-scale bumpy lenses 662

observed on the scanned fault outcrops, could be the scale-free process at the origin of fault 663

roughness. Indeed, even if shear cracks involve significantly higher energy release rates than tensile 664

cracks (Atkinson, 1991), some similarities in the elastic influences on the stress field at the crack tip 665

can be obtained theoretically (Gao and Rice, 1986; Gao et al., 1991; Atkinson, 1991; Schmittbuhl et 666

al., 2002). 667

Even if a unique self-affine regime is maintained up to the map-scale of the rupture traces, it 668

remains difficult to support the fact that their roughness is also controlled by the process at work for 669

the scanned fault scarps. However, it is noteworthy that the rupture traces seem to be also formed by 670

multi-scale segments (see for example the Figure 5 in Klinger, 2010) whose growth and coalescence 671

could be also controlled by long range elastic interactions. 672

673

6. Conclusion 674

The roughness properties of fault surfaces is summarized in Figure 14, where the Fourier power 675

spectra of the Corona Heights fault surfaces perpendicular to the slip direction and the spectra 676

obtained for the continental earthquake surface ruptures are plotted together. First, even if it is 677

difficult to point out a unique power-law fit due to the pre-factor variability, a single anisotropic self-678

Page 27: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

27

affine model ( 6.0// H and 8.0H ) best fit our 3-D scanned fault measurements from 679

mm05.0 scale to m10 . Secondly, an identical self-affine behavior is consistent with the map-scale 680

of the 2-D rupture traces. In order to connect both data sets while considering the variability of the 681

pre-factor, we propose that a unique self-affine geometrical model is maintained over nine decade of 682

length scales (between 0.05 mm and at least the thickness of the seismogenic crust, i.e. ~20 km). Even 683

if the lack of data between the scanned fault surfaces and the rupture traces remains to be investigated, 684

recent works of Bistacchi et al. (2011) bridging the length scale gap appear to support our 685

interpretation. 686

In addition, even if in both cases, for the scanned fault surfaces and the ruptures traces, we have 687

focused our analysis on the active portion of the fault zone, it appears that once a small amount of 688

offset has been achieved, the geometric complexity is maintained regardless of the amount of further 689

slip accommodated. Consequently, we propose that processes that create this roughness and processes 690

that destroy it must reach a dynamic equilibrium. 691

692

Acknowledgments: This study was supported by the Agence Nationale pour la Recherche grant ANR-693

JCJC-0011-01, ANR ECCO TRIGGERLAND and ANR RiskNat SUPNAF. Field work was 694

supported in part by NSF Grant EAR-0711575. The authors are very grateful to Michel Bouchon and 695

Fabrice Cotton for encouragements and scientific stimulations. William L. Power is really thanked for 696

his detailed and constructive comments on an early version of this work. The first author gratefully 697

acknowledges Andrew P. Rathbun and Megan Avants for their constructive comments that helped to 698

improve the content and clarity of the manuscript, and James D. Kirkpatrick for fruitful discussions 699

concerning various aspects of this paper. 700

701

702

703

704

Page 28: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

28

References 705

Alava, M. J., Nukala, P. K. V. V., Zapperi, S., 2006. Statistical models of fracture, Advances in 706

Physics, Vol. 5, No. 3-4, 349-476, doi: 10.1080/00018730300741518. 707

Angheluta, L., Candela, T., Mathiesen, J., Renard, F., 2011. Effect of surface morphology on the 708

dissipation during shear and slip along a rock-rock interface that contains a visco-elastic core. Pure. 709

Appl. Geophy, doi 10.1007/s00024-011-0272-8. 710

Aochi, H., Madariaga, R., 2003. The 1999 Izmit, Turkey, Earthquake: Nonplanar fault structure, 711

dynamic rupture process, and strong ground motion. Bull. Seismol. Soc. Am., 93, 1249-1266. 712

Atkinson, B., 1991. Fracture Mechanics of Rock, Academic, San Diego, California. 713

Barka, A., 1996. Slip Distribution along the North Anatolian Fault Associated with the Large 714

Earthquakes of the Period 1939 to 1967. Bull. Seismol. Soc. Am., 86, 1238-1254. 715

Barabàsi, A.-L., Stanley, H.E., 1995. Fractal Concepts in Surface Growth. Cambridge University 716

Press. 717

Ben-Zion, Y. and Sammis, C.G., 2003. Characterization of fault zones, Pure Appl. Geophys. 160, 718

677–715, doi: 10.1007/PL00012554. 719

Beanland, S., and Clark, M., 1994. The Owens Valley fault zone, eastern California, and surface 720

rupture associated with the 1872 earthquake: U.S. Geological Survey Bulletin 1982, 29 p. 721

Berberian, M., Jackson, J.A., Qorashi, M., Khatib, M.M., Priestley, K., Talebian, M., and Ghafuri-722

Ashtiani, M., 1999. The 1997 May 10 Zirkuh (Qa’enat) earthquake (Mw 7.2): faulting along the 723

Sistan suture zone of eastern Iran, Geophysical Journal International, 136, 671-694. 724

Bhat, H., Dmowska, R., Rice, J., Kame, N., 2004. Dynamic slip transfer from the Denali to 725

Totschunda faults, Alaska: Testing theory for fault branching. Bull. Seismol. Soc. Am., 94(6B), 726

S202–S213, doi: 10.1785/0120040601. 727

Bistacchi, A., Massironi, M., Menegon, L., 2010. Three-dimensional characterization of a crustal-728

scale fault zone: The Pusteria and Sprechenstein fault system (Eastern Alps), Journal of Structural 729

Geology, 32, 2022-2041. 730

Page 29: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

29

Bistacchi, A., Griffith, W. A., Smith, S. A., di Toro, G., Jones, R., Nielsen, S., 2011. Fault roughness 731

at seismogenic depths from LIDAR and photogrammetric analysis, Pure and Applied Geophysics, doi 732

10.1007/s00024-011-0301-7. 733

Bonamy, D., Bouchaud, E., 2011. Failure of heterogeneous materials: A dynamic phase transition? 734

Physics Reports, 498, 1-44, doi:10.1016/j.physrep.2010.07.006. 735

Bouchaud, E., G. Lapasset, and J. Planès, 1990. Fractal dimension of fractured surfaces: A universal 736

value?, Europhys. Lett., 13, 73. 79. 737

Bouchaud, E., 1997. Scaling properties of cracks, J. Phys. Condens. Matter, 9, 4319 – 4344. 738

Bouchon, M., 1997. The state of stress on some faults of the San Andreas system as inferred from 739

near field strong motion data. J. Geophys. Res.-Solid Earth, 102, 11731-11744, 740

doi:10.1029/97JB00623. 741

Bouchon, M., Karabulut, H., Bouin, M.-P., Schmittbuhl, J., Vallée, M., Archuleta, R., Das, S., 742

Renard, F., Marsan, D., 2010. Faulting characteristics of supershear earthquakes. Tectonophysics, 743

doi:10.1016/j.tecto.2010.06.011. 744

Brodsky, E.E., Gilchrist, J.G., Sagy, A., Colletini, C., 2011. Faults Smoothly Gradually as a Function 745

of Slip, Earth & Planetary Science Letters, 302, 185-193. 746

Brown, S. R.; Scholz, C. H., 1985. Broad bandwidth study of the topography of natural rock surfaces. 747

J. Geophys. Res. -Solid Earth, 90, 12575-12582. 748

Burchfiel, B.C., Peizhen, Z., Yipeng, W., Weiqi, Z., Fangmin, S., Qidong, D., Molnar, P., Royden, L., 749

1991. Geology of the Haiyuan fault zone, Ningxia-Hui autonomous region, China, and its relation to 750

the evolution of the northeastern margin of the Tibetan plateau. Tectonics, 10, 1091-1110. 751

Candela, T., Renard, F., Bouchon, M., Marsan, D., Schmittbuhl, J., Voisin, C., 2009. Characterization 752

of fault roughness at various scales: implications of three-dimensional high resolution topography 753

measurements. Pure. Appl. Geophy., 166, 1817-1851. 754

Candela, T., F. Renard, J. Schmittbuhl, M. Bouchon, Brodsky, E.E., 2011a. Fault slip distribution and 755

fault roughness, Geophysical Journal International, doi: 10.1111/j.1365-246X.2011.05189.x. 756

Page 30: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

30

Candela, T., F. Renard, M. Bouchon, J. Schmittbuhl, Brodsky, E.E., 2011b. Stress drop during 757

earthquakes: effect of fault roughness scaling, Bull. Seismol. Soc. Am., 101, doi: 758

10.1785/0120200298. 759

Carcaillet, J., Manighetti, I., Chauvel, C., Schlagenhauf, A., Nicole, J. M., 2008. Identifying past 760

earthquakes on an active normal fault (Magnola, Italy) from the chemical analysis of its exhumed 761

carbonate fault plane. Earth Planet. Sci. Lett., 271, 145–158, doi:10.1016/j.epsl.2008.03.059. 762

Chen, G., Spetzler, H. A., 1993. Topographic characteristics of laboratory-induced shear fractures. 763

Pure Appl. Geophys., 140, 123–135. 764

Chester, F. M., Chester, J. S., 2000. Stress and deformation along wavy frictional faults, Journal of 765

Geophysical Research, 105, 23421-23430. 766

Chester, J. S., Fletcher, R. C., 1997. Stress distribution and failure in anisotropic rock near a bend on a 767

weak fault, Journal of Geophysical Research, 102, 693-708. 768

Dieterich, J. H., Smith, D. E., 2009. Nonplanar faults: mechanics of slip and off-fault damage, Pure. 769

Appl. Geophys., 166, 1799–1815, doi:10.1007/s00024-009-0517-y. 770

Dozer, D. I., 1986. Earthquake processes in the Rainbow Mountain-Fairview Peak-Dixie Valley, 771

Nevada, Region 1954-1969. J. Geophys. Res.-Solid Earth, 91, 12572-12586. 772

Dunham, E. M., D. Belanger, L. Cong, and J. E. Kozdon, 2011. Earthquake ruptures with strongly 773

rate-weakening friction and off-fault plasticity, 2: Nonplanar faults, Bulletin of the Seismological 774

Society of America, 101(5), 2308-2322, doi:10.1785/0120100076. 775

Dysthe D. K., Podladchikov, Y., Renard, F., Feder, J., Jamtveit, B., 2002. Universal Scaling in 776

Transient Creep. Phys. Rev. Lett., doi: 10.1103/PhysRevLett.89.246102. 777

Feder J., 1988. Fractals, Plenum, New York. 778

Faulkner, D. R., Jackson, C. A. L., Lunn, R. J., Schlische, R. W., Shipton, Z. K., Wibberley, C. A. J., 779

and Withjack, M. O., 2010. A review of recent developments concerning the structure, mechanics and 780

fluid properties of fault zones. Journal of Structural Geology, 32, 1557-1575. 781

Florensov, N. A., Solonenko, V. P., 1965. The Gobi-Altai Earthquake, 424 pp., U.S. Dep. of 782

Commer., Washington, D. C. 783

Page 31: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

31

Gao, H., and J. Rice, 1986. Shear-stress intensity factors for a planar crack with slightly curved front, 784

J. Appl. Mech., 53(4), 774. 778. 785

Gao, H., J. Rice, and J. Lee, 1991. Penetration of a quasi-statically slipping crack into a seismogenic 786

zone of heterogeneous fracture-resistance, J. Geophys. Res., 96, 21,535. 21,548. 787

Gaudemer, Y., Tapponnier, P., Meyer, B., Peltzer, G., Guo, S., Chen, Z., Dai, H., Cifuentes, I., 1995. 788

Partitioning of crustal slip between linked active faults in the eastern Qilian Shan, and evidence for a 789

major seismic gap, the “Tianzhu gap”, on the western Haiyuan fault, Gansu (China). Geophys. J. Int. 790

120, 599–645. 791

Griffith, W. A., Nielsen, S., di Toro, G., Smith, S. A. F., 2010. Rough faults, distributed weakening, 792

and off-fault deformation. Journal of Geophysical Research, 115, B08409. 793

Hansen, A., Schmittbuhl, J., 2003. Origin of the Universal Roughness Exponent of Brittle Fracture 794

Surfaces: Stress-Weighted Percolation in the Damage Zone, Vol.90, No 4, doi: 795

10.1103/PhysRevLett.90.045504. 796

Hubert-Ferrari, A., Armijo, R., King, G., Meyer, B., and Barka, K., 2002. Morphology, displacement, 797

and slip rates along the North Anatolian Fault, Turkey. J. Geophys. Res., 107(B10), 2235, 798

doi:10.1029/2001JB000393.Johnson, A. M., Fletcher, R. C., 1994. Folding of viscous layers, 799

Columbia University Press, New York, 461 p. 800

Karig, D.E., 1983. Accreted terranes in the Northern part of the Philippine archipelago, Tectonics, 2, 801

211-236. 802

Klinger, Y., X. W. Xu, P. Tapponnier, J. Van der Woerd, C. Lasserre, and G. King, 2005. High-803

resolution satellite imagery mapping of the surface rupture and slip distribution of the Mw 7.8, 14 804

November 2001 Kokoxili Earthquake, Kunlun Fault, northern Tibet, China, Bull. Seismol. Soc. Am., 805

95(5), 1970–1987, doi:10.1785/0120040233. 806

Klinger, Y., R. Michel, and G. C. P. King, 2006. Evidence for an earthquake barrier model from Mw 807

similar to 7.8 Kokoxili (Tibet) earthquake slip-distribution, Earth Planet. Sci. Lett., 242(3–4), 354–808

364, doi:10.1016/j.epsl.2005.12.003. 809

Page 32: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

32

Klinger, Y., 2010. Relation between continental strike-slip earthquake segmentation and thickness of 810

the crust. J. Geophys. Res.-Solid Earth, 115, B07306, doi:10.1029/2009JB006550. 811

Kondo, H., Awata, Y., Emre, Ö., Doğan, A., Özalp, S., Tokay, F., Yıldırım, C., Yoshioka, T., 812

Okumura, K., 2005. Slip distribution, fault geometry, and fault segmentation of the 1944 Bolu-Gerede 813

earthquake rupture, North Anatolian Fault, Turkey. Bull. Seismol. Soc. Am., 95, 1234–1249, 814

doi:10.1785/0120040194. 815

Kondo, H., Özaksoy, V., Yıldırım, C., 2010. Slip history of the 1944 Bolu-Gerede earthquake rupture 816

along the North Anatolian fault system: Implications for recurrence behavior of multisegment 817

earthquakes. J. Geophys. Res.-Solid Earth, 115, B04316, doi:10.1029/2009JB006413. 818

Kurushin, R. A., Bayasgalan, A., Ölziybat, M., Enhtuvshin, B., Molnar, P., Bayarsayhan, C., Hudnut, 819

K., Jian, L., 1997. The surface rupture of the 1957 Gobi-Altay, Mongolia, Earthquake. Spec. Pap. 820

Geol. Soc.Am., 320, 142 p. 821

Jachens, R. C., Langenheim, V. E., Matti, J. C., 2002. Relationship of the 1999 Hector Mine and 1992 822

Landers fault ruptures to offsets on neogene faults and distribution of late cenozoic basins in the 823

eastern California shear zone, Bulletin of the Seismological Society of America, 92, 1592-1605. 824

Lee, J.J., Bruhn, R.L., 1996. Structural anisotropy of normal fault surfaces. J. Struct. Geol., 18, 1043-825

1059. 826

Mai, P.M., Beroza, G.C., 2002. A spatial random field model to characterize complexity in 827

earthquake slip. J. Geophys. Res.-Solid Earth, 107, doi:10.1029/2001JB000588. 828

Mandelbrot, B. B., 1983. The Fractal Geometry of Nature, 468 pp., Freeman, New York. 829

Manighetti, I., Campillo, M., Bouley, S., Cotton, F., 2007. Earthquake scaling, fault segmentation, 830

and structural maturity. Earth Planet. Sci. Lett., 253(3–4), 429–438, doi:10.1016/j.epsl.2006.11.004. 831

Meakin, P., 1998. Fractals: Scaling and Growth far From Equilibrium. Cambridge Univ. Press, New 832

York. 833

Méheust, Y., 2002. Ecoulements dans les fractures ouvertes, Ph. D. Thesis, Univ. Paris VI and Ecole 834

Normale Superieure, Paris. 835

Page 33: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

33

Mitchell, T.M., Faulkner, D.R., 2009. The nature and origin of off-fault damage surrounding strike-836

slip fault zones with a wide range of displacements: A field study from the Atacama fault system, 837

northern Chile. Journal of Structural Geology, 31 (8), 802-816. 838

Mitchell, A.H.G, Hernandez, F., and dela Cruz A. P., 1986. Cenozoic evolution of the Philippine 839

archipelago, Journal of Southeast Asian Earth Sciences, 1, 3-22. 840

Okaya, D. A., Thompson, G. A., 1985. Geometry of Cenozoic extensional faulting: Dixie Valley, 841

Nevada. Tectonics, 4, 107-125. 842

Palumbo, L., Benedetti, L., Bourles, D., Cinque, A., FinkelI, R., 2004. Slip history of the Magnola 843

fault (Apennines, Central Italy) from 36Cl surface exposure dating: Evidence for strong earthquakes 844

over the Holocene, Earth Planet. Sci. Lett., 225, 163–176. 845

Poliakov, A. N., Dmowska, R., Rice, J. R., 2002. Dynamic shear rupture interactions with fault bends 846

and off‐axis secondary faulting. J. Geophys. Res.-Solid Earth, 107(B11), 2295, 847

doi:10.1029/2001JB000572. 848

Power, W.L., Tullis, T.E., Brown, S.R., Boitnott, G.N., Scholz, C.H., 1987. Roughness of natural 849

fault surfaces. Geophys. Res. Lett., 14, 29–32. 850

Power, W.L., Tullis, T.E., and Weeks, J. D., 1988. Roughness and wear during brittle faulting, J. 851

Geophys. Res., 93, 15,268–15,278. 852

Power,W. L., Tullis, T.E., 1989. The relationship between slickenside surfaces in fine-grained quartz 853

and the seismic cycle. J. Struct. Geol., 11, 879-893. 854

Power, W.L., Tullis, T.E., 1991. Euclidean and fractal models for the description of rock surface 855

roughness. J. Geophys. Res.-Solid Earth, 96, 415-424, doi:10.1029/90JB02107. 856

Power, W.L., Tullis, T.E., 1992. The contact between opposing fault surfaces at Dixie Valley, 857

Nevada, and implications for fault mechanics. J. Geophys. Res.-Solid Earth, 97, 15425-15435, 858

doi:10.1029/92JB01059. 859

Power, W. L., Durham, W. B., 1997. Topography of natural and artificial fractures in granitic rocks: 860

Implications for studies of rock friction and fluid migration. Int. J. Rock Mech. Min. Sci., 34, 979–861

989. 862

Page 34: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

34

Press, W.H., Teukolsky, S.A., Vetterling, W.T., Flannery, B.P., 2007. Numerical Recipes, 3rd

ed., 863

Cambridge Univ. Press: Cambridge, UK. 864

Renard, F., Schmittbuhl, J., Gratier, J.-P., Meakin, P., E. Merino, 2004. The three-dimensional 865

roughness of stylolites in limestones: roughness analysis and possible genetic implications. Journal of 866

Geophysical Research, 108, B03209, doi:10.1029/2003JB002555. 867

Renard, F., Voisin, C., Marsan, D., Schmittbuhl, J., 2006. High resolution 3D laser scanner 868

measurements of a strike-slip fault quantify its morphological anisotropy at all scales. Geophys. Res. 869

Lett., 33, L04305. 870

Resor, P. G., Meer, V. E., 2009. Slip heterogeneity on a corrugated fault. Earth Planet. Sci. Lett., 288, 871

483-491. 872

Rigenbach, J.C., Pinet, N.., Stephan, J.F., and Delteil, J., 1993. Structural variety and tectonic 873

evolution of strike-slip basins related to the Philippine fault system, Northern Luzon, Philippines, 874

Tectonics, 12, 187-203. 875

Sagy A., Brodsky, E.E., Axen, G.J., 2007. Evolution of fault-surface roughness with slip. Geology, 876

35, 283–286. 877

Savage, H.M., Brodsky, E.E., 2011. Collateral Damage: Capturing Fault Strand Formation in Fracture 878

Profiles, Journal of Geophysical Research, 116, B03405, doi 10.1029/2010JB007665. 879

Schmittbuhl, J., Gentier, S., Roux, R., 1993. Field measurements of the roughness of fault surfaces. 880

Geophys. Res. Lett., 20, 639–641. 881

Schmittbuhl, J., Schmitt, F., Scholz, C.H., 1995a. Scaling invariance of crack surfaces, J. Geophys. 882

Res., 100, 5953-5973. 883

Schmittbuhl, J., Vilotte, J., Roux, S., 1995b. Reliability of self-affine measurements. Phys. Rev. E. 51, 884

131-147. 885

Schmittbuhl, J., Chambon, G., Hansen, A., Bouchon, M., 2006. Are stress distributions along faults 886

the signature of asperity squeeze? Geophys. Res. Lett., 33, doi:10.1029/2006GL025952. 887

Schmittbuhl, J, Steyer, A., Jouniaux, L., Toussaint, R., 2008. Fracture morphology and viscous 888

transport, Int. J. Rock. Mech. Min. Sci., 45, 422-430. 889

Page 35: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

35

Schmittbuhl, J., Delaplace, A., Maloy, K., Perfettini, H., Vilotte, J., 2003. Slow Crack Propagation 890

and Slip Correlations. Pure. Appl. Geophy. 160, 961-976, doi:10.1007/PL00012575. 891

Scholz, C. H., Aviles, C. A., 1986. The fractal geometry of faults and faulting, in 5th Ewing 892

Symposium, Earthquake Source Mechanisms, Geophys. Monogr. Ser., 37, edited by Das, S., 893

Boatwright, J., Scholz, C. H. pp. 147-156, AGU, Washington D. C. 894

Sharp, R. V., 1967. San Jacinto Fault zone in the peninsular ranges of Southern California, Geological 895

Society of America Bulletin, 78, 705-730. 896

Simonsen, I., Vandembroucq, D., Roux, S., 2000. Wave scattering from self-affine surfaces. Phys. 897

Rev. E., 61, 5914–7. 898

Thouvenot, F., 1998. The ML 5.3 Epagny (French Alps) earthquake of 1996 July 15: A long-awaited 899

event on the Vuache Fault, Geophys. J. Int., 135, 876–892. 900

Vallée, M., M. Landès, N. M. Shapiro, and Y. Klinger, 2008, The 14 November 2001 Kokoxili 901

(Tibet) earthquake: High‐frequency seismic radiation originating from the transition between 902

sub‐Rayleigh and supershear rupture velocity regimes, J. Geophys. Res., 113, B07305, 903

doi:10.1029/2007JB005520. 904

Wallace, R. E., Whitney R. A., 1984. Late Quaternary history of the Stillwater Seismic Gap, Nevada. 905

Bull. Seismol. Soc. Am., 74, 301-314. 906

Wechsler, N., Ben-Zion, Y., Christofferson, S., 2010. Evolving Geometrical Heterogeneities of Fault 907

Trace Data. Geophys. J. Int., 182, 551-567. 908

Wesnousky, S. G., 1988. Seismological and structural evolution of strike-slip faults. Nature, 335(22), 909

340–343, doi:10.1038/335340a0. 910

Wesnousky, S., 2006. Predicting the endpoints of earthquake ruptures. Nature, 444(7117), 358–360, 911

doi:10.1038/nature05275. 912

Wesnousky, S., 2008. Displacement and Geometrical Characteristics of Earthquake Surface Ruptures: 913

Issues and Implications for Seismic-Hazard Analysis and the Process of Earthquake Rupture. Bull. 914

Seismol. Soc. Am., 98, 1609-1632. 915

Page 36: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

36

Wei, Z., He, H., Shi, F., Gao, X., Xu, C., 2010. Topographic characteristics of rupture surface 916

associated with the Wenchuan earthquake of Mw7.9, in May 12, 2008. Bull. Seismol. Soc. Am., 100, 917

2669-2680. 918

Zhang, W., Jiao, D., Zhang, P., Molnar, P., Burchfiel, B. C., Qidong, D., Yipeng, W., Fangøin, S., 919

1987. Displacement along the Haiyuan fault associated with the great 1920 Haiyuan, China, 920

earthquake, Bull. Seismol. Soc. Am., 77, 117-131. 921

922

Page 37: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

37

Figure captions 923

924

Figure 1: Two-dimensional roughness profiles from the Corona Heights fault surface parallel and 925

perpendicular to slip. A: Profiles parallel to the slip direction and, B: perpendicular to the slip 926

direction. For both directions, a magnified portion of the profiles has a statistically similar appearance 927

to the entire profiles when using the scaling transformation zzxxH

,//, . For each scale, 928

profiles have been shifted vertically for clarity. 929

930

Figure 2: Corona Heights fault. Multiple bumpy discrete slip surfaces constituting lenses and 931

striations can be detected at all scales, from the measurement resolution of each scanner device to the 932

size of the entire exposure. A: Whole outcrop view. The inset corresponds to the surface shown on 933

Figure 2C. B: Zoom on the fault showing different segments constituting the surface. C&D: Map of 934

fault surfaces scanned using LiDAR topography. The inset in C corresponds to the patch shown on 935

Figure 2D. E and F: Maps of fault surfaces scanned with the laser profilometer. G and H: Zoom on 936

the Figure 2E-F scanned with the white light interferometer. 937

938

Figure 3: Bolu Fault. A: Photograph of the fault zone in the Bolu limestone. Dashed red contour 939

corresponds to the limits of the cloud of points shown in B. C: Photograph of a well-preserved slip 940

surface constituting the fault zone. D: LiDAR data of surface in C. E and F: Zoom on the fault surface 941

and the corresponding topographic map acquired with the laser profilometer, which still includes 942

anisotropic roughness features in the slip direction. 943

944

Figure 4: Digitized surface rupture trace of the three largest segments of the Landers earthquake. The 945

data correspond to Figure 5G in Klinger (2010). The inset zooms on one of the steps removed for the 946

Fourier transform analysis. Each segment is individualized by different colors and the steps are 947

represented in pink. The reconstructed trace by removing steps is displayed in gray. 948

Page 38: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

38

949

Figure 5: Example of the 3-D average power spectrum of one LiDAR fault patch (Corona-B) 950

obtained by averaging in a geometric progression several thousands 2-D power spectra of individual 951

profiles (we have changed the color of the spectra every 20 profiles successively extracted from the 952

surface). 953

954

Figure 6: Typical average power spectra in the slip direction and perpendicular to it for fault patches 955

scanned with the three devices used in our study. Errors bars with 68% confidence interval (one 956

sigma), and power law fits performed in linear portions of each average spectra, are shown. At the 957

LiDAR scale, the black arrow indicates the lower limit used for the fits. This lower limit underlines 958

the length scale at which spectra flatten out when they intersect the noise spectrum calculated by 959

scanning smooth, planar reference surfaces. For the laser profilometer and the WLI, the spectral 960

power levels of our natural fault roughness data fall at a vertical higher position than the noise spectra. 961

However, at the WLI scale, the vertical dashed blue line at the bending of spectra indicates the lower 962

limit of our fits. Even if this systematic change of slope at the length scale mm05.0 seems to be 963

related to intrinsic physical fault roughness properties, we arbitrarily limit our scaling analysis to 964

scales larger than mm05.0 . 965

966

Figure 7: Plot of the Hurst exponents (see Table 1) in slip direction (B) and normal to it (A) of the 41 967

scanned fault patches. The average Hurst exponent is equal to 58.0 along the direction of slip and 968

81.0 perpendicular to it. The shaded area and dashed lines indicate 1 and 2 confidence 969

intervals, respectively. 970

971

Figure 8: Illustration of the pre-factor variability along the Corona Heights fault. Log-log graph 972

gathering the global laser profilometer spectrum with the 6 averaged Fourier power spectra obtained 973

at the LiDAR scale. Note the intercepts range of all 2-D LiDAR spectra, performed from each 974

individual profiles extracted from the six fault patches, highlighted by the two dashed dark power 975

Page 39: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

39

laws. The global laser profilometer spectrum falls in this range of intercepts sampled by the whole 976

population of the individual profiles that we explore at the LiDAR scale. 977

978

Figure 9: Global Fourier power spectra from the five faults analyzed along the slip direction (left) 979

and perpendicular to it (right). Each curve at each scale (LiDAR, laser profilometer, white light 980

interferometer) includes together the average spectra of several sub-surfaces (or fault patches in Table 981

1). Power-law fits (thick gray lines) with a roughness exponent of 6.0// H and 8.0H , 982

connecting the field and laboratory data in both directions, are shown on plot for eye guidance. The 983

black arrow and the vertical dashed blue line indicate, respectively: the level of noise of the LiDAR 984

and the lower limit for the fit performed at the WLI scale. Dotted black lines indicate the range of pre-985

factors,

//,21

//,//, )(,H

kCkPC , extracted from the power-law fits at mk 1 (see all the 986

values in Table 3), and used for the Figure 10. One way to more easily interpret this log-log graph and 987

to compare our results with previous studies like Sagy et al. (2007), is to convert the power spectrum 988

module in term of root-mean-square ( RMS ) roughness amplitude using Parseval’s Theorem 989

(Brodsky et al., 2011). Indeed if 10 //, H , for a profile of length L , the integration of 990

//,21

//,)(H

kCkP over the wavelength (with k/1 ) yields that the RMS roughness 991

correlates as

//,

5.0

//,

//,

2H

LH

CRMS . In this sense, red lines represent power-law fits for 992

three self-similar rough surfaces (i.e. H = 1) with various pre-factors, that are: LRMS 1.0 , 993

LRMS 01.0 , LRMS 001.0 . 994

995

Figure 10: Log-log plots of the range of global pre-factor magnitude of the five faults extracted from 996

Figure 9 (see also Table 3) versus the estimated slip (Table 1). 997

998

Figure 11: Compilation of the surface rupture roughness results: Digitized rupture traces (top) and 999

corresponding Fourier power spectrum (bottom). Because of the abrupt steps biasing the Fourier 1000

Page 40: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

40

transform computation, we have performed this roughness analysis on each individual segment 1001

between two steps of the whole profile. The same color code is respected between the individual 1002

segments and the corresponding spectra. In addition, the gray rupture profile and the corresponding 1003

gray Fourier power spectrum represent a reconstruction of the entire profile. Power-law fits and the 1004

inferred Hurst exponents on the linear part of each curves at large scale (above the cross-over length 1005

scale indicated by the gray vertical bar and marking the beginning of the regime controlled by data re-1006

sampling) are represented on each graph. Note here that for each rupture trace, the curves 1007

corresponding respectively to the selected segments and to the reconstructed profiles are characterized 1008

by a similar scaling exponent but variable pre-factors. 1009

1010

Figure 12: Variability of the pre-factor and slip accumulated for the eight studied rupture traces. A: 1011

Stack of all the Fourier power spectra of the whole surface ruptures stacked, underlining the global 1012

trend of the self-affine behavior at large scale. Power-law fit giving an average Hurst exponent of 1013

1.08.0 is indicated. B: Log-log plot of the average pre-factors (corresponding to those of the 1014

spectra of the whole rupture traces and indicated in Table 2) with their respective typical fluctuations 1015

(i.e. the range of pre-factors extracted from the spectra of the individual segments constituting the 1016

whole rupture traces) of each rupture traces in function of the finite geological offsets (see Table 2). 1017

1018

Figure 13: Angular dependence of the Hurst exponent (left) and the pre-factor (right) computed on a 1019

synthetic anisotropic self-affine surface with two input exponents in perpendicular directions 1020

( 6.0// H and 8.0H ), and for a patch (Corona-A, see Table 1) scanned with the LiDAR on the 1021

Corona Heights fault surface. The Hurst exponents and the pre-factors were calculated on series of 2-1022

D profiles extracted at an angle between the slip direction ( 0 ) and the perpendicular direction 1023

( 90 ). 1024

1025

Figure 14: Comparison of the roughness of the earthquakes surface ruptures with the Corona Heights 1026

fault topography. The Fourier power spectra normal to the slip direction of the Corona Heights fault 1027

Page 41: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

41

surface are plotted on a log-log graph together with those obtained for the eight continental strike-slip 1028

earthquakes surface rupture traces. The Corona Heights fault data are identical to those plotted on 1029

Figure 9, and those of the surface ruptures correspond to that of Figure 12. Both data sets are 1030

connected with a unique power-law fit (grey line) with a Hurst exponent of 0.8. 1031

1032

1033

1034

1035

1036

1037

1038

1039

1040

1041

1042

1043

1044

1045

Page 42: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

42

Table 1. Laser scanner characteristics and fault roughness results.

Fault Name Lithology

& Slip* Sense

Fault

Patches Scanner

§

x

Spatial

precision

#

z

)05.0(

//

H

)05.0(

H

Average

//H H

Vuache-

Sillingy

45°57’14.5’’N

6°2’56’’E

Limestone

10-30m

Strike

slip

Surf-1 GS 100

(Trimble)

20

mm 5mm

4.5

mm 0.60 0.78

0.60

0.07

0.81

0.02

Surf-7 GS 100

(Trimble)

20

mm 5mm

4.5

mm 0.68 0.82

Surf-6

LMS

Z420i

(Riegl)

30

mm

7.5mm 10.2

mm 0.50 0.82

Surf-

JPG

S10

(Trimble) 1

mm

0.25mm 0.9

mm 0.63 0.83

Small Lab.

profilo-

meter

20

µm 1µm

<

1µm

0.65 0.81 0.61

0.05

0.80

0.01 Vu-1-G 0.57 0.80

Vu-A-1

WLI

2µm

0.025 µm

3nm

0.58 0.76 0.60

0.04

0.79

0.03

Vu-A-2 2µm 0.55 0.78

Vu-A-7 1µm 0.65 0.84

Vu-A-8 1µm 0.61 0.79

Corona

Heights

37°45’55’’N

122°26’14’’E

Chert

several m

to >1km

Strike

slip

Corona-

A

HDS 3000

Leica

5

mm

1.25mm 2

mm

0.57 0.85

0.65

0.04

0.83

0.03

Corona-

B 0.65 0.81

Corona-

C 0.67 0.87

Corona-

D 0.64 0.81

Corona-

E 0.69 0.79

Corona-

F 0.67 0.85

P3 Lab.

profilo-

meter

20

µm 1µm

<

1µm

0.66 0.86 0.63

0.04

0.85

0.03 Co-

AGU 0.60 0.85

Co-A-4

WLI

2µm 0.025

µm 3nm

0.62 0.82 0.62

0.03

0.83

0.02 Co-A-9 1µm 0.63 0.85

Bolu

40°41’07’’N

31°34’04’’E

Limestone

20m to

85km

Strike

slip

Stack

2345

Ilris-3D

Optech

20

mm 5mm

20

mm

0.48 0.79

0.50

0.07

0.78

0.02

Stack67 0.50 0.78

Stack12 0.44 0.78

W-

detail-2 0.45 0.76

E-detail-

3 0.49 0.80

E-detail-

2 0.46 0.80

E-detail-

1 0.65 0.74

Bolu-1 Lab.

profilo-

meter

20

µm 1µm

<

1µm

0.58 0.79 0.55

0.05

0.76

0.04 Bolu-2 0.51 0.74

1046

Page 43: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

43

1047

Fault Name Lithology

& Slip* Sense

Fault

Patches Scanner

§

x

Spatial

precision #z

)05.0(

//

H

)05.0(

H

Average

//H H

Dixie Valley

39°56’48’’N

117°56’43’’E

Rhyolites

several m

to 3-6km

Normal

Dixie-1

HDS 3000

Leica

5

mm

1.25mm 2

mm

0.66 0.79 0.59

0.08

0.82

0.03

Dixie-2 0.63 0.80

Dixie-3 0.61 0.84

Dixie-4 0.47 0.84

Map-1 Lab.

profilo-

meter

20

µm 1µm

<

1µm

0.57 0.82 0.59

0.03

0.81

0.03 Map-2 0.61 0.81

Dixie-D

WLI

2µm

0.025 µm

3nm

0.56 0.88 0.56

0.05

0.88

0.01

Dixie-H 1µm 0.50 0.87

Dixie-E 2µm 0.61 0.89

Dixie-C 2µm 0.59 0.89

Magnola

42°7’N

13°28’31’’E

Limestone

several m

to >500m

Normal

A32 Ilris-3D

Optech

20

mm 5mm 20

mm 0.59 0.77

M2

Lab.

profilo-

meter

20

µm 1µm

<

1µm 0.60 0.83

1048 * Except for the Vuache-Sillingy fault surface, a lower and upper bound of the displacement is given. 1049

Although total geological cumulated slip for the fault zone as a whole can be kilometric (i.e. the upper 1050

bound), scanned individual surfaces within the fault zone may have experienced considerably less slip 1051

(i.e. the lower bound). 1052

§ Spatial length scale resolution ( x ). 1053

# Vertical resolution ( z ). 1054

1055

1056

1057

1058

1059

1060

1061

1062

Page 44: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

44

Table 2. Characteristics and roughness results of earthquake rupture maps used in this study.

(references for each earthquake are given in the text)

Name Year Magnitude Rupture

length (km)

Total geological

offset (km) RH Pre-factor

(m3)

Owens Valley

(USA) 1872 Mw 7.5-7.8 81 20-30 0.6 0.1 4 x 10-2

Haiyuan

(China) 1920 Ms 8 to 8.7 200 15 to 95 0.8 0.1 7 x 10-3

Gobi-Altay

(Mongolia) 1957 M 8.3 235 2 to 20 0.7 0.1 2 x 10-2

Superstition

Hills (USA) 1987 M 6.6 18 24 0.7 0.1 2 x 10-3

Luzon

(Philippine) 1990 Mw 7.8 107 50 to 200 0.8 0.1 1 x 10-2

Landers

(USA) 1992 Mw 7.2 65 3.1 to 4.6 0.8 0.1 1 x 10-3

Zirkuh

(Iran) 1997 Mw 7.2 104 60 0.9 0.1 3 x 10-4

Hector mine

(USA) 1999 Mw 7.1 39 3.4 0.7 0.1 7 x 10-3

1063 1064

1065

1066

1067

1068

1069

1070

1071

1072

1073

1074

1075

1076

1077

1078

1079

1080

Page 45: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

45

Table 3. Pre-factor (exhumed slip surfaces)

Parallel to slip Perpendicular to slip

Vuache-Sillingy Pre-factor (m3)

min 8 x 10-8 2 x 10-6

max 2 x 10-5 1 x 10-4

Corona Heights Pre-factor (m3)

min 4 x 10-7 1 x 10-5

max 5 x 10-5 2 x 10-4

Bolu Pre-factor (m3)

min 2 x 10-8 3 x 10-7

max 5 x 10-5 2 x 10-4

Dixie Valley Pre-factor (m3)

min 1 x 10-7

2 x 10-5

max 1 x 10-5

Magnola Pre-factor (m3)

min 1 x 10-7 5 x 10-6

max 5 x 10-4 8 x 10-4

1081 1082

1083

1084

1085

1086

1087

1088

1089

1090

1091

1092

1093

1094

1095

1096

1097

Page 46: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

46

Figure 1 1098

1099

1100

Page 47: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

47

Figure 2 1101

1102

1103

Page 48: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

48

Figure 3 1104

1105

Page 49: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

49

Figure 4 1106

1107

1108

1109

1110

1111

1112

1113

1114

1115

1116

1117

1118

1119

1120

1121

1122

1123

1124

1125

Page 50: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

50

Figure 5 1126

1127

1128

1129

1130

1131

1132

1133

1134

1135

1136

1137

1138

1139

1140

1141

1142

1143

Page 51: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

51

Figure 6 1144

1145

1146

1147

Page 52: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

52

Figure 7 1148

1149

1150

1151

1152

1153

1154

1155

1156

1157

1158

Page 53: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

53

Figure 8 1159

1160

1161

1162

1163

1164

1165

1166

1167

1168

Page 54: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

54

Figure 9 1169

1170

1171

1172

1173

1174

1175

1176

Page 55: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

55

1177

1178

1179

1180

1181

1182

1183

1184

1185

1186

1187

1188

Page 56: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

56

Figure 10 1189

1190

1191

1192

1193

1194

1195

1196

1197

1198

1199

1200

1201

1202

Page 57: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

57

Figure 11 1203

1204

1205

1206

1207

1208

1209

1210

1211

Page 58: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

58

1212

1213

1214

1215

1216

1217

1218

1219

1220

Page 59: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

59

Figure 12 1221

1222

1223

1224

1225

1226

1227

1228

1229

Page 60: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

60

Figure 13 1230

1231

1232

1233

1234

1235

1236

1237

1238

1239

1240

1241

1242

1243

1244

1245

1246

1247

1248

Page 61: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

61

Figure 14 1249

1250

1251

1252

1253

1254

1255

1256

1257

1258

1259

1260

1261

1262

1263

1264

Page 62: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

62

Appendix A: Potential bias in roughness data 1265

A.1. Reliability of the roughness results at the LiDAR scale: effect of the noise in the acquisition 1266

system 1267

Before analyzing the main biases inherent to LiDAR data acquisition, we emphasize that raw 1268

scanner data acquired with the home-made laser profilometer and the WLI are considered as quasi-1269

ideal since the level of the intrinsic white noise of these instruments is well below that recovered from 1270

the fault surfaces (see Figure 6). A detailed description of the conditions in which measurements with 1271

this scanner device departs from the reality is given by Méheust (2002). In addition, the slope of the 1272

spectra computed on WLI scans at spatial scales larger than mm05.0 are consistent with those of 1273

laser profilometry for the same range of length scales (see Figure 9). Given that LiDAR data include 1274

the largest bias, we focus our noise analysis on this instrument. Consequently, the results provided by 1275

the following analyses can be considered as end-members of the noise effect estimation for the three 1276

instruments used in our study. 1277

In the spirit of the work of Schmittbuhl et al. (1995) on the reliability of a self-affine 1278

measurement on 2-D rough profile, Candela et al. (2009) have reviewed different statistical 1279

methodologies which allow the assessment and characterization of the anisotropic self-affine behavior 1280

of fault topography. This work was mainly devoted to precisely define the intrinsic error of the 1281

statistical methods (such as the Fourier power spectrum) to estimate the scaling properties of fault 1282

surface roughness. 1283

Here, a new test is performed by taking into account the error encountered in the spatial position 1284

( ZYX ,, ) of each points measured by the 3-D laser scanner. We use a synthetic anisotropic self-1285

affine surface (Candela et al., 2009) of m55 with an original regular point spacing of mm5 1286

(Figure A1), and with two different Hurst exponents in perpendicular direction ( 6.0// H and 1287

8.0H ). In order to simulate the error inherent to LiDAR data acquisition on the spatial position 1288

of each points ( ZYX ,, ), we add Gaussian white noise with a distribution of ]5.2,0[ mm on the 1289

original position of each point, to obtain the perturbed grid ( ZYX ,, 11). Then the height

1Z of each 1290

Page 63: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

63

point at these new positions (11,YX ) is computed by interpolating (bilinear interpolation) the four 1291

nearest points of each of these new positions on the original ideal model (Figure A1). In a final step, 1292

Gaussian white noise, with a standard deviation equal to mm5 , is randomly added on the 1293

interpolated heights 1Z , to yield the error of the LiDAR data in the vertical position. 1294

After generating this noisy cloud of points (111 ,, ZYX ), we extract profiles oriented along slip 1295

direction and perpendicular to it as is done for the measurements, and estimate both Hurst exponents. 1296

The results are then averaged over 100 realizations of synthetic surfaces. Due to the noise, the Fourier 1297

power spectra flatten at short length scales (as for examples in the slip direction shown in Figure A1), 1298

which results in a slight underestimation of the Hurst exponents. For both directions, along slip and 1299

perpendicular to it, we find that the median estimates of //H and H of the biased synthetic surfaces 1300

are 06.059.0 and 11.074.0 respectively, compared to the noise-free data where the Hurst 1301

exponents were 07.06.0 and 05.078.0 , respectively. Note that the error bar of the estimated 1302

Hurst exponent of the biased synthetic surfaces is approximately twice larger in the direction 1303

perpendicular to slip relative to that in the slip direction. In both directions, even if the Hurst exponent 1304

is slightly underestimated for the noisy data, its value is still included in the range given by the 1305

standard deviation of the noise-free data. Therefore, the noise in the LiDAR data could be estimated 1306

as well as the reliability of Hurst exponent values. 1307

1308

A.2. Reliability of the roughness results at surface rupture scale: effect of re-sampling 1309

For each earthquake, once the surface rupture map has been digitized, the data set is re-sampled 1310

to a regular spacing to ensure consistent spatial sampling, independent of the length of each rupture. 1311

This re-sampling is performed to avoid bias due to local wiggles of the rupture trace (Klinger, 2010). 1312

We verify here how this re-sampling process affects the spectra of the Fourier transform. The original 1313

digitized rupture trace of the Hector mine earthquake, taken as an example, is re-sampled with various 1314

constant values of x in the range 60-620 m (Figure A2). In the Fourier power spectra of this set of 1315

digitized rupture traces, two regimes can be observed. At small scales, i. e. large wave numbers, 1316

Page 64: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

64

(between approximately m120 and m1200 ) the behavior can be attributed to data re-sampling. At 1317

large scales, i. e. small wave numbers (above m1200 ), a power law giving a Hurst exponent 1318

74.0RH represents the best fit. The cross-over length scale between the two regimes corresponds 1319

to the maximum spacing between two points in the original data. Whatever the value of x taken for 1320

re-sampling the data, the cross-over length scale remains at the same position. We propose that the 1321

regime at large scales, characterizing roughness properties of the digitized ruptures traces is therefore 1322

not affected by data re-sampling; both the slope and the pre-factor of each spectrum are identical 1323

(Figure A2). 1324

The same re-sampling procedure has been performed on ideal synthetic self-affine profiles in 1325

order to precisely define if the scaling property could be modified (Figure A3). An original self-affine 1326

profile (6.001.0 LRMS ) with regular spacing of m500 and a total length of km100 (extracted 1327

from a synthetic surface as previously presented), is modified by adding to the X coordinates a 1328

Gaussian white noise perturbation with a distribution of ]250,0[ m (Figure A3). This altered profile is 1329

re-sampled with different values of x in the range [40-350] m. The Fourier spectra (Figure A3) 1330

indicate that for the different values of x , the large scale regime (above m750 ) characterizing the 1331

input self-affine behavior with a Hurst exponent of 6.0 is not biased. This test validates that the re-1332

sampling procedure makes it possible to keep the scaling information of the rupture traces at spatial 1333

scales above 1200 m. 1334

1335

Page 65: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

65

Figure A1: Effect of noise inherent to LiDAR data acquisition. A: Ideal synthetic m55 self-affine 1336

surface with an original regular point spacing of mm5 . B: Illustration of the addition of a noise in the 1337

regularly spaced original grid ),,( ZYX . C: Comparison of the Fourier power spectra in direction of 1338

slip obtained for the ideal synthetic surface and the noisy synthetic surface. Both vertical dashed grey 1339

lines indicate limits taken for fitting the Hurst exponent. D: Distribution of measured Hurst exponents, 1340

on 100 simulations, for the direction of slip and perpendicular to it. Black bars show the ideal 1341

simulated fault surface models, and the gray bars correspond to the noisy simulated fault surface 1342

models. The solid lines (black for the ideal case and gray for the natural case) represent the fits of the 1343

measurements to a normal distribution with mean and standard deviation given at the top left 1344

for the noisy natural case and the top right for the noise-free case. 1345

1346

Figure A2: Effect of re-sampling effect on earthquake surface rupture roughness: example of Hector 1347

Mine earthquake. A: Digitized surface rupture traces of the Hector Mine earthquake. The original 1348

rupture map with an irregular point spacing (pink profile) is re-sampled in order to ensure consistent 1349

spatial sampling with a regular spacing )( x . The inset indicates the position of the zoom located on 1350

the right, showing the irregular point spacing on the top and the gradual increase of the re-sampling 1351

(or decreasing of x ) with a regular point spacing of profiles downwards. B: Fourier power spectra of 1352

the digitized rupture traces shown in A. Spectra colors are the same than in A. On the right: the 1353

spectra have been shifted vertically to improve the visibility. The blue vertical bar on both graphs 1354

highlights the cross-over length scale, at approximately m1200 , between both regimes: the lower 1355

regime is attributed to data re-sampling, the upper regime characterizes roughness properties of the 1356

digitized profiles with a Hurst exponent 74.0RH . 1357

1358

Figure A3: Effect of re-sampling on synthetic self-affine profiles. A: Example of synthetic rough 1359

profiles with a standard deviation that scales as 6.001.0 LRMS , analogue of digitized surface 1360

rupture traces. The green profile at the top of the left figure is the original ideal synthetic profile with 1361

Page 66: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

66

a regular point spacing x of m500 . A Gaussian white noise perturbation with a distribution 1362

]250,0[ m is added on the original regular spacing to obtain a noisy profile (pink curve) similar to 1363

that of original ruptures maps. This noisy synthetic profile with irregular point spacing is re-sampled 1364

with a regular spacing x . The inset zooms on the synthetic profiles located on the right. B: Fourier 1365

power spectra of the synthetic self-affine profiles shown in A. Colors of each curves correspond to 1366

those of profiles. On the right: the spectra have been shifted vertically to improve the visibility. 1367

1368

1369

1370

1371

1372

1373

1374

1375

1376

1377

1378

1379

1380

1381

1382

1383

1384

1385

1386

1387

1388

Page 67: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

67

Figure A1 1389

1390

1391

Page 68: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

68

Figure A2 1392

1393

1394

Page 69: Roughness of fault surfaces over nine decades of …seisweb/emily_brodsky/reprints/...134 (Candela et al., 2009) faults have been updated and extended with three new faults (Corona

69

Figure A3 1395

1396

1397

1398