Quick viewing(Text Mode)

The Shear Deformation Zone and the Smoothing of Faults with Displacement Clement Perrin, Felix Waldhauser, Christopher Scholz

The Shear Deformation Zone and the Smoothing of Faults with Displacement Clement Perrin, Felix Waldhauser, Christopher Scholz

The Shear Deformation Zone and the Smoothing of Faults With Displacement Clement Perrin, Felix Waldhauser, Christopher Scholz

To cite this version:

Clement Perrin, Felix Waldhauser, Christopher Scholz. The Shear Deformation Zone and the Smooth- ing of Faults With Displacement. Journal of Geophysical Research : Solid Earth, American Geophys- ical Union, 2021, 126 (5), pp.e2020JB020447. ￿10.1029/2020JB020447￿. ￿hal-03284736￿

HAL Id: hal-03284736 https://hal.archives-ouvertes.fr/hal-03284736 Submitted on 12 Jul 2021

HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non, lished or not. The documents may come from émanant des établissements d’enseignement et de teaching and research institutions in France or recherche français ou étrangers, des laboratoires abroad, or from public or private research centers. publics ou privés. 1 The Shear Deformation Zone and the Smoothing of Faults with Displacement

2

3 Clément Perrin1,2,*, Felix Waldhauser1 and Christopher H. Scholz1

4

5 1 Lamont Doherty Earth Observatory at Columbia University, New York, USA

6 2 Present address : Laboratoire de Planétologie et Géodynamique, UMR6112,

7 Observatoire des Sciences de l’Univers de Nantes Atlantique, UMS3281, Université de

8 Nantes, Université d’Angers, CNRS, 2 rue de la Houssinière - BP 92208, 44322 Nantes

9 Cedex 3, France

10 * Corresponding author: [email protected]

11

12 Keypoints:

13 Across strike distributions of aftershocks of large describe the width of the

14 shear deformation zone around large faults.

15

16 The zone of active shear deformation scales with roughness and narrows as a

17 power law with fault displacement.

18

19 stress drops decrease with fault displacement and hence fault roughness or

20 displacement rate.

21

22 Keywords:

23 Fault maturity; shear deformation zone; large earthquakes; aftershock distributions;

24 scaling laws.

25

1 26 Plain Language Summary

27 Active fault zones worldwide are 3D features made of a parent fault and secondary

28 faults and fractures that damaged the surrounding medium. During and soon after a

29 large earthquake, these structures are reactivated, highlighted by numerous smaller

30 events –aftershocks. Their distribution allows us to characterize the zone of shear

31 deformation around the fault plane. In this study, we show that the width of the shear

32 deformation zone is narrower around mature faults than around immature faults. It

33 decreases as a power law with cumulative fault displacement which we infer to be the

34 result of the smoothing of the fault with wear through geological times. We also find that

35 the stress-drop of mainshocks decrease with fault smoothness or slip rate, both of which

36 correlate with maturity. Our study provides some relations to better understand and

37 anticipate the size of off-fault deformation reactivated during and after an earthquake,

38 based on geological fault parameters.

39

40 Abstract

41 We use high-resolution earthquake locations to characterize the three-dimensional

42 structure of active faults in and how it evolves with fault structural maturity.

43 We investigate the distribution of aftershocks of several recent large earthquakes that

44 occurred on continental strike slip faults of various structural maturity (i.e., various

45 cumulative fault displacement, length, initiation age and slip rate). Aftershocks define a

46 tabular zone of shear deformation surrounding the mainshock rupture plane.

47 Comparing this to geological observations, we conclude that this results from the re-

48 activation of secondary faults. We observe a rapid fall off of the number of aftershocks at

49 a distance range of 0.06–0.22 km from the main fault surface of mature faults, and 0.6-1

50 km from the fault surface of immature faults. The total width of the active shear

2 51 deformation zone surrounding the main fault plane reaches 1-2.5 km and 6-9 km for

52 mature and immature faults, respectively. We find that the width of the shear

53 deformation zone decreases as a power law with cumulative fault displacement.

54 Comparing with a dynamic rough fault model, we infer that the narrowing of the shear

55 deformation zone agrees quantitatively with earlier estimates of the smoothing of faults

56 with displacement, both of which are aspects of fault wear. We find that earthquake

57 stress drop decreases with fault displacement and hence with increased smoothness

58 and/or slip rate. This may result from fault healing or the effect of roughness on friction.

59

60 1. Introduction

61 A fault zone is a complex brittle-frictional system that wears as slip occurs on it. It is

62 formed of three main features, that evolve with fault growth (Fig. 1): (i) the cataclastic

63 core contains the cataclastic detritus of wear of the slipping surfaces of the fault. Its

64 width (WC in Fig. 1) increases linearly with fault displacement at a rate that depends on

65 the strength of the wall rock (Scholz, 1987, 2019, pp 132). For displacements greater

66 than a few hundred meters, growth of the fault core levels off at a thickness of a few tens

67 of meters (Scholz, 2019, pp 132); (ii) Beyond the fault core lies a region of pervasive

68 tensile fracturing which defines the “dilatant damage zone” (WD, Fig. 1; e.g. Faulkner et

69 al., 2011; Savage & Brodsky, 2011; Vermilye & Scholz, 1998). The fracture density in this

70 zone dies off as a power law with distance from the fault (e.g., Ostermeijer et al., 2020

71 and references therein). The dilatant damage zone width increases linearly with fault

72 displacement, and typically levels out at several hundred meters for fault displacements

73 exceeding several hundred meters (Savage & Brodsky, 2011); (iii) Including and

74 extending beyond the dilatant damage zone is what we call the “shear deformation

75 zone” (WS; Fig. 1) which is defined by a region of enhanced seismicity, first pointed out

3 76 by Powers & Jordan (2010). This zone shows a region of high seismic activity near the

77 fault with a power law fall-off beyond a corner at WS1 to a full half-width of WS2 (Fig. 1).

78

79 The above definitions allow us to distinguish two types of damage zones: the “dilatant

80 damage zone” dominated by volumetric strains, and the “shear deformation zone”

81 dominated by shear strains. The tensile (Mode I) cracks in the dilatant damage zone are

82 dilatant cracks that align parallel to the maximum compression direction and

83 perpendicular to the minimum principal stress. Hence the orientation of cracks provides

84 evidence for the several different mechanisms responsible for their formation (Wilson et

85 al., 2003). The shear deformation zone is characterized by secondary faults (Mode II and

86 III cracks) and hence are oriented parallel to the maximum Coulomb stress. For example,

87 in the case of a strike-slip fault, this zone is defined by a conjugate set of secondary faults

88 (Little, 1995), in which one set is parallel to the primary fault.

89

90 The evolution of these three zones in figure 1 defines what is called fault maturity. The

91 three zones can be viewed as regions controlled by wear processes, and the fault

92 structural maturity can hence be measured by its degree of wear, which depends

93 primarily on the net fault displacement, assuming a constant slip-vector throughout the

94 fault’s history. However, previous studies have shown that, in the absence of data on net

95 fault displacement, several other fault parameters such as the fault initiation age and the

96 geological slip rate can be also used as a proxy of net displacement in evaluating the

97 overall maturity of the fault (e.g. Choy et al., 2006; Choy & Kirby, 2004; Dolan &

98 Haravitch, 2014; Hecker et al., 2010; Ikari et al., 2011; Manighetti et al., 2007; Niemeijer

99 et al., 2010; Perrin, Manighetti, Ampuero, et al., 2016; Stirling et al., 1996; Wesnousky,

100 1988). As these parameters increase, the fault grows and becomes more “mature”. Prior

4 101 studies have suggested that the structural maturity may have a strong impact on

102 earthquake behavior, such as magnitude, stress drop, distribution of slip, rupture

103 velocity, ground motion amplitude, and number of ruptured segments (e.g., Cao & Aki,

104 1986; Dolan & Haravitch, 2014; Hecker et al., 2010; Malagnini et al., 2010; Manighetti et

105 al., 2007; Perrin, Manighetti, Ampuero, et al., 2016; Radiguet et al., 2009; Stirling et al.,

106 1996; Wesnousky, 1988).

107

108 The widths of the fault core and dilatant damage zones saturate at fault lengths

109 comparable to the seismogenic thickness, due to the switch from crack like to pulse like

110 earthquake propagation (Ampuero & Mao, 2017). For larger faults, the evolution of fault

111 maturity involves only changes in the shear deformation zone. In this paper we are

112 concerned with the scaling of large faults (i.e., which breach the brittle seismogenic

113 width) and their associated large earthquakes, so we are only concerned with the shear

114 deformation zone. There is evidence that indicates that large faults become smoother

115 with net displacement (Stirling et al., 1996; Wesnousky, 1988). This smoothing is

116 probably the prime attribute of fault maturity. Here we show that the width of the shear

117 deformation zone of large faults decreases with fault displacement, as a consequence of

118 this smoothing.

119

120 Precise earthquake locations can be used to image the internal structure of fault and the

121 zone of brittle deformation, often at a resolution similar to field observations (e.g.

122 Powers & Jordan, 2010; Hauksson, 2010; Valoroso et al., 2014). Powers & Jordan (2010)

123 studied the association of small earthquakes with large faults in California. They found

124 that the frequency of small earthquakes is highest in a narrow region surrounding faults

125 and then falls off as a power law at greater distances. They modeled this behavior with a

5 126 fault model with rough (fractal) topography (Dieterich & Smith, 2009), showing that

127 such a rough fault model would produce high stresses near the fault that could account

128 for the seismicity. The earthquakes they used was from the interseismic period of the

129 faults. Although stacked profiles from many fault sections show the association of

130 seismicity with the faults, individual sections and map views, both in Powers & Jordan

131 (2010) and Hauksson (2010) show that such a tight correlation with the fault is not

132 typical for individual fault segments, which often display a wide variability in

133 distributions. As Hauksson (2010) observed, this variability is likely due to many

134 factors, such as heterogeneity in lithology, the effect of nearby faults both mapped and

135 unmapped, and fault offsets and bends.

136

137 As Hauksson (2010) observed further, aftershocks of large earthquakes, in contrast,

138 mostly show a tight clustering around the main fault. At depth, they are mainly

139 distributed at the edges of regions experiencing high coseismic slip and in surrounding

140 areas of low coseismic slip (e.g., Mendoza & Hartzell, 1988). Also, earthquakes that occur

141 during interseismic periods often delineate the edges of stuck patches sometimes

142 surrounded by creeping areas, and potential rupture surfaces at depth (e.g., Perrin et al.,

143 2019; Schaff et al., 2002; Waldhauser et al., 2004). In both cases, the majority of

144 aftershocks of large earthquakes sample a portion of the main fault zone structure at

145 depth at the edge of the rupture surface, and are often used to delineate it. In an

146 earthquake model with a smooth rupture surface, stress-drop increases to a maximum

147 near the fault surface, thus precluding aftershocks in the near-fault region (Kostrov &

148 Das, 1984). Near-fault aftershocks therefore are an indication of a rough fault, as near-

149 fault stresses are generated by dynamic slip on rough topography. They are greatest

150 right after the mainshock, after which they are relaxed by aftershocks and other

6 151 relaxation mechanisms. In this study, in order to estimate the roughness of active faults,

152 and how they evolve with displacement, we use high-precision aftershock locations of

153 large earthquakes on faults with different net displacements.

154

155 2. Data analysis

156 2.1 Width of the shear deformation zone from aftershock locations

157 We use high-resolution earthquake catalogs available in the literature to analyze the

158 aftershock distribution of seven large (Mw≥6) continental strike slip earthquakes that

159 occurred on faults with a wide range of net displacement: the 1984 Morgan Hill, 2004

160 Parkfield, and 2014 South Napa earthquakes in northern and central California

161 (Waldhauser, 2009; Waldhauser & Schaff, 2008), and the 1987 Superstition Hills, 1992

162 Landers, 1999 Hector Mine, and 2010 El Mayor Cucapah earthquakes in Southern

163 California/Mexico (Hauksson et al., 2012). These earthquakes were selected because

164 high-precision aftershock catalogues, relocated by the double difference method

165 (Waldhauser & Ellsworth, 2000) and with relative lateral location errors derived from

166 bootstrap analysis that are typically less than ≤ 0.03 km on average (Supp. Fig. S1). In

167 order to perform a homogeneous analysis of all sequences, we selected all aftershocks

168 with ML≥1 that occurred within 2 months of each mainshock.

169

170 For each aftershock sequence we determine the three-dimensional fault geometry by

171 applying a principal component analysis (PCA) to all events within boxes that are

172 between 3 and 10 km long along strike and between 3 and 20 km wide across strike (i.e.

173 centered on the surface fault trace), stepping at 1 km intervals along the fault trace

174 (Perrin et al., 2019). For each box, we obtain a plane that best fits the locations of

175 aftershocks. For simplicity, we assume a constant dip of the calculated planes in each

7 176 box (see also Perrin et al., 2019). In general, the calculated fault plane parameters are in

177 good agreement with existing fault parameters derived from source inversion models

178 (Mai & Thingbaijam, 2014 and references therein; see details in Supp. Table 1 and Supp.

179 Fig. S2).

180

181 We then compute the orthogonal distance between each hypocenter and the calculated

182 fault plane segment, count the number of aftershocks as a function of distance from the

183 fault plane within bins of 0.03 km, and normalize the results by the total number of

184 aftershocks in each box (gray lines in Fig. 2b, d, e, f, g and Supp. Fig. S3). Finally, we

185 compute the mean of these normalized counts for each distance bin (black line in Fig. 2

186 and Supp. Fig. S3) to describe the smoothed distribution of aftershocks as a function of

187 distance from the fault plane for each of the seven earthquake sequences.

188

189 We use a power law function to fit the mean normalized aftershock distribution (red

190 curve in Fig. 3 and Supp. Fig. S4) to find the two parameters that describe the width of

191 the shear deformation zone. WS1 is the distance from the fault plane where the fall-off in

nd 192 number of events exhibits the maximum curvature (i.e., maximum in 2 derivative). WS2

193 is the distance from the fault plane where the number of events departs from the power

194 law fit and either flatten out (see Fig. 3a) or drop towards zero (see Fig. 3e). Both

195 parameters represent significant changes in event density, with WS2 indicating the

196 transition from the zone of active faulting to a largely undeformed crustal host rock (see

197 section 4.3 for a comparison with Powers and Jordan, 2010). This transition generally

198 occurs when less than 20% of all boxes used to count aftershocks include events (grey

199 circles in figure 3 and Supp. Fig. S4). We use this criterion to define WS2 for all aftershock

200 distributions considered here.

8 201

202 As examples of our approach, analyses of the 2004 Parkfield and 1999 Hector Mine

203 aftershocks are presented in figure 2 and 3 (see Supp. Fig. S3 and S4 for all earthquake

204 cases). The power law fit to the near-fault aftershock data at Parkfield (red curve, Fig.

205 3a) has a maximum curvature at about 0.06 km away from the fault plane (WS1). Beyond

206 WS2 = 0.56 km, the distribution departs from the red fit, implying that no significant

207 additional aftershock activity (ML≥1) related to the active fault zone is detected. In cases

208 with multiple ruptures on sub-parallel or sub-perpendicular faults such as Hector Mine,

209 Landers, Superstition Hills and El Mayor Cucapah, we estimated WS1 and WS2 for the

210 individual fault segments that broke during the earthquake (Fig. 2c, d, e, f, g, 3b, c, d, e

211 and Supp. Fig. S3 and S4) and then averaged the values so that a single value could be

212 assigned to each aftershock sequence. The variance in the parameters obtained for

213 individual segments is typically small. Table 1 lists WS1 and WS2 we obtained for the

214 seven earthquakes.

215

216 Also listed in Table 1 are data from the immature Awatere fault. This right-lateral strike-

217 slip fault is a first order splay of the Alpine fault of New Zealand. It crosses the coast at a

218 sea-cliff that offers an almost complete exposure of the entire fault zone from the fault

219 core through the shear deformation zone. This was mapped by Little (1995), who

220 showed that the shear deformation zone consists of a set of right-lateral and conjugate

221 left-lateral secondary faults that decreased in frequency with distance from the primary

222 fault in the same manner as the aftershocks do in our study. The right-lateral set of the

223 secondary faults is nearly sub-parallel to the main fault and the left-lateral set about 60°

224 from that. Values of WS1 and WS2 obtained from that data are indicated in Table 1. An

225 exposure at mid-crustal depths of a strike-slip fault in Austria shows just the same

9 226 orientation of secondary faults (Frost et al., 2009). These observations provide the

227 ‘ground truth’ for the structures upon which the near-fault aftershocks occur.

228

229 2.2 Fault parameters

230 We collect key parameters describing the degree of evolution of the faults that broke

231 during the selected earthquakes. Since faults propagate laterally through time, their

232 structural maturity varies also along strike (Perrin, Manighetti, & Gaudemer, 2016;

233 Perrin, Manighetti, Ampuero, et al., 2016). Thus, it is necessary to use fault parameters

234 that describe the local fault maturity where the rupture occurred. This is particularly

235 true for long faults, such as the , for which fault initiation age varies

236 greatly along the fault, from ancient fault sections in Central California (24 to 29 Ma at

237 Parkfield; e.g. Atwater & Stock, 1998; Liu et al., 2010) to younger fault sections in

238 (<12Ma, e.g. Powell & Weldon, 1992; Sims, 1993), and which

239 therefore have different local net displacements. Table 1 presents the fault parameters

240 used in this study. The seven fault sections span a wide range of structural maturity,

241 with various initiation ages (1.1 to 29 Ma), cumulative displacements (1 to 315 km) and

242 geological slip rates (0.2 to 26 mm/yr).

243

244 Some fault parameters are not fully constrained, especially for the fault associated with

245 the South Napa earthquake. The corresponds to the northern extension

246 of the , which transfers its slip northwards via the Contra Costa shear

247 zone (Brossy et al., 2010; Catchings et al., 2016; Unruh et al., 2002). We estimated its

248 cumulative slip using a linear fault tip model (Scholz & Lawler, 2004). The West Napa

249 Fault terminates about 35 km north of the South Napa earthquake, and the tip taper for

250 interacting faults is 0.1-0.8 (Scholz & Lawler, 2004). Combining these values give a total

10 251 slip range of 3.5-28 km (Table 1). As the West Napa Fault is genetically linked to the

252 Calaveras fault and situated at its northern tip, we assume that it formed after the

253 formation of the main Calaveras fault (about 12 Ma; Wakabayashi, 1999) and before the

254 younger surrounded Greenville and Green Valley faults situated to the east (about 6 Ma

255 ago; Wakabayashi, 1999). The slip rate on the West Napa fault is estimated to exceed 1

256 mm/yr, based on geomorphic evidence (Wesling & Hanson, 2008) and to be 4±3 mm/yr

257 as derived from geodetic data (d’Alessio, 2005; Field et al., 2015). For clarity, the

258 estimated geological parameters of the West Napa fault are represented as open

259 symbols in figure 5.

260

261 3. Results

262 We investigate the relationship between the aftershock distributions, as characterized

263 by WS1 and WS2, and the various, independently derived fault parameters (Table 1). Our

264 first observation is that WS1 (i.e., the half distance from the calculated fault plane where

265 the aftershock rate saturates) and WS2 (i.e., the full half-width of the shear deformation

266 zone) correlate with each other (Fig. 4), indicating that these two parameters are not

267 independent and that the shape of the shear deformation zone grows in a self-similar

268 way as

269 WS1=(0.2±0.1)*WS2

270 In figure 5 we plot WS1 and WS2 as a function of various fault parameters. When plotted

271 against cumulative fault displacement (Fig. 5a and 5b), they both indicate that the width

272 of the shear deformation zone (as characterized by WS1 and WS2) decreases with net

273 displacement as a power law with an exponent of -0.54 and -0.39 for WS1 and WS2,

274 respectively. We note that the field measurements across the Awatere fault (red symbol

275 in Fig. 5) are in good agreement with the trends highlighted from seismological data.

11 276 This supports the interpretation that the shear deformation zone defined by the width of

277 the aftershock zone corresponds to the width of the zone of active secondary faulting.

278

279 Furthermore, we observe a decrease of WS1 and WS2 with increasing fault initiation age

280 (fig. 5c and 5d) and geological slip rate (fig. 5e and 5f), showing a wider shear

281 deformation zone for younger fault sections and slower slip rates. This is also in good

282 agreement with geological observations across the Awatere fault. We note that the

283 power law exponents in figure 5 should be used with caution given the occasionally

284 large uncertainties in the fault parameters.

285

286 Figure 6a shows the near-fault aftershock distribution as a function of fault

287 perpendicular distance for each of the seven earthquake sequences (see also black

288 curves in Fig. 2 and Supp. Fig. S3). These curves show a clear pattern: intermediate to

289 mature fault sections (warm colors in fig. 6) are characterized by aftershocks

290 concentrated close to the fault plane. The rapid fall-off in activity away from the fault

291 plane describes a narrow deformation zone at the scale of hundreds of meters. In

292 contrast, intermediate to immature faults (cold colors in fig. 6) exhibit a wider

293 deformation zone at the kilometer scale where events are more widely distributed

294 within the surrounding medium (fig. 6a and b). The distributions remain proportional to

295 one another, indicating that the shape of the shear deformation zone remains constant.

296 Field measurements of the cumulative number of faults across the Awatere fault (black

297 squares, fig. 6a; Little, 1995) show a similar trend and corroborate our seismological

298 observations.

299

300

12 301 4. Discussion

302 4.1. The nature of the shear deformation zone and the smoothing of faults

303 In this study we show that, for fault displacements greater than 1 km, the width of the

304 shear deformation zone (defined by WS1 and WS2) decreases with fault displacement. So,

305 what determines the width of the zone of near-fault aftershocks that define the shear

306 deformation zone? If the fault is smooth, the mainshock would produce a stress shadow

307 centered at the fault and there would be no near-fault aftershocks (Kostrov & Das,

308 1984). Slip on a rough fault, on the other hand, would result in high stresses in a narrow

309 region adjacent to the fault, within which earthquakes could be generated. This was the

310 explanation used by Powers & Jordan (2010) to explain near-fault seismicity during the

311 interseismic period. There they used the static rough fault model of Dieterich & Smith

312 (2009).

313

314 Here we use instead the model of Aslam & Daub (2018) which calculates the stresses

315 resulting from dynamic ruptures propagating on a rough surface. They characterized the

316 fault as a self-affine fractal with Hurst exponent H and fault roughness measured by the

317 RMS of the ratios between the amplitude and the wavelength of the fault surface

318 topography (. This is a fairly realistic rendition of the observed topography of faults

319 (e.g. Candela et al., 2012). Their model predicts large changes in the Coulomb Failure

320 Function (CFF= ; of more than 20 MPa) within a well-defined narrow region, Wnf

321 close to the fault. The modeled receiver faults they assumed are parallel to the primary

322 fault, which is consistent with the orientation of most of the secondary faults observed

323 by Little (1995) and Frost et al. (2009). They find that the width of this near-fault region

324 of high stresses is insensitive to H but decreases with . Assuming that our observed

325 shear deformation zone represents the near-fault region of high stresses, this suggests

13 326 that the observed decrease of WS2 with fault displacement is the result of smoothing of

327 the fault with slip, similarly to observations made during laboratory experiments

328 (Goebel et al., 2017)

329

330 Aslam & Daub (2018) found that the half width of the near fault zone Wnf is ~ 2.7 km and

331 ~ 0.9 km for profiles with fault roughness RMS values  of 0.01 and 0.001, respectively.

332 Comparing these results with the range of WS2 in figure 5 indicates that more than an

333 order of magnitude of roughness change will be required to explain these observations.

334 This indicates a wear rate far greater than that observed for the roughness of individual

335 fault segments measured in the outcrop scale range for fault displacements in the range

336 10-1-103 m (Brodsky et al., 2011). Similar measurements made on fault segments in the

337 high porosity and friable Navajo and Entrada sandstones of Utah (Dascher-Cousineau et

338 al., 2018) indicate greater smoothing rates but predict fault roughness at our scales

339 orders of magnitude smoother than indicated by our results as interpreted with Aslam &

340 Daub (2018). However, faults are composed of many segments or sub-faults at many

341 scales (Ferrill et al., 1999; Klinger, 2010; Manighetti et al., 2015; Scholz, 1998; Wallace,

342 1989) and of nested sub-faults offset from one another (Ben-Zion & Sammis, 2003; de

343 Joussineau & Aydin, 2009; Segall & Pollard, 1980). The length distribution of sub-faults

344 follows a power law (Scholz, 1998) and they are offset by jogs that are self-similar (de

345 Joussineau & Aydin, 2009). The combination of the latter relations would produce a

346 fractal topography at a hierarchy higher than that of the roughness of the individual sub-

347 fault. A fault smoothing relationship based on roughness at this supra-segment level,

348 measured as fault step spacing, shows an approximately inverse linear reduction of fault

349 roughness with fault displacement (Stirling et al., 1996). This is a much higher rate of

350 smoothing than that of individual fault segment surfaces, and is comparable to that we

14 351 infer from the narrowing of the shear deformation zone. Converting the roughness

352 parameter step frequency of Stirling et al. (1996) to  using the typical step width 1-2

353 km (Wesnousky, 1988) and their smoothing curve allows us to recast the two fiducial

354 points of Aslam & Daub, 2018 from (Wnf ,to (Wnf , D) coordinates (i.e. values of 0.001

355 and 0.01 corresponding to fault displacements “D” of 100 and 10 km, respectively).

356 These are plotted in Figure 7 (blue circles) where they are compared with our WS2

357 measurements from figure 5b (squares). The good agreement supports the hypothesis

358 that narrowing of the shear deformation zone with fault displacement is the

359 consequence of the smoothing of the fault with wear (see also discussion in section 4.3).

360

361 4.2. The effect of fault maturity on earthquake properties

362 Once a fault reaches the seismogenic thickness, with a length of ~ 10 km and slip ~ 100

363 m, its core and dilatant damage zones are fully formed and further maturation occurs by

364 the fault becoming smoother through the processes of frictional wear. It follows that the

365 characteristics of earthquakes may also evolve with the smoothing of the fault. It has

366 been suggested, for example, that earthquake stress drop, apparent stress, and radiation

367 efficiency vary with fault maturity (Choy et al., 2006; Hecker et al., 2010; Ross et al.,

368 2018). These are not entirely independent parameters. The radiation efficiency, 휂푅, is

369 given by:

퐸푅 2휇퐸푅 2휎푎 370 휂푅 = = = , 퐸푅+퐸퐺 Δ휎푀표 Δ휎

371 where 퐸푅 is radiated energy, 퐸퐺 is energy dissipated in damage, Δ휎 is (static) stress

372 drop, 휎푎is apparent stress, 푀표is seismic moment, and 휇 is shear modulus. Here we focus

373 on stress drop, which is the most commonly reported of these properties.

15 374 Five of the earthquakes in our study have three or more estimates of stress drop (Δ휎).

375 Following the procedure of Hardebeck (2020), we assume that log10(Δ휎) has a normal

376 distribution (Baltay et al., 2011) and calculate its mean and the standard error of the

377 mean, as reported in Table 2. The corresponding values of mean Δ휎 are plotted vs. the

378 cumulative fault slip in figure 8. The correlation Δ휎  D-0.23, compared to our earlier

−0.39 379 finding that 푊푆2 ∝ 퐷 implies that stress drop scales with 푊푆2 and hence, from figure

380 7, that stress drop scales with fault roughness. Hecker et al. (2010) measured the

381 maximum slip to length ratio of prehistoric earthquake scarps on intraplate dip-slip

382 faults in the western U.S. They found that this ratio, an indicator of stress drop, tends to

383 decrease with net fault displacement, consistent with our finding.

384

385 However, net fault slip usually correlates with slip rate (see Fig. 5), so we also find a

386 similar correlation between stress drop and that parameter, as shown in figure 9a (the

387 relation between stress drop and fault initiation age is shown in Supp. Fig. S5). This is

388 shown also in log-linear coordinates in figure 9b to compare it with the form expected

389 from fault healing. The correlation of stress drop with the log of slip rate or its inverse,

390 the recurrence time, is well known (Kanamori & Allen, 1986; Scholz et al., 1986) as is its

391 correlation with strain rate (Hauksson, 2015). The size of the effect seen in figure 9b is

392 about -2 MPa/decade in slip rate, in rough agreement with the other studies.

393 Scholz et al. (1986) pointed out that this behavior could either reflect a rate effect due to

394 fault healing as predicted by the rate- and state-dependent friction laws (Dieterich,

395 1972) or a roughness effect in which friction increases with surface roughness, as

396 indicated by several experimental rock friction studies (Biegel et al., 1992; Byerlee,

397 1967), or some combination of the two.

398

16 399 Observations of repeating earthquakes in which seismic moment increased with

400 recurrence time promised a separation of these effects (Marone et al., 1995; Vidale et al.,

401 1994). Simulations with one dimensional spring-slider models with rate-state friction

402 indicated that these data could be explained with fault healing at a rate of 1-3

403 MPa/decade. However, subsequent studies using a 3D model of a repeating earthquake

404 occurring on a velocity-weakening patch imbedded within a creeping velocity-

405 strengthening fault showed that the recurrence time – moment scaling could occur

406 without any increase in stress drop (Chen & Lapusta, 2009). Observations at Parkfield,

407 California, showed both positive and negative moment-recurrence time scaling of

408 repeating earthquakes (Chen et al., 2010). Both types of scaling can be predicted with

409 the Chen & Lapusta, (2009) model by varying the ratio of nucleation to patch size.

410

411 In the Chen & Lapusta, (2009) model, aseismic slip propagates from the external

412 creeping region into the velocity-weakening patch such that only a fraction of its slip is

413 seismic. The ratio of seismic to aseismic slip in the velocity-weakening patch is sensitive

414 to parameters such as the radius of the velocity-weakening patch and its ratio to the

415 nucleation radius. As a result, it behaves very differently from a simple velocity-

416 weakening block slider model in which the slip is entirely seismic. The recurrence time –

417 moment scaling can be reproduced by varying the patch radius without varying stress

418 drop. In our problem, however, we are considering the fault to be fully coupled

419 seismically so that its slip is entirely seismic. In that case, the simple spring slider model

420 should be more applicable. If so, the analysis of Beeler et al. (2001) finds that stress drop

421 should increase logarithmically with recurrence time at about 0.5 to 1.5 MPa per decade,

422 assuming typical laboratory values for the friction parameters a and b. This is in

17 423 reasonable agreement with the slope in Fig. 9b of -2 MPa/decade in slip rate, although,

424 as we said earlier, the effect of fault roughness on stress drop cannot be ruled out.

425

426 4.3. Strengths, limitations and comparison with other studies.

427 In our study, we use aftershock distributions to study how the early postseismic events

428 are reactivating the long-term damage zone around active faults. We define WS2 as the

429 distance from the fault plane where the mean aftershock distributions start to flatten

430 out or drop abruptly (Fig. 3 and Supp. Fig. S4). Either the flattening or the abrupt

431 termination of the number of events with distance away from the fault would indicate

432 that there are no faults or fractures big enough to accommodate a large number of

433 events. Powers and Jordan (2010) observed a flattening of the interseismic event

434 distribution away from the fault which they attributed to background seismicity. Since

435 we are only using 2 months long aftershock sequences, with significantly larger number

436 of events compared to the interseismic period, we don’t systematically observe the

437 curve flattening and cannot reliably estimate the off-fault background seismicity level as

438 Powers & Jordan (2010) did for interseismic periods. If significant off-fault seismicity

439 exists during our observational period, then it may have occurred shortly after the

440 mainshock, and went undetected due to the low signal to noise ratio in the coda of the

441 mainshocks. However, a few of our aftershock distributions show curve flattening (e.g.

442 Parkfield, South Napa, Morgan Hill). These sequences occurred on intermediate to

443 mature faults where creep has been observed (e.g. Harris & Segall, 1987; Schaff et al.,

444 2002; Xu et al., 2018). Our hypothesis is that these aftershock distributions have fewer

445 absolute number of events because they are related to smaller mainshocks (M~6), and

446 also lower stress drops (see Fig. 8). Moreover, the presence of creeping sections might

447 increase the regional background seismicity level.

18 448

449 In this study, we selected aftershocks of Ml>1, thus including events with magnitudes

450 that are below magnitudes of completeness, Mc, levels for most aftershock sequences,

451 especially in southern California. Assuming that small events are randomly distributed

452 in the stress field, recovering and including all small events in our analysis would

453 increase the number of events in all boxes. This would result in shifts along the y axis in

454 figure 6a, but would not change the shape of the curves. It might lead to slight increases

455 in the WS2 values, which therefore should be considered here as lower bounds of our

456 shear deformation zone widths. This is in good agreement with figure 7, where some

457 measurements corresponding to southern California sequences seem to be under-

458 estimated compared to the expected theoretical models.

459

460 Our estimates of the shear deformation zone width might also be affected by the

461 aftershock productivity, which depends on the magnitude of the mainshock. The number

462 of aftershocks does increase with magnitude, but also linearly with fault area (Wetzler et

463 al., 2016; Yamanaka & Shimazaki, 1990), and therefore the density of aftershocks per

464 unit area is constant. This is what we show in supplementary figure S6, where the mean

465 number of events in boxes per km2 are quite similar and do not show any magnitude

466 dependency (Parkfield and Morgan Hill being slightly off because of the creeping areas

467 that bound the rupture zones, which produce abundant interseismic earthquakes on the

468 fault, contaminating the data at small distances). The apparent correlation between

469 mainshock magnitude and WS2 in our study results rather from the happenstance that

470 the larger earthquakes happen to have occurred on low maturity intraplate faults

471 whereas the smaller ones occurred on mature faults.

472

19 473 As mentioned earlier, Powers & Jordan (2010) estimated the normal distance of the

474 earthquakes away from a single vertical fault plane inferred from the surface trace of the

475 fault. They defined the zone of shear deformation from the near fault seismicity during

476 the interseismic period and averaged over the entire fault length, possibly neglecting

477 local variations in fault plane orientation and dip, leading to wider zones compared to

478 our results. In addition, geological heterogeneity over long distances along the fault and

479 other effects (such as triggering or regional stress release caused by other earthquakes)

480 may play a role, as pointed out by Hauksson (2010). Finally, our approach of fitting a

481 single fault plane to the local aftershock distribution as we step along the fault assumes

482 that the rupture occurred mainly on one fault strand and that aftershocks are

483 homogeneously distributed in the surrounded medium. While individual secondary

484 faults are distinguishable around mature fault sections (see for instance oblique linear

485 faults around Morgan Hill in Supp. Fig. S3), it is more difficult to identify them around

486 immature fault sections, within the broader zone of deformation, especially near the

487 rupture tips (see the northern part of the El Mayor Cucapah earthquake for instance;

488 Supp. Fig. S3).

489

490 Previous studies have estimated damage zone width or low velocity zones across several

491 faults in California using various seismic and/or geodetic methods (see Yang, 2015 for a

492 review). We compare some measurements with our results in Supp. Table S2. First, we

493 note that, different methods result in different estimates of “damage zone” widths.

494 Secondly, all studies using seismic fault zone trapped waves (FZTW) techniques find a

495 similar low velocity zone that is 100-200 m wide, regardless the fault structural

496 maturity. The FZTW derived width correspond to the constant width of dilatant damage

497 zones around large faults, already described by geological studies (e.g., Faulkner et al.,

20 498 2011; Savage & Brodsky, 2011; also called the inner damage zone in Perrin, Manighetti,

499 Ampuero, et al., 2016). However, other seismic and geodetic studies (sometimes used

500 together; see Cochran et al., 2009), find greater damage zone widths, that are in general

501 agreement with our results. This discrepancy between results from different methods

502 was also pointed out by Spudich & Olsen (2001). We tested the relations found in Figure

503 5a and 5b to deduce the size of the shear damage zone of the San Jacinto fault, for which

504 both a narrow damage zone using FZTW methods (e.g., Lewis et al., 2005) and wider

505 damage zone using tomography (e.g., Allam et al., 2014) have been described. Assuming

506 a cumulative slip of 24 km (Sharp, 1967), we obtain a half width of the shear

507 deformation zone ranging from 0.2 (i.e. WS1) to 1.5 km (i.e. WS2), which is in better

508 agreement with tomographic studies (half width of 1.5 to 2.5 km; Allam et al., 2014).

509 Thus, we believe that the studies using FZTW are mainly sensitive to the most damaged

510 inner part of the fault zone, also called the dilatant damage zone, which we don’t seem to

511 be able to resolve with our aftershock data. The other seismic and geodetic studies

512 found damage zone widths that are between 2(WS1) and 2(WS2), which we define as the

513 inner and outer bounds of the shear deformation zone.

514

515 A recent geodetic study from Scott et al. (2018) has shown that the dilatant damage zone

516 is equivalent to slightly wider than the shear deformation zone. These results come from

517 surface measurements (i.e., LiDAR) along the active fault zone responsible for the 2016

518 Kumamoto earthquake. It is possible that extensional strains are greater at the surface

519 than at depth. As cracks close down rapidly under pressure, dilatant strains in the

520 dilatant damage zone will decrease more rapidly with depth (Yoshioka & Scholz, 1989)

521 than the shear strains. This may explain why their results from surface measurements

522 differ from ours measured at depth.

21 523

524 In our study we average the near-fault aftershock distributions assuming that the degree

525 of fault maturity is locally constant along each broken fault sections we investigate.

526 However, it has been shown that the fault maturity can vary along strike and this can

527 affect the heterogeneity of fracture density around the fault core (Ostermeijer et al.,

528 2020), the distribution of secondary faults at greater distances away from the fault

529 (Perrin, Manighetti, & Gaudemer, 2016) and finally the behavior of earthquakes (e.g.,

530 Cappa et al., 2014; Huang, 2018; Perrin, Manighetti, Ampuero, et al., 2016).

531 Consequently, for long or multiple-fault earthquake ruptures, the fault normal

532 distribution of aftershocks could change along strike, and present a shear deformation

533 zone that widens towards the most immature parts of the rupture. Such a behavior

534 would be in good agreement with mapped faults at the surface showing that the off-fault

535 damage zone widens in the direction of long-term fault propagation (e.g., Manighetti et

536 al., 2001; Perrin, Manighetti, & Gaudemer, 2016). Future work is needed to relate such

537 observations to the occurrence of seismicity.

538

539 5. Conclusion

540 Our study presents strong correlations between independent datasets that describe the

541 near-fault distribution of aftershocks following large earthquakes and geological fault

542 parameters such as cumulative displacement, initiation age and slip rate. We find that

543 for large faults, defined as those that have ruptured the entire brittle crust, the zone of

544 active shear deformation narrows as a power law with cumulative fault displacement,

545 hence with fault maturity. This result is predicted by a dynamic rough fault model

546 combined with an empirical roughness/displacement relation, indicating that the

547 narrowing of the shear deformation zone is the result of fault smoothing with

22 548 cumulative displacement. Earthquake stress-drop also decreases with fault

549 displacement and hence fault smoothness or slip rate.

550

551 ACKNOWLEDMENTS

552 This work is supported by the Brinson Foundation (CP) and the National Science

553 Foundation (NSF-EAR #1520680). Earthquake catalog data are available at

554 http://ddrt.ldeo.columbia.edu and https://scedc.caltech.edu/research-

555 tools/altcatalogs.html. We thank the editor R. Abercrombie, associate editor M. Cooke

556 and reviewers E. Nissen, N. Abolfathian and K. Dascher-Cousineau for their fruitful

557 comments that greatly improved the manuscript. This paper is LDEO contribution 8477.

558

559 REFERENCES

560 Allam, A. A., Ben-Zion, Y., Kurzon, I., & Vernon, F. (2014). Seismic velocity structure in the

561 Hot Springs and Trifurcation areas of the , California, from

562 double-difference tomography. Geophysical Journal International, 198(2), 978–999.

563 https://doi.org/10.1093/gji/ggu176

564 Allmann, B. P., & Shearer, P. M. (2009). Global variations of stress drop for moderate to

565 large earthquakes. Journal of Geophysical Research: Solid Earth, 114(B1).

566 https://doi.org/10.1029/2008JB005821

567 Ampuero, J.-P., & Mao, X. (2017). Upper limit on damage zone thickness controlled by

568 seismogenic depth. In M. Thomas, T. Mitchell, & H. S. Bhat (Eds.), Invited chapter in

569 “Fault Zone Dynamic Processes: Evolution of Fault Properties During Seismic

570 Rupture.”

571 Ancheta, T. D., Darragh, R. B., Stewart, J. P., Seyhan, E., Silva, W. J., Chiou, B. S.-J., et al.

572 (2014). NGA-West2 database. Earthquake Spectra, 30(3), 989–1005.

23 573 Aslam, K. S., & Daub, E. G. (2018). Effect of Fault Roughness on Aftershock Distribution:

574 Elastic Off-Fault Material Properties. Journal of Geophysical Research: Solid Earth,

575 123(11), 9689–9711. https://doi.org/10.1029/2018JB016214

576 Atwater, T., & Stock, J. (1998). Pacific North America plate tectonics of the Neogene

577 southwestern United States: An update. International Geology Review, 40(5), 375–

578 402.

579 Baltay, A. S., Ide, S., Prieto, G., & Beroza, G. (2011). Variability in earthquake stress drop

580 and apparent stress. Geophysical Research Letters, 38(6), n/a-n/a.

581 https://doi.org/10.1029/2011GL046698

582 Baltay, A. S., Hanks, T. C., & Abrahamson, N. A. (2019). Earthquake Stress Drop and Arias

583 Intensity. Journal of Geophysical Research: Solid Earth, 124(4), 3838–3852.

584 https://doi.org/10.1029/2018JB016753

585 Beeler, N. M., Hickman, S. H., & Wong, T. -f. (2001). Earthquake stress drop and

586 laboratory-inferred interseismic strength recovery. Journal of Geophysical Research:

587 Solid Earth, 106(B12), 30701–30713. https://doi.org/10.1029/2000JB900242

588 Ben-Zion, Y., & Sammis, C. G. (2003). Characterization of Fault Zones. Pure and Applied

589 Geophysics, 160(3), 677–715. https://doi.org/10.1007/PL00012554

590 Biegel, R. L., Wang, W., Scholz, C. H., Boitnott, G. N., & Yoshioka, N. (1992).

591 Micromechanics of rock friction 1. Effects of surface roughness on initial friction

592 and slip hardening in westerly granite. Journal of Geophysical Research, 97(B6),

593 8951. https://doi.org/10.1029/92JB00042

594 Blisniuk, K., Rockwell, T., Owen, L. A., Oskin, M., Lippincott, C., Caffee, M. W., & Dortch, J.

595 (2010). Late slip rate gradient defined using high-resolution

596 topography and 10Be dating of offset landforms on the southern San Jacinto Fault

597 zone, California. Journal of Geophysical Research: Solid Earth, 115(8).

24 598 https://doi.org/10.1029/2009JB006346

599 Brodsky, E. E., Gilchrist, J. J., Sagy, A., & Collettini, C. (2011). Faults smooth gradually as a

600 function of slip. Earth and Planetary Science Letters, 302(1–2), 185–193.

601 https://doi.org/10.1016/j.epsl.2010.12.010

602 Brossy, C., Kelson, K., & Ticci, M. (2010). Digital Compilation of Data for the Contra

603 CostaShear Zone for the Quaternary Fault Map Database:

604 Collaborative Research with William Lettis & Associates, Inc. and the US Geological

605 Survey. Final Technical Report submitted to the US Geological Survey NEHRP, Award

606 no. 07HQGR0063. https://doi.org/10.1017/CBO9781107415324.004

607 Byerlee, J. D. (1967). Theory of friction based on brittle fracture. Journal of Applied

608 Physics, 38, 2928–2934.

609 Candela, T., Renard, F., Klinger, Y., Mair, K., Schmittbuhl, J., & Brodsky, E. E. (2012).

610 Roughness of fault surfaces over nine decades of length scales. Journal of

611 Geophysical Research: Solid Earth, 117(8), 1–30.

612 https://doi.org/10.1029/2011JB009041

613 Cao, T., & Aki, K. (1986). Seismicity simulation with a rate- and state-dependent friction

614 law. Pure and Applied Geophysics, 124(3), 487–513.

615 https://doi.org/10.1007/BF00877213

616 Cappa, F., Perrin, C., Manighetti, I., & Delor, E. (2014). Off-fault long-term damage: A

617 condition to account for generic, triangular earthquake slip profiles. Geochemistry,

618 Geophysics, Geosystems, 15(4), 1476–1493.

619 https://doi.org/10.1002/2013GC005182

620 Catchings, R. D., Goldman, M. R., Li, Y. G., & Chan, J. H. (2016). Continuity of the West

621 Napa–Franklin Fault Zone Inferred from Guided Waves Generated by Earthquakes

622 Following the 24 August 2014 M w 6.0 South Napa Earthquake. Bulletin of the

25 623 Seismological Society of America, 106(6), 2721–2746.

624 https://doi.org/10.1785/0120160154

625 Chen, K. H., Bürgmann, R., Nadeau, R. M., Chen, T., & Lapusta, N. (2010). Postseismic

626 variations in seismic moment and recurrence interval of repeating earthquakes.

627 Earth and Planetary Science Letters, 299(1–2), 118–125.

628 https://doi.org/10.1016/j.epsl.2010.08.027

629 Chen, T., & Lapusta, N. (2009). Scaling of small repeating earthquakes explained by

630 interaction of seismic and aseismic slip in a rate and state fault model. Journal of

631 Geophysical Research: Solid Earth, 114(B1).

632 https://doi.org/10.1029/2008JB005749

633 Choy, G. L., & Kirby, S. H. (2004). Apparent stress, fault maturity and seismic hazard for

634 normal-fault earthquakes at subduction zones. Geophysical Journal International,

635 159(3), 991–1012. https://doi.org/10.1111/j.1365-246X.2004.02449.x

636 Choy, G. L., McGarr, A., Kirby, S. H., & Boatwright, J. (2006). An overview of the global

637 variability in radiated energy and apparent stress. In R. Abercrombie & E. Al. (Eds.),

638 Geophysical Monograph Series (pp. 43–57). Washington DC: AGU.

639 https://doi.org/10.1029/170GM06

640 Cochran, E. S., Li, Y. G., Shearer, P. M., Barbot, S., Fialko, Y., & Vidale, J. E. (2009). Seismic

641 and geodetic evidence for extensive, long-lived fault damage zones. Geology, 37(4),

642 315–318. https://doi.org/10.1130/G25306A.1

643 Critelli, S., & Nilsen, T. H. (2000). Provenance and stratigraphy of the Eocene Tejon

644 Formation, Western Tehachapi Mountains, San Emigdio Mountains, and Southern

645 San Joaquin Basin, California. Sedimentary Geology, 136, 7–27.

646 Crowell, J. C. (1979). The San Andreas fault system through time. Journal of the

647 Geological Society, 136, 293–302. https://doi.org/10.1144/gsjgs.136.3.0293

26 648 d’Alessio, M. A. (2005). Slicing up the : Block kinematics and fault

649 slip rates from GPS-derived surface velocities. Journal of Geophysical Research,

650 110(B6), B06403. https://doi.org/10.1029/2004JB003496

651 Dascher-Cousineau, K., Kirkpatrick, J. D., & Cooke, M. L. (2018). Smoothing of Fault Slip

652 Surfaces by Scale-Invariant Wear. Journal of Geophysical Research: Solid Earth,

653 123(9), 7913–7930. https://doi.org/10.1029/2018JB015638

654 Dibblee, T. W. J. (1961). Evidence of strike-slip movement on northwest-trending faults

655 in the western Mojave Desert, California. U.S. Geological Survey Professional Paper,

656 424-B, B197–B199.

657 Dieterich, J. H. (1972). Time-dependent friction in rocks. Journal of Geophysical Research,

658 77(20), 3690–3697. https://doi.org/10.1029/JB077i020p03690

659 Dieterich, J. H., & Smith, D. E. (2009). Nonplanar faults: Mechanics of slip and off-fault

660 damage. Pure and Applied Geophysics, 166, 1799–1815.

661 https://doi.org/10.1007/s00024-009-0517-y

662 Dokka, R. K. (1983). Geology Displacements on late Cenozoic strike-slip faults of the

663 central Mojave Desert, California Displacements on late Cenozoic strike-slip faults

664 of the central Mojave Desert, California, (May), 305–308.

665 https://doi.org/10.1130/0091-7613(1983)11<305

666 Dokka, R. K., & Travis, C. J. (1990). Late Cenozoic strike‐slip faulting in the Mojave

667 Desert, California. Tectonics. https://doi.org/10.1029/TC009i002p00311

668 Dolan, J. F., & Haravitch, B. D. (2014). How well do surface slip measurements track slip

669 at depth in large strike-slip earthquakes? The importance of fault structural

670 maturity in controlling on-fault slip versus off-fault surface deformation. Earth and

671 Planetary Science Letters, 388(C), 38–47.

672 https://doi.org/10.1016/j.epsl.2013.11.043

27 673 Dorsey, R. J., Axen, G. J., Peryam, T. C., & Kairouz, M. E. (2012). Initiation of the southern

674 Elsinore fault at ∼1.2 Ma: Evidence from the Fish Creek-Vallecito Basin, southern

675 California. Tectonics, 31(2), 1–21. https://doi.org/10.1029/2011TC003009

676 Faulkner, D. R., Mitchell, T. M., Jensen, E., & Cembrano, J. (2011). Scaling of fault damage

677 zones with displacement and the implications for fault growth processes. Journal of

678 Geophysical Research: Solid Earth, 116(5), 1–11.

679 https://doi.org/10.1029/2010JB007788

680 Ferrill, D. A., Stamatakos, J. A., & Sims, D. W. (1999). Normal fault corrugation:

681 Implications for growth and seismicity of active normal faults. Journal of Structural

682 Geology, 21, 1027–1038. https://doi.org/10.1016/S0191-8141(99)00017-6

683 Field, E. H., Biasi, G. P., Bird, P., Dawson, T. E., Felzer, K. R., Jackson, D. D., et al. (2015).

684 Long‐Term Time‐Dependent Probabilities for the Third Uniform California

685 Earthquake Rupture Forecast (UCERF3). Bulletin of the Seismological Society of

686 America, 105(2A), 511–543. https://doi.org/10.1785/0120140093

687 Fletcher, J. B., & McGarr, A. (2006). Distribution of stress drop, stiffness, and fracture

688 energy over earthquake rupture zones. Journal of Geophysical Research: Solid Earth,

689 111(B3), n/a-n/a. https://doi.org/10.1029/2004JB003396

690 Fletcher, K. E. K., Rockwell, T. K., & Sharp, W. D. (2011). Late Quaternary slip rate of the

691 southern Elsinore fault, Southern California: Dating offset alluvial fans via 230Th/U

692 on pedogenic carbonate. Journal of Geophysical Research: Earth Surface, 116(2).

693 https://doi.org/10.1029/2010JF001701

694 Frost, E., Dolan, J., Sammis, C., Hacker, B., Cole, J., & Ratschbacher, L. (2009). Progressive

695 strain localization in a major strike‐slip fault exhumed from midseismogenic

696 depths: Structural observations from the Salzach‐Ennstal‐Mariazell‐Puchberg

697 fault system, Austria. Journal of Geophysical Research, 114(B4), B04406.

28 698 https://doi.org/10.1029/2008JB005763

699 Garfunkel, Z. (1974). Model for the Late Cenozoic Tectonic History of the Mojave Desert ,

700 California , and for Its Relation to Adjacent Regions. Bulletin of the Geological Society

701 of America, 85(December), 1931. https://doi.org/10.1130/0016-

702 7606(1974)85<1931

703 Goebel, T. H. W., Kwiatek, G., Becker, T. W., Brodsky, E. E., & Dresen, G. (2017). What

704 allows seismic events to grow big?: Insights from b-value and fault roughness

705 analysis in laboratory stick-slip experiments. Geology, 45(9), 815–818.

706 https://doi.org/10.1130/G39147.1

707 Graham, S. A., Stanley, R. G., Bent, J. V., & Carter, J. B. (1989). Oligocene and Miocene

708 paleogeography of central California and displacement along the San Andreas fault.

709 Geological Society of America Bulletin, 101, 711–730.

710 Gurrola, L. D., & Rockwell, T. K. (1996). Timing and slip for prehistoric earthquakes on

711 the Superstition Mountain Fault, , southern California. Journal of

712 Geophysical Research: Solid Earth, 101(B3), 5977–5985.

713 https://doi.org/10.1029/95JB03061

714 Hanks, T. C., & Bakun, W. H. (2008). M-logA Observations for Recent Large Earthquakes.

715 Bulletin of the Seismological Society of America, 98(1), 490–494.

716 https://doi.org/10.1785/0120070174

717 Hardebeck, J. L. (2020). Are the Stress Drops of Small Earthquakes Good Predictors of

718 the Stress Drops of Moderate‐to‐Large Earthquakes? Journal of Geophysical

719 Research: Solid Earth, 125(3). https://doi.org/10.1029/2019JB018831

720 Harris, R. A., & Segall, P. (1987). Detection of a Locked Zone at Depth on the Parkfield,

721 California, Segment of the San Andreas Fault. Journal of Geophysical Research,

722 92(10), 7945–7962. https://doi.org/10.1029/JB092iB08p07945

29 723 Hauksson, E. (2010). Spatial Separation of Large Earthquakes, Aftershocks, and

724 Background Seismicity: Analysis of Interseismic and Coseismic Seismicity Patterns

725 in Southern California. Pure and Applied Geophysics, 167(8–9), 979–997.

726 https://doi.org/10.1007/s00024-010-0083-3

727 Hauksson, E. (2015). Average Stress Drops of Southern California Earthquakes in the

728 Context of Crustal Geophysics: Implications for Fault Zone Healing. Pure and Applied

729 Geophysics, 172(5), 1359–1370. https://doi.org/10.1007/s00024-014-0934-4

730 Hauksson, E., Jones, L. M., & Hutton, K. (2002). The 1999 Mw 7.1 Hector Mine, California,

731 earthquake sequence: Complex conjugate strike-slip faulting. Bulletin of the

732 Seismological Society of America, 92(4), 1154–1170.

733 https://doi.org/10.1785/0120000920

734 Hauksson, E., Yang, W., & Shearer, P. M. (2012). Waveform relocated earthquake catalog

735 for Southern California (1981 to June 2011). Bulletin of the Seismological Society of

736 America, 102(5), 2239–2244. https://doi.org/10.1785/0120120010

737 Hecker, S., Dawson, T. E., & Schwartz, D. P. (2010). Normal-Faulting Slip Maxima and

738 Stress-Drop Variability: A Geological Perspective. Bulletin of the Seismological

739 Society of America, 100(6), 3130–3147. https://doi.org/10.1785/0120090356

740 Huang, Y. (2018). Earthquake Rupture in Fault Zones With Along-Strike Material

741 Heterogeneity. Journal of Geophysical Research: Solid Earth, 123(11), 9884–9898.

742 https://doi.org/10.1029/2018JB016354

743 Hudnut, K. W., & Sieh, K. (1989). Behavior of the Superstition Hills fault during the past

744 330 years. Bulletin of the Seismological Society of America, 79(2), 304–329.

745 Ikari, M. J., Marone, C., & Saffer, D. M. (2011). On the relation between fault strength and

746 frictional stability. Geology, 39(1), 83–86. https://doi.org/10.1130/G31416.1

747 Jachens, R. C., Langenheim, V. E., & Matti, J. C. (2002). Relationship of the 1999 Hector

30 748 Mine and 1992 Landers fault ruptures to offsets on neogene faults and distribution

749 of late Cenozoic basins in the eastern California shear zone. Bulletin of the

750 Seismological Society of America, 92(4), 1592–1605.

751 https://doi.org/10.1785/0120000915

752 de Joussineau, G., & Aydin, A. (2009). Segmentation along strike-slip faults revisited.

753 Pure and Applied Geophysics, 166(10–11), 1575–1594.

754 https://doi.org/10.1007/s00024-009-0511-4

755 Kanamori, H., & Allen, C. (1986). Earthquake repeat time and average stress drop, in

756 Earthquake Source Mechanics. Geophysical Monograph-AGU, 37, 227–236.

757 Kaverina, A., Dreger, D. S., & Price, E. J. (2002). The Combined Inversion of Seismic and

758 Geodetic Data for the Source Process of the 16 October 1999 Mw 7.1 Hector Mine,

759 California, Earthquake. Bulletin of the Seismological Society of America, 92(4), 1266–

760 1280. https://doi.org/10.1785/0120000907

761 Kim, A., & Dreger, D. S. (2008). Rupture process of the 2004 Parkfield earthquake from

762 near-fault seismic waveform and geodetic records. Journal of Geophysical Research,

763 113(B7), B07308. https://doi.org/10.1029/2007JB005115

764 Kirby, S. M., Janecke, S. U., Dorsey, R. J., Housen, B. A., Langenheim, V. E., McDougall, K. A.,

765 & Steely, A. N. (2007). brawley and ocotillo formations: Evidence for

766 initial strike-slip deformation along the San Felipe and San Jacinto fault zones,

767 southern California. Journal of Geology, 115(1), 42–62.

768 https://doi.org/10.1086/509248

769 Klinger, Y. (2010). Relation between continental strike-slip earthquake segmentation

770 and thickness of the crust. Journal of Geophysical Research, 115(B7), B07306.

771 https://doi.org/10.1029/2009JB006550

772 Kostrov, B. V., & Das, S. (1984). Evaluation of stress and displacement fields due to an

31 773 elliptical plane shear crack. Geophysical Journal International, 78(1), 19–33.

774 https://doi.org/10.1111/j.1365-246X.1984.tb06469.x

775 Lewis, M. A., Peng, Z., Ben-Zion, Y., & Vernon, F. L. (2005). Shallow seismic trapping

776 structure in the San Jacinto fault zone near Anza, California. Geophysical Journal

777 International, 162(3), 867–881. https://doi.org/10.1111/j.1365-

778 246X.2005.02684.x

779 Little, T. A. (1995). Brittle deformation adjacent to the Awatere strike-slip fault in New

780 Zealand: Faulting patterns, scaling relationships, and displacement partitioning.

781 Geological Society of America Bulletin, 107(11), 1255–1271.

782 https://doi.org/10.1130/0016-7606(1995)107<1255:BDATTA>2.3.CO;2

783 Liu, M., Wang, H., & Li, Q. (2010). Inception of the eastern California shear zone and

784 evolution of the Pacific- boundary: From kinematics to

785 geodynamics. Journal of Geophysical Research, 115(B7), B07401.

786 https://doi.org/10.1029/2009JB007055

787 Lutz, A. T., Dorsey, R. J., Housen, B. A., & Janecke, S. U. (2006). Stratigraphic record of

788 Pleistocene faulting and basin evolution in the Borrego Badlands, San Jacinto fault

789 zone, Southern California. Bulletin of the Geological Society of America, 118(11–12),

790 1377–1397. https://doi.org/10.1130/B25946.1

791 Ma, S., Custódio, S., Archuleta, R. J., & Liu, P. (2008). Dynamic modeling of the 2004 Mw

792 6.0 Parkfield, California, earthquake. Journal of Geophysical Research: Solid Earth,

793 113(2), 1–16. https://doi.org/10.1029/2007JB005216

794 Mai, P. M., & Thingbaijam, K. K. S. (2014). SRCMOD: An Online Database of Finite-Fault

795 Rupture Models. Seismological Research Letters, 85(6), 1348–1357.

796 https://doi.org/10.1785/0220140077

797 Malagnini, L., Nielsen, S., Mayeda, K., & Boschi, E. (2010). Energy radiation from

32 798 intermediate- to large-magnitude earthquakes: Implications for dynamic fault

799 weakening. Journal of Geophysical Research: Solid Earth, 115, 1–30.

800 https://doi.org/10.1029/2009JB006786

801 Manighetti, I., King, G. C. P., Gaudemer, Y., Scholz, C. H., & Doubre, C. (2001). Slip

802 accumulation and lateral propagation of active normal faults in Afar. Journal of

803 Geophysical Research, 106(B7), 13667. https://doi.org/10.1029/2000JB900471

804 Manighetti, I., Campillo, M., Bouley, S., & Cotton, F. (2007). Earthquake scaling, fault

805 segmentation, and structural maturity. Earth and Planetary Science Letters, 253(3–

806 4), 429–438. https://doi.org/10.1016/j.epsl.2006.11.004

807 Manighetti, I., Caulet, C., De Barros, D., Perrin, C., Cappa, F., & Gaudemer, Y. (2015).

808 Generic along-strike segmentation of Afar normal faults, East Africa: Implications

809 on fault growth and stress heterogeneity on seismogenic fault planes. Geochem.

810 Geophys. Geosyst., 16, 443–467. https://doi.org/10.1002/2014GC005691.Received

811 Marone, C., Vidale, J. E., & Ellsworth, W. L. (1995). Fault healing inferred from time

812 dependent variations in source properties of repeating earthquakes. Geophysical

813 Research Letters, 22(22), 3095–3098. https://doi.org/10.1029/95GL03076

814 Matthews, V. (1976). Correlation of Pinnacles and Neenach volcanic formations and

815 their bearing on San Andreas fault problem. The American Association of Petroleum

816 Geologists Bulletin, 60, 2128–2141.

817 Mendoza, C., & Hartzell, S. H. (1988). Aftershock patterns and main shock faulting.

818 Bulletin of the Seismological Society of America, 78(4), 1438–1449.

819 Mikumo, T., & Miyatake, T. (1995). Heterogeneous distribution of dynamic stress drop

820 and relative fault strength recovered from the results of waveform inversion: the

821 1984 Morgan Hill, California, earthquake. Bulletin of the Seismological Society

822 America, 85(1), 178–193.

33 823 Niemeijer, A. R., Marone, C., & Elsworth, D. (2010). Fabric induced weakness of tectonic

824 faults. Geophysical Research Letters, 37(3), n/a-n/a.

825 https://doi.org/10.1029/2009GL041689

826 Oskin, M., Perg, L., Blumentritt, D., Mukhopadhyay, S., & Iriondo, A. (2007). Slip rate of

827 the Calico fault: Implications for geologic versus geodetic rate discrepancy in the

828 Eastern California Shear Zone. Journal of Geophysical Research: Solid Earth, 112(3),

829 1–16. https://doi.org/10.1029/2006JB004451

830 Ostermeijer, G. A., Mitchell, T. M., Aben, F. M., Dorsey, M. T., Browning, J., Rockwell, T. K.,

831 et al. (2020). Damage zone heterogeneity on seismogenic faults in crystalline rock; a

832 field study of the Borrego Fault, . Journal of Structural Geology,

833 104016. https://doi.org/10.1016/j.jsg.2020.104016

834 Perrin, C., Manighetti, I., Ampuero, J.-P., Cappa, F., & Gaudemer, Y. (2016). Location of

835 largest earthquake slip and fast rupture controlled by along-strike change in fault

836 structural maturity due to fault growth. Journal of Geophysical Research: Solid Earth,

837 121(5), 3666–3685. https://doi.org/10.1002/2015JB012671

838 Perrin, C., Manighetti, I., & Gaudemer, Y. (2016). Off-fault tip splay networks: A genetic

839 and generic property of faults indicative of their long-term propagation. Comptes

840 Rendus Geoscience, 348(1), 52–60. https://doi.org/10.1016/j.crte.2015.05.002

841 Perrin, C., Waldhauser, F., Choi, E., & Scholz, C. H. (2019). Persistent fine-scale fault

842 structure and rupture development: A new twist in the Parkfield, California, story.

843 Earth and Planetary Science Letters, 521, 128–138.

844 https://doi.org/10.1016/j.epsl.2019.06.010

845 Peyrat, S., Olsen, K., & Madariaga, R. (2001). Dynamic modeling of the 1992 Landers

846 earthquake. Journal of Geophysical Research: Solid Earth, 106(B11), 26467–26482.

847 https://doi.org/10.1029/2001JB000205

34 848 Powell, R. E., & Weldon, R. J. (1992). Evolution of the San Andreas Fault. Annual Review of

849 Earth and Planetary Sciences, 20(1), 431–468.

850 https://doi.org/10.1146/annurev.ea.20.050192.002243

851 Powers, P. M., & Jordan, T. H. (2010). Distribution of seismicity across strike-slip faults in

852 California. Journal of Geophysical Research: Solid Earth, 115(5).

853 https://doi.org/10.1029/2008JB006234

854 Price, E. J., & Burgmann, R. (2002). Interactions between the Landers and Hector Mine,

855 California, Earthquakes from Space Geodesy, Boundary Element Modeling, and

856 Time-Dependent Friction. Bulletin of the Seismological Society of America, 92(4),

857 1450–1469. https://doi.org/10.1785/0120000924

858 Radiguet, M., Cotton, F., Manighetti, I., Campillo, M., & Douglas, J. (2009). Dependency of

859 Near-Field Ground Motions on the Structural Maturity of the Ruptured Faults.

860 Bulletin of the Seismological Society of America, 99(4), 2572–2581.

861 https://doi.org/10.1785/0120080340

862 Revenaugh, J., & Reasoner, C. (1997). Cumulative offset of the San Andreas fault in

863 central California: A seismic approach. Geology, 25(2), 123–126.

864 https://doi.org/10.1130/0091-7613(1997)025<0123:COOTSA>2.3.CO;2

865 Rockwell, T. K., Lindvall, S., Herzberg, M., Murbach, D., Dawson, T., & Berger, G. (2000).

866 Paleoseismology of the Johnson Valley, Kickapoo, and Homestead Valley faults:

867 Clustering of earthquakes in the Eastern California shear zone. Bulletin of the

868 Seismological Society of America, 90(5), 1200–1236.

869 https://doi.org/10.1785/0119990023

870 Ross, Z. E., Kanamori, H., Hauksson, E., & Aso, N. (2018). Dissipative Intraplate Faulting

871 During the 2016 Mw 6.2 Tottori, Japan Earthquake. Journal of Geophysical Research:

872 Solid Earth, 123(2), 1631–1642. https://doi.org/10.1002/2017JB015077

35 873 Rubin, C. M., & Sieh, K. (1997). Long dormancy, low slip rate, and similar slip-per-event

874 for the Emerson fault, eastern California shear zone. Journal of Geophysical

875 Research, 102(B7), 15319. https://doi.org/10.1029/97JB00265

876 Savage, H. M., & Brodsky, E. E. (2011). Collateral damage: Evolution with displacement of

877 fracture distribution and secondary fault strands in fault damage zones. Journal of

878 Geophysical Research, 116(B3), B03405. https://doi.org/10.1029/2010JB007665

879 Schaff, D. P., Bokelmann, G. H. R., Beroza, G. C., Waldhauser, F., & Ellsworth, W. L. (2002).

880 High-resolution image of Calaveras Fault seismicity. Journal of Geophysical

881 Research: Solid Earth, 107(B9), ESE 5-1-ESE 5-16.

882 https://doi.org/10.1029/2001JB000633

883 Scholz, C. H. (1987). Wear and gouge formation in brittle faulting. Geology, 15(6), 493.

884 https://doi.org/10.1130/0091-7613(1987)15<493:WAGFIB>2.0.CO;2

885 Scholz, C. H. (1998). A further note on earthquake size distributions. Bulletin of the

886 Seismological Society of America, 88(5), 1325–1326.

887 Scholz, C. H. (2019). The mechanics of earthquakes and faulting (Third edit). Cambridge

888 university press.

889 Scholz, C. H., & Lawler, T. M. (2004). Slip tapers at the tips of faults and earthquake

890 ruptures. Geophysical Research Letters, 31(21), L21609.

891 https://doi.org/10.1029/2004GL021030

892 Scholz, C. H., Aviles, C. A., & Wesnousky, S. G. (1986). Scaling differences between large

893 interplate and intraplate earthquakes. Bulletin of the Seismological Society of

894 America, 76(1), 65–70.

895 Scott, C. P., Arrowsmith, J. R., Nissen, E., Lajoie, L., Maruyama, T., & Chiba, T. (2018). The

896 M 7 2016 Kumamoto, Japan, Earthquake: 3-D Deformation Along the Fault and

897 Within the Damage Zone Constrained From Differential Lidar Topography. Journal

36 898 of Geophysical Research: Solid Earth, 123(7), 6138–6155.

899 https://doi.org/10.1029/2018JB015581

900 Segall, P., & Pollard, D. D. (1980). Mechanics of discontinuous faults. Journal of

901 Geophysical Research, 85(B8), 4337. https://doi.org/10.1029/JB085iB08p04337

902 Sharp, R. V. (1967). San Jacinto Fault Zone in the of Southern

903 California. Geological Society of America Bulletin, 78, 705–730.

904 Sims, J. (1993). Chronology of displacement on the San Andreas fault in central

905 California : Evidence from reversed positions of exotic rock bodies near Parkfield,

906 California. In R. E. Powell, R. J. Weldon, & J. C. Matti (Eds.), The San Andreas Fault

907 System : Displacement, Palinspastic Reconstruction and Geologic Evolution. Boulder,

908 Colorado: Geological Society of America Memoir 178.

909 Spudich, P., & Olsen, K. B. (2001). Fault zone amplified waves as a possible seismic

910 hazard along the Calaveras Fault in central California. Geophysical Research Letters,

911 28(13), 2533–2536. https://doi.org/10.1029/2000GL011902

912 Stirling, M. W., Wesnousky, S. G., & Shimazaki, K. (1996). Fault trace complexity,

913 cumulative slip, and the shape of the magnitude-frequency distribution for strike-

914 slip faults: a global survey. Geophysical Journal International, 124(3), 833–868.

915 https://doi.org/10.1111/j.1365-246X.1996.tb05641.x

916 Toké, N. A., Arrowsmith, J. R., Rymer, M. J., Landgraf, A., Haddad, D. E., Busch, M., et al.

917 (2011). Late slip rate of the San Andreas fault and its accommodation by

918 creep and moderate-magnitude earthquakes at Parkfield, California. Geology, 39(3),

919 243–246. https://doi.org/10.1130/G31498.1

920 Twardzik, C., Das, S., & Madariaga, R. (2014). Inversion for the physical parameters that

921 control the source dynamics of the 2004 Parkfield earthquake. Journal of

922 Geophysical Research: Solid Earth, 119(9), 7010–7027.

37 923 https://doi.org/10.1002/2014JB011238

924 Unruh, J. R., Kelson, K., & Barron, A. (2002). Critical Evalutation of the Northern

925 Termination of the Calaveras Fault, Eastern San Francisco Bay Area, California. Final

926 Technical Report submitted to the US Geological Survey, Award no. 1434-HQ- 97-GR-

927 03146.

928 Valoroso, L., Chiaraluce, L., & Collettini, C. (2014). Earthquakes and fault zone structure.

929 Geology, 42(4), 343–346. https://doi.org/10.1130/G35071.1

930 Vermilye, J. M., & Scholz, C. H. (1998). The process zone: A microstructural view of fault

931 growth. Journal of Geophysical Research, 103(B6), 12223.

932 https://doi.org/10.1029/98JB00957

933 Vidale, J. E., Ellsworth, W. L., Cole, A., & Marone, C. (1994). Variations in Rupture Process

934 With Recurrence Interval in a Repeated Small Earthquake. Nature, 368(6472), 624–

935 626.

936 Wakabayashi, J. (1999). Distribution of displacement on and evolution of a young system

937 , California Sierra SN :, 18(6), 1245–1274.

938 Wald, D. J., Helmberger, D. V., & Hartzell, S. H. (1990). Rupture Process of the 1987

939 Superstition Hills Earthquake from the Inversion of Strong-Motion Data. Bulletin of

940 the Seismological Society America, 80(5), 1079–1098.

941 Waldhauser, F. (2009). Near-Real-Time Double-Difference Event Location Using Long-

942 Term Seismic Archives, with Application to Northern California. Bulletin of the

943 Seismological Society of America, 99(5), 2736–2748.

944 https://doi.org/10.1785/0120080294

945 Waldhauser, F., & Ellsworth, W. L. (2000). A Double-difference Earthquake location

946 algorithm: Method and application to the Northern Hayward Fault, California.

947 Bulletin of the Seismological Society of America, 90(6), 1353–1368.

38 948 https://doi.org/10.1785/0120000006

949 Waldhauser, F., & Schaff, D. P. (2008). Large-scale relocation of two decades of Northern

950 California seismicity using cross-correlation and double-difference methods.

951 Journal of Geophysical Research: Solid Earth, 113(8), 1–15.

952 https://doi.org/10.1029/2007JB005479

953 Waldhauser, F., Ellsworth, W. L., Schaff, D. P., & Cole, A. (2004). Streaks, multiplets, and

954 holes: High-resolution spatio-temporal behavior of Parkfield seismicity. Geophysical

955 Research Letters, 31(18), L18608. https://doi.org/10.1029/2004GL020649

956 Wallace, R. E. (1989). Segmentation in fault zones. In Proc. USGS Conference of

957 Segmentation of Faults, U. S. Geol. Surv. Open-file Rept. 89-315, edited.

958 Wesling, J. R., & Hanson, K. L. (2008). Mapping of the West Napa Fault Zone for Input into

959 the Northern California Quaternary Fault Database. U.S. Geological Survey, Open File

960 Report (Vol. 05HQAG0002).

961 Wesnousky, S. G. (1988). Seismological and structural evolution of strike-slip faults.

962 Nature. https://doi.org/10.1038/335340a0

963 Wetzler, N., Brodsky, E. E., & Lay, T. (2016). Regional and stress drop effects on

964 aftershock productivity of large megathrust earthquakes. Geophysical Research

965 Letters, 43(23), 12,012-12,020. https://doi.org/10.1002/2016GL071104

966 Wilson, J. ., Chester, J. ., & Chester, F. . (2003). Microfracture analysis of fault growth and

967 wear processes, Punchbowl Fault, San Andreas system, California. Journal of

968 Structural Geology, 25(11), 1855–1873. https://doi.org/10.1016/S0191-

969 8141(03)00036-1

970 Xu, W., Wu, S., Materna, K., Nadeau, R., Floyd, M., Funning, G., et al. (2018). Interseismic

971 Ground Deformation and Fault Slip Rates in the Greater San Francisco Bay Area

972 From Two Decades of Space Geodetic Data. Journal of Geophysical Research: Solid

39 973 Earth, 123(9), 8095–8109. https://doi.org/10.1029/2018JB016004

974 Yamanaka, Y., & Shimazaki, K. (1990). Scaling relationship between the number of

975 aftershocks and the size of the main shock. Journal of Physics of the Earth, 38(4),

976 305–324. https://doi.org/10.4294/jpe1952.38.305

977 Yang, H. (2015). Recent advances in imaging crustal fault zones: a review. Earthquake

978 Science, 28(2), 151–162. https://doi.org/10.1007/s11589-015-0114-3

979 Yoshioka, N., & Scholz, C. H. (1989). Elastic properties of contracting surfaces under

980 normal and shear loads 1. Theory. Journal of Geophysical Research, 94(B12).

981 https://doi.org/10.1029/jb094ib12p17681

982 Zeng, Y., & Shen, Z. (2016). A Fault‐Based Model for Crustal Deformation, Fault Slip

983 Rates, and Off‐Fault Strain Rate in California. Bulletin of the Seismological Society

984 of America, 106(2), 766–784. https://doi.org/10.1785/0120140250

985

986

987

988 SUPPORTING INFORMATION REFERENCES

989 Allam, A. A., Ben-Zion, Y., Kurzon, I., & Vernon, F. (2014). Seismic velocity structure in the

990 Hot Springs and Trifurcation areas of the San Jacinto fault zone, California, from

991 double-difference tomography. Geophysical Journal International, 198(2), 978–999.

992 https://doi.org/10.1093/gji/ggu176

993 Beroza, G. C., & Spudich, P. (1988). Linearized inversion for fault rupture behavior:

994 Application to the 1984 Morgan Hill, California, earthquake. Journal of Geophysical

995 Research, 93(B6), 6275. https://doi.org/10.1029/JB093iB06p06275

996 Blümling, P., Mooney, W. D., & Lee, W. H. K. (1985). Crustal structure of the southern

997 Calaveras fault zone, central California, from seismic refraction investigations.

40 998 Bulletin of the Seismological Society of America, 75(1), 193–209.

999 Catchings, R. D., Rymer, M. J., Goldman, M. R., Hole, J. A., Huggins, R., & Lippus, C. (2002).

1000 High-Resolution Seismic Velocities and Shallow Structure of the San Andreas Fault

1001 Zone at Middle Mountain, Parkfield, California. Bulletin of the Seismological Society

1002 of America, 92(6), 2493–2503. https://doi.org/10.1785/0120010263

1003 Cochran, E. S., Li, Y. G., Shearer, P. M., Barbot, S., Fialko, Y., & Vidale, J. E. (2009). Seismic

1004 and geodetic evidence for extensive, long-lived fault damage zones. Geology, 37(4),

1005 315–318. https://doi.org/10.1130/G25306A.1

1006 Cohee, B. P., & Beroza, G. C. (1994). Slip distribution of the and

1007 its implications for earthquake source mechanics. Bulletin of the Seismological

1008 Society America, 84(3), 692–712.

1009 Cotton, F., & Campillo, M. (1995). Frequency domain inversion of strong motions:

1010 Application to the 1992 Landers earthquake. Journal of Geophysical Research: Solid

1011 Earth, 100(B3), 3961–3975. https://doi.org/10.1029/94JB02121

1012 Custódio, S., Liu, P., & Archuleta, R. J. (2005). The 2004 Mw6.0 Parkfield, California,

1013 earthquake: Inversion of near-source ground motion using multiple data sets.

1014 Geophysical Research Letters, 32(23), 1–4.

1015 https://doi.org/10.1029/2005GL024417

1016 Dreger, D. S., Gee, L., Lombard, P., Murray, M. H., & Romanowicz, B. (2005). Rapid Finite-

1017 source Analysis and Near-fault Strong Ground Motions: Application to the 2003 Mw

1018 6.5 San Simeon and 2004 Mw6.0 Parkfield Earthquakes. Seismological Research

1019 Letters, 76(1), 40–48. https://doi.org/10.1785/gssrl.76.1.40

1020 Fialko, Y. (2004). Probing the mechanical properties of seismically active crust with

1021 space geodesy: Study of the coseismic deformation due to the 1992 M w 7.3

1022 Landers (southern California) earthquake. Journal of Geophysical Research,

41 1023 109(B3), B03307. https://doi.org/10.1029/2003JB002756

1024 Fialko, Y., Sandwell, D., Agnew, D., Simons, M., Shearer, P. M., & Minster, B. (2002).

1025 Deformation on Nearby Faults Induced by the 1999 Hector Mine Earthquake.

1026 Science, 297(5588), 1858–1862. https://doi.org/10.1126/science.1074671

1027 Gallovič, F. (2016). Modeling Velocity Recordings of the M w 6.0 South Napa, California,

1028 Earthquake: Unilateral Event with Weak High ‐ Frequency Directivity.

1029 Seismological Research Letters, 87(1), 2–14. https://doi.org/10.1785/0220150042

1030 Hartzell, S. H., & Heaton, T. H. (1986). Rupture history of the 1984 Morgan Hill,

1031 California, earthquake from the inversion of strong motion records. Bulletin -

1032 Seismological Society of America, 76(3), 649-674.

1033 Hauksson, E., Yang, W., & Shearer, P. M. (2012). Waveform relocated earthquake catalog

1034 for Southern California (1981 to June 2011). Bulletin of the Seismological Society of

1035 America, 102(5), 2239–2244. https://doi.org/10.1785/0120120010

1036 Hernandez, B., Cotton, F., & Campillo, M. (1999). Contribution of radar interferometry to

1037 a two-step inversion of the kinematic process of the 1992 Landers earthquake.

1038 Journal of Geophysical Research: Solid Earth, 104(B6), 13083–13099.

1039 https://doi.org/10.1029/1999JB900078

1040 Ji, C. (2004). Slip history the 2004 (Mw 5.9) Parkfield Earthquake (Single-Plane Model).

1041 Retrieved from

1042 http://www.tectonics.caltech.edu/slip_history/2004_ca/parkfield2.html

1043 Ji, C., Wald, D. J., & Helmberger, D. V. (2002). Source Description of the 1999 Hector Mine,

1044 California, Earthquake, Part II: Complexity of Slip History. Bulletin of the

1045 Seismological Society of America, 92(4), 1208–1226.

1046 https://doi.org/10.1785/0120000917

1047 Jonsson, S., Zebker, H., Segall, P., & Amelung, F. (2002). Fault Slip Distribution of the

42 1048 1999 Mw 7.1 Hector Mine, California, Earthquake, Estimated from Satellite Radar

1049 and GPS Measurements. Bulletin of the Seismological Society of America, 92(4),

1050 1377–1389. https://doi.org/10.1785/0120000922

1051 Kaverina, A., Dreger, D. S., & Price, E. J. (2002). The Combined Inversion of Seismic and

1052 Geodetic Data for the Source Process of the 16 October 1999 Mw 7.1 Hector Mine,

1053 California, Earthquake. Bulletin of the Seismological Society of America, 92(4),

1054 1266–1280. https://doi.org/10.1785/0120000907

1055 Larsen, S., Reilinger, R., Neugebauer, H., & Strange, W. (1992). Global positioning system

1056 measurements of deformations associated with the 1987 Superstition Hills

1057 Earthquake: Evidence for conjugate faulting. Journal of Geophysical Research,

1058 97(B4), 4885. https://doi.org/10.1029/92JB00128

1059 Lewis, M. A., & Ben-Zion, Y. (2010). Diversity of fault zone damage and trapping

1060 structures in the Parkfield section of the San Andreas Fault from comprehensive

1061 analysis of near fault seismograms. Geophysical Journal International, 183(3),

1062 1579–1595. https://doi.org/10.1111/j.1365-246X.2010.04816.x

1063 Lewis, M. A., Peng, Z., Ben-Zion, Y., & Vernon, F. L. (2005). Shallow seismic trapping

1064 structure in the San Jacinto fault zone near Anza, California. Geophysical Journal

1065 International, 162(3), 867–881. https://doi.org/10.1111/j.1365-

1066 246X.2005.02684.x

1067 Li, Y. G., & Vernon, F. L. (2001). Characterization of the San Jacinto fault zone near Anza,

1068 California, by fault zone trapped waves. Journal of Geophysical Research: Solid

1069 Earth, 106(B12), 30671–30688. https://doi.org/10.1029/2000JB000107

1070 Li, Y. G., Aki, K., Vidale, J. E., Lee, W. H. K., & Marone, C. J. (1994). Fine Structure of the

1071 Landers Fault Zone: Segmentation and the Rupture Process. Science, 265(5170),

1072 367–370. https://doi.org/10.1126/science.265.5170.367

43 1073 Li, Y. G., Aki, K., & Vernon, F. L. (1997). San Jacinto Fault Zone guided waves: A

1074 discrimination for recently active fault strands near Anza, California. Journal of

1075 Geophysical Research: Solid Earth, 102(B6), 11689–11701.

1076 https://doi.org/10.1029/97JB01050

1077 Li, Y. G., Aki, K., Vidale, J. E., & Xu, F. (1999). Shallow structure of the Landers Fault Zone

1078 from explosion-generated trapped waves. Journal of Geophysical Research: Solid

1079 Earth, 104(B9), 20257–20275. https://doi.org/10.1029/1999JB900194

1080 Li, Y. G., Vidale, J. E., Aki, K., & Xu, F. (2000). Depth-dependent structure of the Landers

1081 fault zone from trapped waves generated by aftershocks. Journal of Geophysical

1082 Research: Solid Earth, 105(B3), 6237–6254.

1083 https://doi.org/10.1029/1999JB900449

1084 Li, Y. G., Vidale, J. E., Oglesby, D. D., Day, S. M., & Cochran, E. (2003). Multiple-fault

1085 rupture of the M 7.1 Hector Mine, California, earthquake from fault zone trapped

1086 waves. Journal of Geophysical Research: Solid Earth, 108(B3), 1–25.

1087 https://doi.org/10.1029/2001JB001456

1088 Li, Y. G., Vidale, J. E., & Cochran, E. S. (2004). Low-velocity damaged structure of the San

1089 Andreas Fault at Parkfield from fault zone trapped waves. Geophysical Research

1090 Letters, 31(12), n/a-n/a. https://doi.org/10.1029/2003GL019044

1091 Mai, P. M., & Thingbaijam, K. K. S. (2014). SRCMOD: An Online Database of Finite-Fault

1092 Rupture Models. Seismological Research Letters, 85(6), 1348–1357.

1093 https://doi.org/10.1785/0220140077

1094 Mendoza, C., & Hartzell, S. (2013). Finite-Fault Source Inversion Using Teleseismic P

1095 Waves: Simple Parameterization and Rapid Analysis. Bulletin of the Seismological

1096 Society of America, 103(2A), 834–844. https://doi.org/10.1785/0120120069

1097 Mooney, W. D., & Colburn, R. H. (1985). Seismic-Refraction Profile Across the San

44 1098 Andreas, Sargent, and Calaveras Faults, West-Central California. Bulletin of the

1099 Seismological Society of America, 75(1), 175–191.

1100 Peng, Z., Ben-Zion, Y., Michael, A. J., & Zhu, L. (2003). Quantitative analysis of seismic

1101 fault zone waves in the rupture zone of the 1992 Landers, California, earthquake:

1102 evidence for a shallow trapping structure. Geophysical Journal International,

1103 155(3), 1021–1041. https://doi.org/10.1111/j.1365-246X.2003.02109.x

1104 Perrin, C., Waldhauser, F., Choi, E., & Scholz, C. H. (2019). Persistent fine-scale fault

1105 structure and rupture development: A new twist in the Parkfield, California, story.

1106 Earth and Planetary Science Letters, 521, 128–138.

1107 https://doi.org/10.1016/j.epsl.2019.06.010

1108 Powers, P. M., & Jordan, T. H. (2010). Distribution of seismicity across strike-slip faults in

1109 California. Journal of Geophysical Research: Solid Earth, 115(5).

1110 https://doi.org/10.1029/2008JB006234

1111 Salichon, J., Lundgren, P., Delouis, B., & Giardini, D. (2004). Slip History of the 16 October

1112 1999 Mw 7.1 Hector Mine Earthquake (California) from the Inversion of InSAR,

1113 GPS, and Teleseismic Data. Bulletin of the Seismological Society of America, 94(6),

1114 2015–2027. https://doi.org/10.1785/0120030038

1115 Spudich, P., & Olsen, K. B. (2001). Fault zone amplified waves as a possible seismic

1116 hazard along the Calaveras Fault in central California. Geophysical Research Letters,

1117 28(13), 2533–2536. https://doi.org/10.1029/2000GL011902

1118 Wald, D. J., & Heaton, T. H. (1994). Spatial and temporal distribution of slip for the 1992

1119 Landers, California, earthquake. Bulletin of the Seismological Society of America,

1120 84(3), 668-691.

1121 Wald, D. J., Helmberger, D. V., & Hartzell, S. H. (1990). Rupture Process of the 1987

1122 Superstition Hills Earthquake from the Inversion of Strong-Motion Data. Bulletin of

45 1123 the Seismological Society America, 80(5), 1079–1098.

1124 Waldhauser, F., & Ellsworth, W. L. (2000). A Double-difference Earthquake location

1125 algorithm: Method and application to the Northern Hayward Fault, California.

1126 Bulletin of the Seismological Society of America, 90(6), 1353–1368.

1127 https://doi.org/10.1785/0120000006

1128 Waldhauser, F., & Schaff, D. P. (2008). Large-scale relocation of two decades of Northern

1129 California seismicity using cross-correlation and double-difference methods.

1130 Journal of Geophysical Research: Solid Earth, 113(8), 1–15.

1131 https://doi.org/10.1029/2007JB005479

1132 Wei, S., Fielding, E., Leprince, S., Sladen, A., Avouac, J.-P., Helmberger, D., et al. (2011).

1133 Superficial simplicity of the 2010 El Mayor–Cucapah earthquake of Baja California

1134 in Mexico. Nature Geoscience, 4(9), 615–618. https://doi.org/10.1038/ngeo1213

1135

1136

46 1137 Figures and figure captions:

1138

1139 Figure 1: Simplified view of the architecture of a fault zone and the density of

1140 fractures (red) and seismicity (blue) away from the fault core.

1141

47 1142

1143

48 1144 Figure 2: Aftershocks distribution of the (a, b) 2004 Parkfield and (c, d, e, f, g)

1145 1999 Hector Mine earthquakes (see Supp. Fig. S3 for all cases). (a) Map view showing

1146 the San Andreas fault (black lines) and the surface rupture (thick grey line) of the 2004

1147 Parkfield earthquake (epicenter indicated by the red star). Red dots are aftershocks that

1148 occurred within 2 months after the mainshock (Waldhauser & Schaff, 2008; Perrin et al.,

1149 2019). (b) Gray profiles are fault-normal earthquake distributions measured from the

1150 best fitting plane in each moving box along the rupture trace (see text for explanations).

1151 Black curve is the mean of the gray profiles. Inset: cross section going through the

1152 hypocenter area. Depth in y-axis; across strike distance in x-axis. Black line is best fitting

1153 plane minimizing fault-normal distance to aftershock hypocenters (red dots); (c) same

1154 as (a) but for the 1999 Hector Mine earthquake (earthquake catalog from Hauksson et

1155 al., 2012). Boxes include earthquakes used in d-g. (d-g) Same as (b) but for the 1999

1156 Hector Mine earthquake. The four sub-figures are based on earthquakes included in

1157 boxes shown in (c).

1158

49 1159

1160 Figure 3: Determination of WS1 and WS2 parameters from the aftershocks

1161 distribution of the (a) Parkfield 2004 and (b, c, d, e) Hector Mine 1999

1162 earthquakes (see Supp. Fig. S4 for all cases). Blue circles represent the mean

1163 hypocenter distances from each fault section within 0.03 km wide bins (see black curve

1164 in Fig. 2 and Supp. Fig. S3). Grey circles represent the mean distances of events that are

1165 only included in less than 20% of all boxes. The red curve is the best (power law) fit to

1166 the blue circles, taking into account the power law singularity on the fault (see Supp. Fig.

1167 S4 for details). The vertical gray dashed lines labeled WS1 and WS2 point out the locations

1168 where the power law fit shows the maximum curvature and where the data (blue

50 1169 circles) deviate from the power law fit by dropping or flattening (i.e., the transition

1170 between seismically active fault damage zone and the undeformed crustal medium),

1171 respectively.

51 Half width of Earthquake Half width of Long-term the shear Name of fault Initiation Cumulative References Location name, date, seismicity fall slip rate deformation section(s) age (Ma) slip (km) for fault parameters magnitude off WS1 (km) (mm/yr) zone WS2 (km) El Mayor Elsinore (southern Dorsey et al., 2012; K. E. K. Fletcher USA Cucapah, 0.6  0.03 4.51  0.03 ~ 1.1 1 to 2 1 to 2 section) et al., 2011 and references therein 2010, Mw 7.2 Dibblee, 1961; Dokka, 1983; Dokka 0.86 Hector Mine, & Travis, 1990; Garfunkel, 1974; USA +0.26/-0.13 3.36  0.03 Lavic Lake-Bullion < 10 10 to 20 ~ 0.8 1999, M 7.1 Jachens et al., 2002; Oskin et al., w (mean value) 2007 Dibblee, 1961; Dokka, 1983; Dokka 0.96 Emerson-Camp Rock- Landers, & Travis, 1990; Garfunkel, 1974; USA +0.16/-0.27 3.96  0.03 Homestead Valley- < 10 3.5 to 4.6 0.2 to 0.7 1992, M 7.3 Jachens et al., 2002; Rockwell et w (mean value) Johnson Valley al., 2000; Rubin & Sieh, 1997 Morgan Hill, Stirling et al., 1996; Wakabayashi, USA 0.12  0.03 1.32  0.03 Calaveras ~ 12 60 to 70 3 to 25 1984, Mw 6.1 1999 and references therein Atwater & Stock, 1998; Critelli & Nilsen, 2000; Crowell, 1979; Parkfield, San Andreas (central Graham et al., 1989; Liu et al., USA 0.06  0.03 0.56  0.075 24 to 29 ~ 315 ~ 26 2004, Mw 6.0 section) 2010; Matthews, 1976; Revenaugh & Reasoner, 1997; Toké et al., 2011 West Napa (considered d’Alessio, 2005; Field et al., 2015; South Napa, USA 0.22  0.03 0.90 0.06 as a splay of the 6 to 12† 3.8 to 28† 1 to 7 Wakabayashi, 1999; Wesling & 2014, M 6.1 w Calaveras fault zone) Hanson, 2008; Zeng & Shen, 2016 Blisniuk et al., 2010; Dorsey et al., Superstition San Jacinto (southern 2012; Gurrola & Rockwell, 1996; USA Hills, 1987, 0.78  0.03 > 3 < 2 ~ 4 ~ 4 section) Hudnut & Sieh, 1989; Kirby et al., M 6.6 w 2007; Lutz et al., 2006 0.12 to 1.83 ~2.80 New - (field (field Awatere < 4 < 2 ~ 5 Little, 1995 and references therein Zealand measurements) measurements)

52 1172

1173 Table 1: Fault and aftershock distribution parameters for the seven earthquake sequences analyzed in this study. Uncertainties in WS1

1174 and WS2 correspond to size of bin (0.03 km) used in this analysis. For multiple broken fault sections (Hector Mine, Landers, El Mayor Cucapah),

1175 WS1 and WS2 are mean values, when possible, and the uncertainties represent the minimum and maximum range of values (see detailed

1176 measurements in Supp. Figure S3). Field measurements of the Awatere fault from Little et al. (1995) are also indicated (see text for details).

1177 * available at: https://gbank.gsj.jp/activefault/

1178 † cumulative slip deduced from scaling relations in Scholz & Lawler, 2004 (West Napa fault). Initiation age of the West Napa fault is estimated

1179 from

1180 the age of surrounding faults in California. See text for details

53 1181

1182 Figure 4: Relations between WS1 and WS2 for the seven earthquakes analyzed in

1183 this study. Grey lines indicate two affine functions between WS1 and WS2 that bound our

1184 data. Uncertainties are reported in Table 1.

1185

54 1186

1187 Figure 5: Relations between WS1 (circles; a, c, e), WS2 (squares; b, d, f) of seven

1188 earthquake sequences and the cumulative slip (a, b), initiation age (c, d) and

1189 geological slip rate (e, f) of their host fault taken from literature (solid symbols)

55 1190 and estimated from scaling relations (open symbols for fault parameters

1191 associated with the South Napa earthquakes) (see Table 1 and text for details).

1192 Power laws are indicated by grey lines. For comparison, red symbols indicate geological

1193 surface measurements along the Awatere fault (from Little, 1995). D is cumulative fault

1194 displacement, A is initiation age of the fault, SR is slip rate, k is a different constant in

1195 each relation.

1196

56 1197

1198 Figure 6: (a) Normalized distribution of the mean number of aftershocks as a

1199 function of normal distance from the fault (see also black curves in Fig. 2 and Supp.

1200 Fig. S3). Each colored curve represents one earthquake sequence. Curves with the same

1201 color are distinct fault sections that broke during one earthquake. Colors are warmer as

1202 the structural maturity of the broken fault sections increases. For comparison, black

1203 squares indicate the cumulative number of faults (right y axis) measured at surface from

1204 the Awatere fault (modified from Little, 1995). (b) Sketch summarizing the fault-normal

57 1205 distributions of aftershocks for immature (blue curve) and mature (red curve) fault

1206 sections. (c) Interpretative cross-section describing the structural makeup of immature

1207 (blue) and mature (red) faults. As fault structural maturity increases, inner and outer

1208 bounds of the shear deformation zone (WS1 and WS2, respectively) decrease, as

1209 expressed by a decreasing width of the fault-normal aftershock distribution. See text for

1210 more discussion.

1211

58 1212

1213 Figure 7: Comparison between the outer bound of the shear deformation zone (WS2;

1214 data from fig. 5b) and the predictions of the model of Aslam & Daub, 2018 that were

1215 calibrated to fault displacement with roughness-displacement relation of Stirling et al.,

1216 1996 (blue circles, dashed lines and shaded area). See text for explanation of the latter.

1217 Blue circles are for step widths of 1 and 2 km, which are typical step width range

1218 (Wesnousky, 1988).

1219

1220

59 1221

Earthquake Number of Mean Mean Mean References name measurements log10() standard stress drop error (in MPa) Hector Mine, 8 0.672 0.011 4.7 ± 1 Allmann & Shearer, 2009; 1999 Ancheta et al., 2014; Baltay et al., 2019; Hanks & Bakun, 2008; Hauksson et al., 2002; Kaverina et al., 2002; Price & Burgmann, 2002 Landers, 6 0.792 0.016 6.2 ± 1 Allmann & Shearer, 2009; 1992 Ancheta et al., 2014; Baltay et al., 2019; J. B. Fletcher & McGarr, 2006; Peyrat et al., 2001; Price & Burgmann, 2002 Morgan Hill, 3 0.447 0.032 2.8 ± 1 Mikumo & Miyatake, 1995 1984 Parkfield, 4 0.362 0.011 2.3 ± 1 Allmann & Shearer, 2009; 2004 Kim & Dreger, 2008; Ma et al., 2008; Twardzik et al., 2014 Superstition 3 0.763 0.011 5.8 ± 1 Ancheta et al., 2014; Hills, 1987 Baltay et al., 2019; Wald et al., 1990 1222 1223 Table 2: Mean stress drop values available from the literature for five of the seven

1224 earthquakes analyzed in this study.

1225

1226

60 1227

1228 Figure 8: Earthquake stress drop (Δ휎) as a function of cumulative displacement of the

1229 broken fault section (D) for earthquakes listed in Table 2.

1230

1231

1232

61 1233

1234 1235 Figure 9: Earthquake stress drop (Δ휎) as a function of geological slip rate (SR) for

1236 earthquakes listed in Table 2, represented in (a) log-log and (b) semi-log coordinates.

1237

62 Figure 1. log fracture and seismicity density

WS1

3 2 1

log 1 cataclasite core Wc distance W from fault 2 dilatant damage zone d W 3 shear deformation zone S2 Figure 2. N A’ 2 a 0 −2 2004 Across strike Across distance [km] distance A −20 −15 −10 −5 0 5 10 15 20 Along strike distance [km] 10 0 A A’ b 0 10 −1

10 −2 −10 [km]

10 −3 −20 −5 0 5 −4 10 [km]

Number of events (normalized) Number of events −5 10 1 10 −2 10 −1 10 0 10 Normal distance from the best fitting plane [km]

d D’ g c 5 E’ 1999 G’ f F’ 0 D e N Across strike Across E distance [km] distance -5 F G

-30 -20 -10 0 10 20 30 10 -1 10 -1 d D D’ e E E’ 0 0 10 -2 -10 -10

-2 [km]

[km] 10 -20 -20 10 -3 The Blind -5 0 5 The Lavic -5 0 5 section [km] Lake section [km]

10 -4 10 -3 10 -2 10 -1 10 0 10 1 10 -2 10 -1 10 0 10 1 -1 10 -1 10 f g G G’ F F’ 0 0 10 -2 -10 [km] Number of events (normalized) Number of events -10

10 -2 [km] 10 -3 -20 -20 -5 0 5 -5 0 5 10 -4 The South [km] The Bend [km] Bullion section section 10 -3 10 -5 10 -2 10 -1 10 0 10 1 10 -2 10 -1 10 0 10 1 Normal distance from the best fitting plane [km] Figure 3. 10 0 WS1 Parkfield 2004 WS1 : 0.06 km 10 -1 WS2 : 0.56 km

-2 10 WS2

10 -3

10 -4 a Mean number of events (normalized) Mean number of events 10 -5 10 -2 10 -1 10 0 10 1 Normal distance from the fault plane [km]

10 -1 10 -1

WS1 WS1

10 -2 10 -2 Hector Mine 1999 Hector Mine 1999 WS2 ? Blind WS2 ? Lavic Lake fault section fault section WS1 : 0.86 km WS1 : 1.12 km WS2 : - WS2 : - b c Mean number of events (normalized) Mean number of events 10 -3 10 -3 10 -2 10 -1 10 0 10 1 10 -2 10 -1 10 0 10 1

10 -1 10 -1 WS1

WS1

10 -2

WS2 Hector Mine 1999 10 -2 Hector Mine 1999 10 -3 The Bend WS2 ? South Bullion fault section fault section WS1 : 0.73 km WS1 : 0.74 km -4 WS2 : - 10 WS2 : 3.36 km d e Mean number of events (normalized) Mean number of events 10 -3 10 -5 10 -2 10 -1 10 0 10 1 10 -2 10 -1 10 0 10 1 Normal distance from the fault plane [km] Normal distance from the fault plane [km] Figure 4. 10

Hector Mine Sup. Landers 1 Hills El Mayor South Cucapah (km) Napa S1

2 W S Morgan =0.3W 0.1 W S1 Hill S2 Parkfield =0.1W W S1

0.01 0.1 1 10 WS2 (km) Figure 5. 10 a b El Mayor 5 Cucapah Awatere Landers Hector Landers 1 Mine Hector El Mayor Awatere Mine Cucapah Sup. Hills Sup. (km)

(km) Hills S1 South S2

W Morgan 0.1 Napa W Hill Morgan South Napa Hill -0.54 Parkfield -0.39 WS1 = k*D WS2 = k*D 0.01 0.5 Parkfield 0.1 1 10 100 1000 0.1 1 10 100 1000 Cumulative fault displacement (km) Cumulative fault displacement (km) 10 c d

5 El Mayor Landers Cucapah 1 El Mayor Awatere Landers Cucapah Hector Mine Sup. Hills Hector Sup. Awatere Mine (km) Hills South (km) S1 Napa S2 Morgan W 0.1 W Hill Morgan Hill Parkfield -0.80 -0.62 W =k*A W =k*A South S1 S2 Napa Parkfield 0.01 0.5 0.1 1 10 100 0.1 1 10 100 Fault initiation age (Ma) Fault initiation age (Ma) 10 e f El Mayor 5 Cucapah Hector Landers 1 Landers Mine Awatere Sup.Hills Sup. Hector El Mayor Mine Awatere (km) Hills Cucapah (km) S1 Morgan S2

W Morgan South Hill W 0.1 Napa Hill

Parkfield -0.71 -0.46 South W =k*SR Parkfield S1 WS2 =k*SR Napa 0.01 0.5 0.11 10 100 0.1 1 10 100 Geological fault slip rate (mm/yr) Geological fault slip rate (mm/yr) Figure 6. 10 0 Field measurements across 100 a the Awatere fault (Little, 1995)

10 -1 Number of faults (%)

10 10 -2

10 -3 Parkfield 1 Increasing Morgan Hill South Napa -4 structural 10 maturity of Hector Mine fault sections Landers El Mayor Cucapah Superstition Hills Mean number of events (normalized) Mean number of events 10 -5 10 -2 10 -1 10 0 10 1 Normal distance from the fault plane [km]

b c WS2 Immature Immature fault

fault section crust section Undeformed Damage intensity Mature fault WS1 - + section

Increasing time Mature Mean number of EQs fault and slip crust section

Distance from the fault plane Undeformed Cataclasite fault core Figure 7. Data from this study 5 (km) S2 W

Theoretical models j=1.0 j=2.0 0.5 0.1 1 10 100 1000 Cumulative fault displacement (km) Figure 8. Earthquake stress drop (MPa) 10 1 1 Landers Sup. Hills Cumulative displacement of the the of displacement Cumulative broken fault section (km) section fault broken Hector Hector Mine 10 Morgan Hill 100 ∆σ=8.3*D Parkeld -0.23 1000 Figure 9. Earthquake stress drop (MPa) 10 1 . 10 0.1 a Landers Hector Hector Mine Geological slip rate (mm/yr) Geological slip rate (m rate slip Geological (mm/yr) rate slip Geological 1 Sup. Hills Morgan Hill ∆σ=5.4*SR Parkfield -0.22 100

Earthquake stress drop (MPa) 10 2 4 6 8 0 . 10 0.1 b Landers Hector Hector Mine 1 Sup. Hills ∆σ=-2*log(SR)+5.5 Morgan Hill m/yr) Parkfield 100