Distribution of seismicity across strike slip faults in California

Size: px
Start display at page:

Download "Distribution of seismicity across strike slip faults in California"

Transcription

1 Click Here for Full Article JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 115,, doi: /2008jb006234, 2010 Distribution of seismicity across strike slip faults in California Peter M. Powers 1 and Thomas H. Jordan 2 Received 1 December 2008; revised 5 November 2009; accepted 16 December 2009; published 11 May [1] The distribution of seismicity about strike slip faults provides measurements of fault roughness and damage zone width. In California, seismicity decays with distance from strike slip faults according to a power law (1 + x 2 /d 2 ) g/2. This scaling relation holds out to a fault normal distance x of 3 6 km and is compatible with a rough fault loading model in which the inner scale d measures the half width of a volumetric damage zone and the roll off rate g is governed by stress variations due to fault roughness. According to Dieterich and Smith s 2 D simulations, g approximates the fractal dimension of alongstrike roughness. Near fault seismicity is more localized on faults in northern California (NoCal, d = 60 ± 20 m, g = 1.65 ±.05) than in southern California (SoCal, d = 220 ± 40 m, g = 1.16 ±.05). The Parkfield region has a damage zone half width (d = 120 ± 30 m) consistent with the SAFOD drilling estimate; its high roll off rate (g = 2.30 ±.25) indicates a relatively flat roughness spectrum: k 1 versus k 2 for NoCal, k 3 for SoCal. Our damage zone widths (the first direct estimates averaged over the seismogenic layer) can be interpreted in terms of an across strike fault core multiplicity that is 1 in NoCal, 2 at Parkfield, and 3 in SoCal. The localization of seismicity near individual faults correlates with cumulative offset, seismic productivity, and aseismic slip, consistent with a model in which faults originate as branched networks with broad, multicore damage zones and evolve toward more localized, lineated features with low fault core multiplicity, thinner damage zones, and less seismic coupling. Our results suggest how earthquake triggering statistics might be modified by the presence of faults. Citation: Powers, P. M., and T. H. Jordan (2010), Distribution of seismicity across strike slip faults in California, J. Geophys. Res., 115,, doi: /2008jb Introduction [2] California, with its dense, well mapped network of faults and high quality earthquake catalogs, is an excellent setting to investigate seismicity variations in space and time. Earthquake catalogs constructed using improved hypocenter relocation techniques [Ellsworth et al., 2000; Hauksson and Shearer, 2005; Shearer et al., 2005; Thurber et al., 2006] are revealing new details about the 3 D geometry of fault networks [Yule and Sieh, 2003; Carena et al., 2004] and ruptures of large earthquakes [Liu et al., 2003], properties of nascent faults [Bawden et al., 1999], earthquake streaks observed on creeping sections [Rubin et al., 1999; Waldhauser et al., 1999, 2004; Shearer et al., 2005; Thurber et al., 2006], and the space time behavior of earthquake swarms [Vidale and Shearer, 2006]. These studies, as well as extensive research on the fractal character of fault systems [Tchalenko, 1970; King, 1983; Okubo and Aki, 1987; Hirata, 1989; Robertson et al., 1995; Ouillon et al., 1996; Kagan, 2007], have raised a number of interesting 1 Department of Earth Sciences, University of Southern California, Los Angeles, California, USA. 2 Southern California Earthquake Center, University of Southern California, Los Angeles, California, USA. Copyright 2010 by the American Geophysical Union /10/2008JB issues regarding the relationship of small earthquakes to major faults. For instance: Are physical properties of a fault zone such as the distribution of secondary faults and fractures, damage zone width, and fault roughness captured in near fault earthquake rates? Does the rate of small earthquakes in the vicinity of a major fault zone reflect the longterm fault slip rate or cumulative offset of the fault? Do other factors such as heat flow and lithology modulate earthquake rate? How might spatial models of near fault seismicity improve subsurface fault maps or models of earthquake triggering? [3] To address these issues, we analyze the variation of seismicity rate normal to near vertical strike slip faults in California and examine its relation to stress heterogeneity, damage zones, and degree of seismic coupling. Strike slip faults were chosen because their locations are constrained by mapped surface traces and their approximate bilateral symmetry of seismicity makes their earthquake distributions simpler to interpret than those of normal and reverse faults. To reveal systematic scaling relationships, we aggregated data from fault segments in a common class, as defined by geographic region, fault length, and aftershock activity. We restricted our analysis to small earthquakes (M w < 5), which we treated as point sources. Examined this way, the nearfault seismicity shows a power law decay away from the 1of25

2 Figure 1. Map of northern California showing locations of seismicity samples (gray boxes with reference numbers; see Table 3). Black lines delineate large faults, and the heavy black line marks the San Andreas fault. Dark gray dots mark the locations of 1.5 < M 2.5 earthquakes ( ). Reference x label background colors reflect the fault normal intercatalog location uncertainty of each segment: s TN for x the Parkfield segments (51 and 52); s UN for all others. fault surface [Powers and Jordan, 2005, 2007a; Hauksson, 2010]. [4] We examined regional variations in this fault normal distribution of seismicity by comparing the results from selected fault segments in northern California, between Parkfield and the San Francisco Bay (Figure 1), and a more extensive distribution of faults in southern California (Figure 2). In the northern California region, the fault segments are on or subparallel to the San Andreas master fault, a larger percentage of the faults are creeping [Irwin, 1990], and little seismicity extends below 10 km [Hill et al., 1990] (Figure 3). In contrast, the southern California segments have deeper seismicity [Hauksson, 2000], show little or no creep [Bodin et al., 1994; Lyons and Sandwell, 2003; Shearer et al., 2005], and often intersect at high angles. In southern California, we also investigated the spatial distribution of seismicity near smaller faults (e.g., splays of large faults and unmapped secondary faults) to see how fault length and along fault variations affect the scaling relations. We declustered each data set and incorporated fault segments that ruptured during large (M w > 6) earthquakes to constrain how the scaling relations are modified by after- 2of25

3 Figure 2. Map of southern California showing locations of seismicity samples (with reference numbers; see Table 3). Black lines delineate large faults, and the heavy black line marks the San Andreas fault. Grey boxes with reference numbers in circles (1 15) indicate the limits of seismicity samples about large strike slip faults; reference numbers in squares indicate the locations of samples about small (16 29) and aftershock dominated (30 41) fault segments. Dark gray dots mark the locations of 1.5 < M 2.5 earthquakes ( ). Reference label background colors reflect the fault normal intercatalog location uncertainty, s x PS, of each segment. shock activity. To assess the bias and variance in the scaling parameters caused by event mislocation, we supplemented the errors estimated by the hypocenter location algorithms with constraints from intercatalog comparisons and statistical simulations. [5] Our results provide measures of shear localization on faults and indirect evidence that fault damage zones extend through the seismogenic crust. We present a model of fault behavior that incorporates slip on a fractal fault, and discuss the implications of the model on fault evolution using data on fault width, on fault earthquake density, cumulative offset, and aseismic slip rate. Through comparisons with studies of exhumed faults and drilling results, we relate the structure of a fault damage zone at depth, as defined by seismicity, to near surface observations. 2. Fault Referenced Seismicity Catalogs [6] Hypocenters from six earthquake catalogs (three in southern, two in northern California, and one for the Parkfield region (Table 1)) were used to constrain near fault seismicity distributions and their uncertainties. The Southern California Seismic Network (SCSN) is the southern part of the California Integrated Seismic Network (CISN), a region within the Advanced National Seismic System (ANSS). The SCSN catalog (here abbreviated as S ; available at 3of25

4 Figure 3. Depth distributions of relocated earthquakes in the fault segment catalogs. Events from (a) catalog H and (b) catalog P in southern California and (c) catalogs U and T in northern California. The darker shaded bars mark the medial 90% of all events in the catalogs. Note that seismicity is generally shallower in northern California (Figure 3c). The downward bias in catalog H (Figure 3a) is likely an artifact of the 3 D velocity model used for earthquake relocation. is the standard catalog for southern California and contains events reported by all networks in the region. It includes some estimates of hypocentral errors, but all events have a quality designation indicative of maximum horizontal and vertical location uncertainties. [7] Hauksson and Shearer [2005] relocated events (catalog H ; available at from the SCSN using the double difference algorithm of Waldhauser and Ellsworth [2000]. They cross correlated waveforms to measure traveltime differences and relocated seismicity in a three dimensional (3 D) velocity model. For events lacking sufficient data for double differencing (<10%), they determined hypocenters using Hauksson s [2000] relocation method. They also evaluated errors using this method, because the double difference code does not compute hypocentral location errors for large data sets. [8] Shearer et al. [2005] relocated events (catalog P ; available at using a sourcespecific station term algorithm [Richards Dinger and Shearer, 2000] that employs a layered (1 D) velocity model. Using event similarity data from a waveform cross correlation analysis, they further refined the locations of spatially related events ( 60%) via a cluster analysis. Hypocentral errors reported by Shearer et al. [2005] correspond to the relative locations of events in each cluster. [9] The catalog of the Northern California Seismic Network (NCSN, abbreviated N ; available at ncedc.org/ncsn) is the standard catalog for the northern part of the CISN. Ellsworth et al. [2000] relocated a subset of the NCSN events in the San Francisco Bay area (catalog U ; available at using the double difference algorithm of Waldhauser and Ellsworth [2000]. The catalog does not include hypocentral location errors, but Ellsworth et al. [2000] report average horizontal and vertical location uncertainties of 0.1 and 0.5 km, respectively. [10] On the creeping section of the San Andreas fault in the vicinity of Parkfield, relocated events (denoted catalog T ; available at BSSA_html/bssa_96 4b/05825 esupp/) were taken from Thurber et al. [2006]. They constructed an improved 3 D wave speed model to first determine station corrections and then relocated events via double difference using a combination of event cross correlation differential times and traveltime differences from NCSN phase picks. The catalog does not include any hypocentral error information. [11] For each of the six catalogs, we limited our analysis to events that occurred from the beginning of 1984 until the end of To allow for intercatalog comparisons, the events were required to have hypocenters in both northern California catalogs (N and U or T and U) or in all three southern California catalogs (S, P, and H). The northern California catalogs, T and U, do not overlap geographically and may only be compared independently to catalog N, whereas the southern California catalogs span the entire lower half of the state and may be compared collectively. Events identified as quarry blasts or ones with intercatalog separations greater than 10 km were discarded. [12] We constructed catalogs of earthquakes for five classes of near vertical strike slip faults (Tables 2 and 3 and Table 1. Earthquake Catalog Sources Sources by Region ID a Type Southern California SCSN S standard Shearer et al. [2005] P relocated Hauksson and Shearer [2005] H relocated Northern California NCSN N standard Ellsworth et al. [2000] U relocated Thurber et al. [2006], Parkfield T relocated a Used to reference catalog in equations and text. 4of25

5 Table 2. Earthquake Catalog Statisitics N T a (events) b M c Southern California (SoCal) SCSN catalog 291, Fault classes Large 11, Small 8, Aftershock dominated 19, Northern California (NoCal) NCSN catalog 47, Fault classes Large 16, Parkfield 3, a All events common to regional and relocated catalogs in Table 1. Figures 1 and 2): large faults of northern California, the San Andreas fault at Parkfield, large faults of southern California, small faults of southern California, and fault segments with abundant aftershocks of major southern California earthquakes. For large faults, we chose relatively straight segments of named, throughgoing faults (e.g., Figure 4), eliminated fault segments with earthquake density of less than one event per km, and avoided fault junctions and zones of structural complexity. In southern California the SCEC Community Fault Model (CFM, available at [Plesch et al., 2007] was our guide for fault selection; in northern California, where faults are well defined by near vertical seismicity distributions, we used surface traces. Where the surface trace or seismicity along a particular fault indicates significant changes in strike or fault strand overlap, we restricted our selection to smaller fault segments; e.g., the Hayward fault (Figure 1; segments 42 43) and Garlock fault (Figure 2; segments 1 3). Large fault lengths, L k in Table 3, average 21 km in northern California and 47 km in southern. [13] We identified small and aftershock dominated faults by sets of earthquakes that define linear structural features of shorter length (average 9 km). The small fault class comprises splays off larger faults and unmapped secondary faults. Aftershock dominated segments were selected from faults activated by the 1992 Joshua Tree (M w 6.1), 1992 Landers (M w 7.3), or 1999 Hector Mine (M w 7.1) earthquakes. Although the seismicity in the aftershock dominated class spans the entire 19 year length of the source catalogs, it is dominated by aftershocks from these large events. The seismicity of the small fault class is more uniformly distributed in time. [14] We used the most recently updated magnitude data from catalogs S and N; these were generally reported as local magnitude, although there are a few events for which magnitudes were computed using other means. The maximum likelihood Gutenberg Richter b values [Aki, 1965] for the northern and southern California catalogs are 0.8 and 0.9, respectively (Table 2). Northern California has a lower magnitude of completeness (M c = 1.2) than southern California (M c = 2.0), reflecting its smaller area and higher station density. Notes to Table 3: a Down weighted level N 0, Fitting distance x max. b From relocated catalogs P, U, and T as described in text. Table 3. Fault Segment Catalogs a Segment Segment Name Size N k (events) L k (km) W k b (km) SoCal Large (N 0 = 1000, x max = 6 km) 1 Garlock (East) Garlock (Central) Garlock (West) Lenwood Lockhart San Andreas (Mojave) Santa Cruz Catalina Ridge Palos Verdes Newport Inglewood (North) Newport Inglewood (South) Elsinore Temecula San Jacinto (Anza) 5, Elsinore Coyote Mt Cerro Prieto Imperial 1, San Andreas (Coachella) Total (N T ) 11, Length weighted average 13.2 SoCal Small (N 0 = 700, x max = 2.5 km) 16 Scodie Lineament 1, San Jacinto (Anza) 1, San Jacinto (Anza) 1, San Jacinto (Anza) San Jacinto (Coyote Creek) San Jacinto (Anza) San Jacinto (Coyote Creek) San Jacinto (Anza) San Jacinto (Borrego) Superstition Mt Elmore Ranch Elmore Ranch (western ext.) Elmore Ranch (western ext.) Elsinore Total (N T ) 8, Length weighted average 9.2 SoCal Aftershock Dominated (N 0 = 1500, x max = 3 km) 30 Joshua Tree 1, Joshua Tree 2, Joshua Tree 1, Joshua Tree Landers Landers Landers 1, Landers 2, Landers 1, Hector Mine Hector Mine 2, Hector Mine 2, Total (N T ) 19, Length weighted average 8.2 NoCal Large (N 0 = 1500, x max = 3 km) 42 Hayward (North) Hayward (South) Calaveras (North) 2, Calaveras (Central) Calaveras (South) 1, Sargent 1, San Andreas Creeping (North) 1, San Andreas Creeping (Central) 3, San Andreas Creeping (South) 5, Total (N T ) 16, Length weighted average 7.9 NoCal Parkfield (No Down Weight, x max = 3 km) 51 San Andreas Parkfield (North) 3, San Andreas Parkfield (South) Total (N T ) 3, Length weighted average 8.4 5of25

6 Figure 4. Example of intercatalog earthquake location variation. (a) Map of the Elsinore fault (segment 10, Figure 2) showing relocated seismicity of catalog P and historic (heavy black lines), Holocene (thin black lines), and late Quaternary (thin gray lines) faults. (b) Depth section across the map showing locations of events in the standard catalog S relative to an initial, 3 D fault model based estimate of the fault trace (heavy dashed line); earthquakes are the same magnitude ranges as on the map. (c) Fault normal distribution of events. (d) Depth section across the map for relocated catalog P. (e) Fault normal seismicity distribution of relocated events. Note the difference in horizontal bias (black arrow in Figures 4c and 4e) of peak seismicity between the standard and relocated catalog. [15] We filtered the hypocentral depths in each fault catalog to focus on the central part of the seismogenic crust. Averaged across all fault segments, 90% of seismicity falls between 2 and 10 km for catalogs U and T and 2.5 and 17 km for catalog P (Figure 3). The upper 5% of events tend to occur within 2 km of the free surface, and so we set 2 km as an upper truncation depth. The lower cutoff shows significant variation reflecting regional differences in seismogenic thickness [Hauksson, 2000; Magistrale, 2002]. We therefore set the lower truncation depth to exclude the deepest 5% of hypocenters in each fault catalog. The differences between the lower and upper truncation depths for the kth fault segment determined the segment width W k ; values for each fault segment catalog are listed with the fault lengths L k in Table 3. [16] For each fault segment, we established a fault oriented coordinate system by fitting a plane to the seismicity. An initial estimate of the fault plane was derived from the CFM or, in the absence of a fault model, from a vertical plane that approximated the mapped surface or epicenter trace. All parameters of the plane were perturbed to obtain a least squares fit to relocated hypocenters from catalogs P, U or T within 2 km of this initial fault plane. For all fault segments considered, a 4 km wide swath captures most 6of25

7 Figure 5. Example of how a fault seismicity based coordinate system localizes events on a fault and minimizes artificial fault normal dispersion. (a) Map of the southern Hayward fault (segment 43, Figure 1) showing relocated seismicity of catalog U; fault age representations are the same as in Figure 4. The fault strand cutting across the lower right corner of the map is the northern Calaveras fault. (b) Fault normal depth section across the fault trace prior to aligning coordinate system to relocated seismicity; earthquakes are the same magnitude ranges as in the map, and the heavy dashed line marks the depth projection of the fault surface trace. (c) Fault normal distribution of events. (d) Fault normal depth section across the fault trace after aligning coordinate system to a best fit plane to relocated seismicity. (e) Realigned fault normal distribution. Note that the fault seismicity based coordinate system yields a narrower event distribution in Figure 5e. near fault earthquakes while ignoring off fault clusters that would contribute to misalignment of the coordinate system with the fault plane. Narrower swaths fail to include nearfault events for fault segments where seismicity exhibits a significant horizontal shift from the initial estimate, as in the case of the Elsinore fault (Figure 4). Rotations permitted by the fitting process further localize events on final, seismicity based fault planes, as is observed on the Hayward fault (Figure 5). [17] The hypocenters from the relevant relocated catalog were transformed into a local Cartesian system defined by an origin at one end of the surface trace of the best fit fault plane, a near vertical z axis, a y axis along the fault strike, and an x axis perpendicular to the fault plane. A final, faultreferenced catalog was then constructed by eliminating events with relocated x coordinates greater than ±15 km, relocated y coordinates beyond the ends of the fault segment, and relocated z coordinates outside the depth limits 7of25

8 Table 4. Intercatalog Error Estimates Catalog Pair (AB) Bias (km) a Standard Deviation (km) b x AB b y AB b z AB s x AB s x AB b s y AB s z AB SoCal Large HS PS HP SoCal Small HS PS HP SoCal Aftershock Dominated HS PS HP NoCal UN (Large) TN (Parkfield) See the auxiliary material for individual fault values. a Event weighted mean absolute values. b The s x AB estimated by uniform reduction. described above. The fault segment catalogs are summarized in Table Intercatalog Analysis [18] A proper description of near fault seismicity distributions requires careful attention to mislocation errors. Information about such errors can be determined from comparisons of hypocenters determined by different methods [e.g., Shearer et al., 2005]. In the present study, we have quantified the intercatalog comparisons on a fault segment basis. For the kth fault segment with N k events common to the catalog pair A and B, we define the intercatalog faultnormal bias by b kx AB ¼ 1 X ðx i A N xi B Þ; k i2k and the fault normal variance by ð kx AB Þ2 ¼ 1 X ðx i A N k 1 xi B i2k bkx AB Þ2 : i Here, x A is the fault normal coordinate of the ith event, and the summation implied by i 2 k is over all N k events associated with the kth fault segment. Similar expressions can be written for the fault parallel and near vertical directions. The intercatalog bias and variance computed for each catalog pair (UN and TN in northern California and HS, PS, and HP in southern California) are listed by individual fault segment in Tables S1 S3 in the auxiliary material. 1 [19] For each coordinate of the fault oriented reference frame, the mean intercatalog bias for a fault class was computed by taking the absolute values of the segment biases, weighting them by the number of events for each segment, and averaging over all segments. Likewise, the 1 Auxiliary materials are available in the HTML. doi: / 2008JB ð1þ ð2þ mean standard deviation was calculated as the square root of the event weighted segment variances. These averages are given in Table 4. For all classes, the intercatalog biases and standard deviations are largest for the z coordinate, reflecting the uncertainty in estimating hypocentral depth. No systematic differences are observed between the two horizontal coordinate statistics, x and y. In southern California, the z coordinate statistics are especially large, in part because shallow events in the early part of Catalog S were often assigned a default depth of 6 km, and also because the velocity model used to relocate events in Catalog H tends to bias events downward (Figure 3). Because we are interested in the fault normal distribution of seismicity, we focus our x discussion on b AB and s x AB. [20] The intercatalog statistics for well instrumented Parkfield region are higher than those of the northern California fault class, particularly the fault normal values x x (e.g., s TN = 0.62 km versus s UN = 0.31 km). These differences are primarily due to more stations and a greater number of earthquakes being used in the relocation procedure, as well as a strong velocity contrast across the San Andreas fault near Parkfield, which is modeled in the Thurber et al. [2006] relocations but not in the standard catalog. [21] The intercatalog statistics for the large faults in x southern California (e.g., s PS = 1.04 km) are also substantially higher than for the northern California fault class, which we ascribe to several factors. The southern region has a more heterogeneous crustal structure than the northern region, such as larger and deeper sedimentary basins, increasing the location errors. Moreover, the faults sampled in northern California were restricted to well instrumented regions of the San Andreas system near the center of the NCSN. In southern California, a number of the large faults are peripheral to the SCSN, and they invariably show bigger intercatalog variations (Figure 2). For instance, s PS for the coastal Newport Inglewood fault (segments 8 9) is 2.1 km, and it reaches 2.7 km for the Cerro Prieto fault (segment 13), which is located in Mexico outside the SCSN. In contrast, the values for the more centrally located San Jacinto fault (segments 17 24) are less than 1 km. [22] In southern California, the large fault class has a higher intercatalog standard deviation than either the smallfault or aftershock classes (e.g., s PS = 1.04, 0.75, 0.55 km, respectively). The fault normal biases show a similar ordering (e.g., b PS = 0.54 km, 0.43 km, 0.23 km). The network x geometry again plays a role, because the latter two classes comprise segments that tend to be more centrally located within the SCSN. In addition, the estimator given by equation (1) accounts only for a constant translational bias; for long segments, other parameters, such as a rotational bias, may be needed to represent the catalog differences, especially for faults on the periphery of the network. The inadequacy of the bias model acts to increase the apparent intercatalog variance. [23] The standard deviations between the two relocated southern California catalogs (HP) are consistently lower than those involving the standard catalog (HS and PS), satisfying the expectation that relocation reduces the hypocentral variance (Table 4). However, histograms of the faultnormal differences for all three catalog combinations show heavy tailed distributions with outliers that dominate the variance estimates. Most of these outliers can be explained x x 8of25

9 Table 5. Catalog Specific Error Estimates a Catalog (A) Bias (km) b,c Standard Deviation b (km) Catalog Error (km) b x A b y A b z A s x A s x A d s y A s z A Horizontal Vertical SoCal Large S (standard) P (relocated) H (relocated) SoCal Small S (standard) P (relocated) H (relocated) SoCal Aftershock Dominated S (standard) P (relocated) H (relocated) NoCal Large N (standard) U (relocated) e NoCal Parkfield N (standard) T (relocated) NR NR a See the auxiliary material for individual fault values. NR, not reported. b Computed assuming statistical independence. c Event weighted mean absolute values. d The s x A estimated by uniform reduction. e Catalog averages as reported by Ellsworth et al. [2000]. by the way the different location algorithms respond to anomalous travel times (e.g., picking blunders, large path anomalies). To account for outliers, we applied a method of uniform reduction [Jeffreys, 1932; Buland, 1986] in which we modeled the differences as the superposition of a Gaussian distribution and a nearly uniform distribution. The standard deviations of the best fit Gaussians are listed in Table 4. The largest reductions, more than 80% in variance, are obtained for the HP intercatalog differences. The reduced standard deviations were used to characterize the event mislocations in our subsequent analysis of the errors in the fault normal scaling parameters. [24] If the event location errors from the three southern California catalogs are assumed to be statistically independent (possibly a poor assumption) then we can determine catalog specific biases, b A kx, and standard deviations, s A kx,by solving the three equations for intercatalog bias: b kx AB ¼ bkx A bkx B ; and the three for intercatalog variance: kx AB 2¼ kx A ð3þ 2þ kx 2; B ð4þ where AB = {PS, HS, HP}. Equations (3) are not linearly independent, and we therefore included the additional constraint that the biases of the individual catalogs should sum to zero, which minimizes the overall bias. Similar sets of equations can be solved for the fault parallel and depth directions. We averaged the event weighted absolute values of the fault segment biases to obtain the values in Table 5 (see the auxiliary material for individual fault segment data). [25] The s values for the relocated catalogs are substantially smaller than those for the standard SCSN catalog, as expected from the intercatalog comparisons, and the s values for the Shearer et al. [2005] catalog are in all cases smaller than those for the Hauksson and Shearer [2005] catalog. In x particular, the values of s P obtained from the reduced intercatalog standard deviations are only about half the size of s x H, which is consistent with the qualitative observation that the cluster analysis relocation method used to develop the P catalog provides significantly better localization of hypocenters into fault like structures [Shearer et al., 2005]. For this preferred southern California catalog, the faultnormal standard deviations are less than 0.1 km for all three fault classes. As noted above, our linear removal of bias did not consider possible intercatalog rotations, which could skew the intercatalog variance to higher values, whereas possible correlations in the hypocenter errors between different catalogs would skew them to lower values. On the balance, s x P 0.1 km appears to be a good estimate. [26] In Table 5, we compare the results of the intercatalog error analysis with formal location errors listed in the individual catalogs. The standard deviations in depth from the latter sources are always larger than the corresponding mean horizontal standard deviation, in rough agreement with the intercatalog analysis, but the magnitudes are rather different. The mislocation errors included in the standard network catalogs are substantially larger than our computed values. The reverse is generally true for the relocated catalogs, though the agreement is much better. The confidence region of fault normal hypocentral error for the relocated catalogs is km with the high end of the range represented by a few fault segments at the periphery of the southern 9of25

10 California network. For the most part, location error is <0.1 km, consistent with the intercatalog analysis. Further checks on the mislocation errors from seismicity modeling, described below, support the intercatalog analysis. 4. Fault Normal Seismicity Distributions [27] Because strike slip faults in California are nearly vertical, we developed our scaling relations using the faultnormal distance x as the independent variable, ignoring bilateral asymmetry in seismicity. We stacked the seismicity data in each fault group and computed earthquake density as a function of distance x using a nearest neighbor method [Silverman, 1986] in which the bins are adjusted to contain q neighboring events. For each data set we experimented with a range of q values and selected one that yielded an adequate point density for deriving fault normal scaling relations (10 q 50). [28] Logarithmic plots of earthquake density versus x for each regional catalog indicate fault normal distributions that have flat peaks within a few hundred meters of the fault, rolloff as an inverse power law for about an order of magnitude in distance, and merge with irregular backgrounds at distances less than 10 km (Figures 6 9). Near the fault, the Figure 6. Fault normal earthquake density distributions for large faults in southern California. Distributions for (a) relocated catalog P, (b) relocated catalog H, and (c) standard catalog S using a nearest neighbor bin interval of q =50 events. The heavy dashed line marks the limit to which we fit data, x max ; beyond this limit, background seismicity dominates. The black line is a maximum likelihood fit of an inverse power law, with asymptotic slope ~, to observations within that limit. The inner scale of the distribution is described by d. ~ Figure 7. Fault normal earthquake density distributions for (a) small faults and (b) aftershock dominated fault segments in southern California. Both Figures 7a and 7b use events from relocated catalog P with a bin interval of q = 50 events. Features are the same as in Figure of 25

11 parameter, we found that the maximum likelihood fits to the observed fault normal distributions were obtained for m 2. We also experimented with exponential and Gaussian distributions but found they were a poor fit to the data. [29] Assuming m = 2, we obtained a maximum likelihood fit of equation (5) to the binned data for each fault group out to a maximum fault normal distance x max, chosen such that the relative contributions from background seismicity were small; the values of x max for each fault class are listed in Figure 8. Fault normal earthquake density distributions for large faults in northern California. Distributions for (a) relocated catalog U and (b) standard catalog N using a nearestneighbor bin interval of q = 20 events. The heavy dashed line marks the limit to which we fit data, x max ; beyond this limit, background seismicity dominates. As in Figures 6 and 7, the black line is a maximum likelihood fit of an inverse power law to observations within that limit. observed distributions can be described by the functional form: ðxþ ¼ 0 d m =m : ð5þ jj x m þ d m In expression (5), d is an inner scale that removes the power law singularity on the fault, g is the asymptotic roll off of seismicity away from the fault, and the exponent m controls the shape of the distribution for x d, i.e., the sharpness of the corner on a logarithmic plot. By varying the latter Figure 9. Fault normal earthquake density distributions for the Parkfield fault segments. Distributions for (a) relocated catalog T and (b) standard catalog N using a nearest neighbor bin interval of q = 10 events. Features are the same as in Figure of 25

12 We assume the observed value, n j, in each bin is Poisson distributed, which yields the log likelihood function [e.g., Boettcher and Jordan, 2004]: XJbins ð 0 ;;dþ ¼ n j ln n j ðxþ nj ðxþ lnðnj!þ ; ð7þ j¼1 Maximizing (7) using a linear approximation to n(x) over each binning interval (adequate for the small intervals used here) yields the estimates ~ 0, ~, and d. ~ These estimates have correlated errors. However, we note that the maximum likelihood estimator for ~ 0 is N max / R x max (1 + x 2 /d 2 ) g/2 dx, 0 where N max is the cumulative number of events out to x max. If N max is large, its relative error is small ( N 1/2 max, the standard population error) and uncorrelated with the errors in ~ and d. ~ The latter are positively correlated, as shown in Figure 10, which plots the maximum likelihood estimates and confidence intervals for the various fault groups in the ~ d ~ plane. [30] We checked the error estimates from the maximum likelihood procedure with those derived from jackknife resampling [Efron, 1979]. Generally speaking, the two were in agreement, but where they differed, we used the larger estimate. We experimented with the lower magnitude cutoff and depth ranges and found the results to be robust. We also tested a range of bin widths, q, and found little variation in our results. [31] An important issue is the weighting of individual fault segments in the seismicity stacking. Owing to the variability in seismicity rates, the number of earthquakes per fault segment ranges from a hundred to several thousand (Table 3), and our results will depend on how each is weighted. In our stacking procedure, we applied a positive weighting factor w k to each event in the kth fault segment catalog, which we computed by w k ¼ min½1; N 0 =N k Š; ð8þ Figure 10. Maximum likelihood solutions and errors for ~ and d. ~ Values for (a) southern California large faults, (b) northern California large faults, and (c) Parkfield showing the positive correlation between scaling parameters. Light gray and black ovals mark the 68% and 95% confidence bounds, respectively. Table 3. The expected number of events in the jth bin of width Dx is the integral: n j ¼ Z jxjj jx jj x ðxþdx: ð6þ where N 0 is a down weighted level that was held constant for each fault group. We varied N 0 from N min, the minimum of all catalog sizes N k in each fault group, to N max, the maximum in each group. The latter bound corresponds to one event one vote (w k = 1), whereas the former corresponds to one catalog one vote (w k 1/N k ). For intermediate values, events from catalogs larger than N 0 were down weighted by the ratio N 0 /N k, while those from smaller catalogs received unit weight. We experimented with a range of down weight levels for each structural group and found that the maximum likelihood estimates for most of the parameters were stable across a wide range between N min and N max (Figure 11). Table 3 lists the actual values used in deriving the parameter values discussed below. [32] For all structural groups, there is a well defined scaling region of at least an order of magnitude in faultnormal distance where the earthquake density shows a power law roll off before it merges with the background seismicity (Figures 6 9). In Table 6, we list by fault group the maximum likelihood estimates of ~ and ~ d. To assess the effects of earthquake clustering, we declustered the fault segment catalogs using the algorithm of Reasenberg [1985] with default parameters (r fact = 10, x k = 0.5, x meff = 1.5, t 0 = 2 days, t max = 10 days, p 1 = 0.99) and obtained the distri- 12 of 25

13 bution parameters for the declustered catalogs and the event clusters. [33] Table 6 shows interesting variations across the fault groups and catalog types. Comparing the relocated catalogs P and U, seismicity decays away from the large faults in southern California at a significantly lower rate (~ = 0.98 ± 0.04) than it decays in northern California (1.60 ± 0.04) or for the small faults in southern California (1.37 ± 0.08). In southern California, clustered events decay more rapidly than independent events, in agreement with the higher decay rate for aftershocks of large southern California earthquakes (~ = 1.50 ± 0.08). [34] For the larger faults, the apparent inner scale ~ d for relocated catalogs is smaller in northern California (0.08 ± 0.01 km) than in southern California (0.23 ± 0.03 km). The former is comparable to the relocation uncertainty. Significantly higher values are obtained for the standard catalogs ( km), consistent with more dispersion due to mislocation. Figure 11. Down weighted analysis results for small faults in southern California: (a) ~ and (b) d ~ vary with down weighted value across the different catalogs. The dashed gray line marks our selected value of N 0 = 700. Note that parameter estimates are largely stable within error for most down weighted values. Only at low values of N 0 do parameters start to vary as more box catalogs, including those with few events, are weighted equally. 5. Analysis of Bias [35] The results in Table 6 are biased by two factors: contributions from background seismicity and hypocentral error. The former is apparent when we relax the assumption of bilateral symmetry. We first identified which side of each fault segment had more events for x x max and then restacked the data for each of the five fault classes, preserving this asymmetry in seismic abundance (Figure 12 and the auxiliary material). The estimates of ~ obtained for the less abundant side were consistently higher than those on the more abundant side. The values of d ~ on the less abundant side of each fault class also increased slightly over those of the symmeterized distributions owing to the positive correlation between ~ and d. ~ We did find that, on the side with fewer events, the scaling region extended to greater distances from the fault, in some cases by an order of magni- Table 6. Apparent Fault Normal Scaling Parameters Catalog Whole Catalog Declustered Clusters ~d (km) ~ ~ d (km) ~ ~ d (km) ~ SoCal Large S (standard) 0.88 ± ± ± ± ± ± 0.12 P (relocated) 0.23 ± ± ± ± ± ± 0.05 H (relocated) 0.27 ± ± ± ± ± ± 0.05 SoCal Small S (standard) 0.79 ± ± ± ± ± ± 0.54 P (relocated) 0.19 ± ± ± ± ± ± 0.17 H (relocated) 0.23 ± ± ± ± ± ± 0.19 SoCal Aftershock Dominated S (standard) 0.55 ± ± 0.11 P (relocated) 0.33 ± ± 0.08 H (relocated) 0.31 ± ± 0.07 NoCal Large N (standard) 0.16 ± ± ± ± ± ± 0.07 U (relocated) 0.08 ± ± ± ± ± ± 0.05 NoCal Parkfield N (standard) 0.89 ± ± ± ± ± ± 1.60 T (relocated) 0.13 ± ± ± ± ± ± of 25

14 distribution close to x max. We show this by modeling the distribution in each fault class as the superposition of a fault normal decay of earthquake rate and a uniform background rate (Figures 12b and 13). Parkfield was excluded from the modeling because the background signal is weak and the scaling parameters therefore unbiased. We use the parameters of the low abundance side of the fault in each class because the contribution from background is significantly lower and the scaling only minimally biased. By varying the ratio of background to decaying events, we recovered the results reported in Table 6, as well as the distributions on the event heavy side of the fault for each fault class, suggesting that the fault normal decay of seismicity is similar across strike slip faults. Values for background corrected scaling parameters are reported in Table 7. [37] We investigated how the results are biased by hypocenter errors in two ways, theoretically and using Monte Carlo simulations. We assumed that the true fault normal Figure 12. Comparison of fault normal earthquake density for (a) low and (b) high productivity sides of a fault for catalog P about large, southern California faults using a nearest neighbor bin interval of q = 20 events. Features are the same as in Figures 6 9. Seismicity decays more rapidly on the low productivity side of a fault and spans almost 2 orders of magnitude. In Figure 12a, maximum likelihood fits of exponential (dashed) and Gaussian (dotted) distributions are shown for comparison. In Figure 12b, we additionally fit the data using the parameters from the lowproductivity side (Figure 12a) and include a uniform background rate, n bg. See the auxiliary material for distributions of other fault classes. tude, supporting our choice of power law model. Figure 12a shows that exponential and Gaussian distributions are a poor fit to the data. Fault maps show that the truncation of the scaling region on the abundant side can generally be explained by the seismicity increase from another fault branch or splay, proximate to the target fault segment, a common feature of the San Andreas system (e.g., Figure 5). [36] Uniform background seismicity associated with proximate faults significantly alters the shape of a density Figure 13. Comparison of fault normal earthquake density for (a) large faults in southern California (catalog P) with (b) that of a synthetic distribution using bin intervals of q = 50 events. Features are the same as in Figures 6 9. The synthetic distribution is one realization of a Monte Carlo simulation in which fault normal earthquake density is modeled as the superposition of an unbiased distribution (g = 1.3, d = 0.3 km; compare Figure 12a) that decays beyond x max (dashed gray line) and a uniform background. The maximum likelihood fit to the synthetic distribution out to x max recovers the scaling parameters determined in the initial analysis of symmeterized distributions. 14 of 25

15 Table 7. Bias Corrected Scaling Parameters Fault Class Catalog Hypocentral Error Correction Background Correction Simulated a d (km) g Theoretical a d (km) d (km) g SoCal large P 0.26 ± ± ± ± 0.05 SoCal small P 0.21 ± ± ± ± 0.10 SoCal aftershock dominated P 0.39 ± ± ± ± 0.10 NoCal large U 0.09 ± ± ± ± 0.05 NoCal Parkfield T 0.13 ± ± ± ± 0.25 a Bias corrections computed assuming s x A = 0.1 km. seismicity is governed by the distribution n(x) in equation (5), and that the catalogs have independent, identically distributed mislocation errors approximated by a zero mean Gaussian probability density function (pdf), g A ðxþ ¼ p 1 ffiffiffiffiffi 2 x A h i exp x 2 =2ð x A Þ2 ; ð9þ where s x A is the standard error for catalog A. The pdf for the observed seismicity can then be computed as the convolution of the two (normalized) distributions: p A ðxþ ¼ Z 1 1 ðx Þg A ðþd= Z 1 1 ðþd: ð10þ A little analysis shows that, if s A x /d is not too large (less than 5 or so), p A (x) can be approximated by equation (5) with an asymptotic slope ~ = g and an inner scale ~ d computed as the intersection of the small x probability density with the large x asymptote, d= d ~ ¼ pffiffiffiffiffiffiffiffi Z 1 2= 1 þð x =m A x=dþm e x 2 =2 dx: ð11þ 0 The bias correction d d ~ derived by solving (11) is only weakly dependent on the shape parameter m, so we fixed it at its best fit value (m = 2). We experimented with s x A values spanning the confidence region of km recorded in the relocated catalogs. Because low values have only a minimal effect and the large values are only applicable to a few fault segments in southern California, the conservative estimate of s x A = 0.1 km determined from our intercatalog analysis is appropriate for A = P, U, and T. Figure 14 plots d ~ versus d for s x A = 0.1 km and we see that the correction is small for d > s x A and decreases with g. [38] The bias corrected estimates of d obtained from (11) are listed in Table 7. The largest correction, for large fault seismicity in northern California, changes the estimated inner scale from 0.09 km to 0.06 km, a difference of only 30 m. Note that the magnitude of bias in this worst pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi case, 0.03 km, is smaller than the quadratic estimator d 2 þ 2 A d = 0.05 km. We verified the small size of the bias correction using a Monte Carlo method in which we generated synthetic catalogs that satisfied (12), perturbed them with Gaussian noise (0.05 s x A 0.2 km), and calculated a likelihood score of their fit to the distribution curves obtained from the actual data. The values of d and g that maximized the likelihood for many ( 50) catalog realizations for s x A = 0.1 km are listed in Table 7; an example of a single realization is presented in Figure 15. The estimates of d are nearly identical to the theoretically corrected values. Moreover, the simulations provided bias corrections for g, which are not zero (as the asymptotic theory predicts) owing to the positive correlation between the estimators of d and g arising from a finite range of x (see Figure 10). However, the corrections to g are also small, 10% (for catalog T) or less. 6. Rough Fault Loading Model [39] Our multicatalog analysis of earthquake hypocenters in California reveals that the seismicity in the vicinity of strike slip faults can be represented by a three parameter distribution: d 2 =2 ðxþ ¼ 0 x 2 þ d 2 ; jxj x max: ð12þ The constant n 0 describes the fault normal seismic intensity (in events/km) on the fault surface. Using the data in Table 3, the intensity n(x) can be normalized by the total fault length S L k, as plotted in Figures 6 9, or by the total fault area S L k W k, which yields a spatial seismic density for the catalog interval T = 19 years. The inner scale d measures the Figure 14. Theoretical relationship between an observed inner scale, d,andthetruevalueofd ~ plotted for various g. Theory assumes that an observed fault seismicity scaling distribution is the product of the true distribution convolved with a Gaussian noise function with standard deviation of 0.1 km. Only at d < 0.2 km do the observed and true values diverge significantly. 15 of 25

16 Figure 15. Sample result from simulation analysis of parameter bias. In each simulation, synthetic distributions were perturbed with Gaussian noise (s x A = 0.1 km) until a combination of g and d was found that maximized the likelihood score of the fit to the original distribution. The simulation result pictured is for large faults of southern California and illustrates a good correlation between a perturbed synthetic distribution (dots) and our observed distribution (dashed line) for catalog P. half width of a near fault region where the seismic intensity is flat ( n 0 ), and the exponent g specifies the power law roll off of seismic intensity in the scaling region d < x x max. [40] Figure 16 presents a conceptual rough faulting loading (RFL) model that we will use to explain the seismicity behavior. Our starting point is the observation that fault surfaces can be described by a self affine (fractal) complexity over a large range of spatial scales [Power and Tullis, 1995; Lee and Bruhn, 1996; Renard et al., 2006] and evolve in time toward surfaces that are less complex in the direction of slip [Wesnousky, 1988; Stirling et al., 1996; Sagy et al., 2007; Finzi et al., 2009]. Here complexity refers to the fractal branching of faults into multiple surfaces [e.g., King, 1983; Hirata, 1989] as well as the fractal roughness of individual fault surfaces [e.g., Lee and Bruhn, 1996; Renard et al., 2006; Sagy et al., 2007]. To build a simple model, we begin by considering a single fault surface whose deviations from the planar approximation x = 0 define a fault normal topography [Saucier et al., 1992; Chester and Chester, 2000; Dieterich and Smith, 2009]. We represent an along strike (constant z) profile of this topography as the realization of a stationary stochastic process X(y) that has zero expectation, hx(y)i = 0, and a variogram D E 1=2: 2 2 ðyþ ¼ ½X ðy þ yþ XðyÞŠ 2 ð13þ the surface is self affine, then x Dy H, where 0 H 1is the Hausdorff measure (sometimes referred to as the Hurst exponent) of the along strike profile [Feder, 1988; Turcotte, 1997]. Assume the self affine scaling of fault roughness breaks down above some outer scale Dy outer related to the Figure 16. Schematic representation of the RFL (rough fault loading) model that explains observations of near fault seismicity distribution. (a) Tectonic loading of a self affine fault generates a heterogeneous stress field that yields a power law decay of seismicity over a scaling region via stress relaxation [Dieterich and Smith, 2009, Figure 3]. (b) Toward the fault core, small scale stress heterogeneities of the rough fault are attenuated by low fracture strength across a damage zone of width 2d km. (c) Illustration of how observed seismicity rates vary with distance from a fault (solid black line). The scaling region likely extends beyond x max, as indicated by the dashed black line, but is masked by interference from proximal fault branches. 16 of 25

Self-similar earthquake triggering, Båth s law, and foreshock/aftershock magnitudes: Simulations, theory, and results for southern California

Self-similar earthquake triggering, Båth s law, and foreshock/aftershock magnitudes: Simulations, theory, and results for southern California JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 117,, doi:10.1029/2011jb008957, 2012 Self-similar earthquake triggering, Båth s law, and foreshock/aftershock magnitudes: Simulations, theory, and results for southern

More information

Space-time clustering of seismicity in California and the distance dependence of earthquake triggering

Space-time clustering of seismicity in California and the distance dependence of earthquake triggering JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 117,, doi:10.1029/2012jb009471, 2012 Space-time clustering of seismicity in California and the distance dependence of earthquake triggering P. M. Shearer 1 Received

More information

Velocity contrast along the Calaveras fault from analysis of fault zone head waves generated by repeating earthquakes

Velocity contrast along the Calaveras fault from analysis of fault zone head waves generated by repeating earthquakes Click Here for Full Article GEOPHYSICAL RESEARCH LETTERS, VOL. 35, L01303, doi:10.1029/2007gl031810, 2008 Velocity contrast along the Calaveras fault from analysis of fault zone head waves generated by

More information

to: Interseismic strain accumulation and the earthquake potential on the southern San

to: Interseismic strain accumulation and the earthquake potential on the southern San Supplementary material to: Interseismic strain accumulation and the earthquake potential on the southern San Andreas fault system by Yuri Fialko Methods The San Bernardino-Coachella Valley segment of the

More information

Geo736: Seismicity and California s Active Faults Introduction

Geo736: Seismicity and California s Active Faults Introduction Geo736: Seismicity and California s Active Faults Course Notes: S. G. Wesnousky Spring 2018 Introduction California sits on the boundary of the Pacific - North American plate boundary (Figure 1). Relative

More information

Spatial and temporal stress drop variations in small earthquakes near Parkfield, California

Spatial and temporal stress drop variations in small earthquakes near Parkfield, California Click Here for Full Article JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 112,, doi:10.1029/2006jb004395, 2007 Spatial and temporal stress drop variations in small earthquakes near Parkfield, California Bettina

More information

Appendix O: Gridded Seismicity Sources

Appendix O: Gridded Seismicity Sources Appendix O: Gridded Seismicity Sources Peter M. Powers U.S. Geological Survey Introduction The Uniform California Earthquake Rupture Forecast, Version 3 (UCERF3) is a forecast of earthquakes that fall

More information

Supporting Information for Break of slope in earthquake-size distribution reveals creep rate along the San Andreas fault system

Supporting Information for Break of slope in earthquake-size distribution reveals creep rate along the San Andreas fault system GEOPHYSICAL RESEARCH LETTERS Supporting Information for Break of slope in earthquake-size distribution reveals creep rate along the San Andreas fault system Inessa Vorobieva, 1,2 Peter Shebalin, 1,2 Clément

More information

Limitations of Earthquake Triggering Models*

Limitations of Earthquake Triggering Models* Limitations of Earthquake Triggering Models* Peter Shearer IGPP/SIO/U.C. San Diego September 16, 2009 Earthquake Research Institute * in Southern California Why do earthquakes cluster in time and space?

More information

Aftershocks are well aligned with the background stress field, contradicting the hypothesis of highly heterogeneous crustal stress

Aftershocks are well aligned with the background stress field, contradicting the hypothesis of highly heterogeneous crustal stress JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 115,, doi:10.1029/2010jb007586, 2010 Aftershocks are well aligned with the background stress field, contradicting the hypothesis of highly heterogeneous crustal stress

More information

Does Aftershock Duration Scale With Mainshock Size?

Does Aftershock Duration Scale With Mainshock Size? GEOPHYSICAL RESEARCH LETTERS, VOL.???, NO., PAGES 1 16, Does Aftershock Duration Scale With Mainshock Size? A. Ziv A. Ziv, Ben-Gurion University of the Negev, Beer-Sheva, 84105, Israel. (e-mail: zival@bgu.ac.il)

More information

Jack Loveless Department of Geosciences Smith College

Jack Loveless Department of Geosciences Smith College Geodetic constraints on fault interactions and stressing rates in southern California Jack Loveless Department of Geosciences Smith College jloveless@smith.edu Brendan Meade Department of Earth & Planetary

More information

Model Uncertainties of the 2002 Update of California Seismic Hazard Maps

Model Uncertainties of the 2002 Update of California Seismic Hazard Maps Bulletin of the Seismological Society of America, Vol. 95, No. 6, pp. 24 257, December 25, doi: 1.1785/12517 Model Uncertainties of the 22 Update of California Seismic Hazard Maps by Tianqing Cao, Mark

More information

Lack of continuity of the San Andreas Fault in southern California: Three-dimensional fault models and earthquake scenarios

Lack of continuity of the San Andreas Fault in southern California: Three-dimensional fault models and earthquake scenarios JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 109,, doi:10.1029/2003jb002643, 2004 Lack of continuity of the San Andreas Fault in southern California: Three-dimensional fault models and earthquake scenarios Sara

More information

State of Stress in Seismic Gaps Along the SanJacinto Fault

State of Stress in Seismic Gaps Along the SanJacinto Fault ELEVEN State of Stress in Seismic Gaps Along the SanJacinto Fault Hirao Kanamori and Harold Magistrale NTRODUCTON Data from the Southern California Seismic Network have been extensively used to map spatial

More information

Geophysical Journal International

Geophysical Journal International Geophysical Journal International Geophys. J. Int. (21) 183, 433 45 doi: 1.1111/j.1365-246X.21.4745.x Quantifying focal mechanism heterogeneity for fault zones in central and southern California Iain W.

More information

CHAPTER 1 BASIC SEISMOLOGY AND EARTHQUAKE TERMINOLGY. Earth Formation Plate Tectonics Sources of Earthquakes...

CHAPTER 1 BASIC SEISMOLOGY AND EARTHQUAKE TERMINOLGY. Earth Formation Plate Tectonics Sources of Earthquakes... CHAPTER 1 BASIC SEISMOLOGY AND EARTHQUAKE TERMINOLGY Earth Formation... 1-2 Plate Tectonics... 1-2 Sources of Earthquakes... 1-3 Earth Faults... 1-4 Fault Creep... 1-5 California Faults... 1-6 Earthquake

More information

Today: Basic regional framework. Western U.S. setting Eastern California Shear Zone (ECSZ) 1992 Landers EQ 1999 Hector Mine EQ Fault structure

Today: Basic regional framework. Western U.S. setting Eastern California Shear Zone (ECSZ) 1992 Landers EQ 1999 Hector Mine EQ Fault structure Today: Basic regional framework Western U.S. setting Eastern California Shear Zone (ECSZ) 1992 Landers EQ 1999 Hector Mine EQ Fault structure 1 2 Mojave and Southern Basin and Range - distribution of strike-slip

More information

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 112, B12309, doi: /2007jb004986, 2007

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 112, B12309, doi: /2007jb004986, 2007 Click Here for Full Article JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 112,, doi:10.1029/2007jb004986, 2007 Applying a three-dimensional velocity model, waveform cross correlation, and cluster analysis to locate

More information

External Grant Award Number 04HQGR0058 IMPROVED THREE-DIMENSIONAL VELOCITY MODELS AND EARTHQUAKE LOCATIONS FOR CALIFORNIA

External Grant Award Number 04HQGR0058 IMPROVED THREE-DIMENSIONAL VELOCITY MODELS AND EARTHQUAKE LOCATIONS FOR CALIFORNIA External Grant Award Number 04HQGR0058 IMPROVED THREE-DIMENSIONAL VELOCITY MODELS AND EARTHQUAKE LOCATIONS FOR CALIFORNIA Clifford H. Thurber University of Wisconsin-Madison 1215 W. Dayton St. Madison,

More information

Earthquake Stress Drops in Southern California

Earthquake Stress Drops in Southern California Earthquake Stress Drops in Southern California Peter Shearer IGPP/SIO/U.C. San Diego September 11, 2009 Earthquake Research Institute Lots of data for big earthquakes (rupture dimensions, slip history,

More information

Kinematics of the Southern California Fault System Constrained by GPS Measurements

Kinematics of the Southern California Fault System Constrained by GPS Measurements Title Page Kinematics of the Southern California Fault System Constrained by GPS Measurements Brendan Meade and Bradford Hager Three basic questions Large historical earthquakes One basic question How

More information

Overview of the Seismic Source Characterization for the Palo Verde Nuclear Generating Station

Overview of the Seismic Source Characterization for the Palo Verde Nuclear Generating Station Overview of the Seismic Source Characterization for the Palo Verde Nuclear Generating Station Scott Lindvall SSC TI Team Lead Palo Verde SSC SSHAC Level 3 Project Tuesday, March 19, 2013 1 Questions from

More information

Anomalous early aftershock decay rate of the 2004 Mw6.0 Parkfield, California, earthquake

Anomalous early aftershock decay rate of the 2004 Mw6.0 Parkfield, California, earthquake GEOPHYSICAL RESEARCH LETTERS, VOL. 33,, doi:10.1029/2006gl026744, 2006 Anomalous early aftershock decay rate of the 2004 Mw6.0 Parkfield, California, earthquake Zhigang Peng, 1,2 John E. Vidale, 1 and

More information

Probing Mid-Mantle Heterogeneity Using PKP Coda Waves

Probing Mid-Mantle Heterogeneity Using PKP Coda Waves Probing Mid-Mantle Heterogeneity Using PKP Coda Waves Michael A.H. Hedlin and Peter M. Shearer Cecil H. and Ida M. Green Institute of Geophysics and Planetary Physics Scripps Institution of Oceanography,

More information

Peter Shearer 1, Robin Matoza 1, Cecily Wolfe 2, Guoqing Lin 3, & Paul Okubo 4

Peter Shearer 1, Robin Matoza 1, Cecily Wolfe 2, Guoqing Lin 3, & Paul Okubo 4 Characterizing fault zones and volcanic conduits at Kilauea and Mauna Loa volcanoes by large-scale mapping of earthquake stress drops and high precision relocations Peter Shearer 1, Robin Matoza 1, Cecily

More information

Tests of relative earthquake location techniques using synthetic data

Tests of relative earthquake location techniques using synthetic data JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 110,, doi10.1029/2004jb003380, 2005 Tests of relative earthquake location techniques using synthetic data Guoqing Lin and Peter Shearer Institute of Geophysics and

More information

Some aspects of seismic tomography

Some aspects of seismic tomography Some aspects of seismic tomography Peter Shearer IGPP/SIO/U.C. San Diego September 7, 2009 Earthquake Research Institute Part 1: Global Tomography P velocity perturbations 200 km 1000 km 2700 km MIT 2006

More information

Observational analysis of correlations between aftershock productivities and regional conditions in the context of a damage rheology model

Observational analysis of correlations between aftershock productivities and regional conditions in the context of a damage rheology model Geophys. J. Int. (2009) 177, 481 490 doi: 10.1111/j.1365-246X.2009.04145.x FAST TRACK PAPER Observational analysis of correlations between aftershock productivities and regional conditions in the context

More information

arxiv:physics/ v2 [physics.geo-ph] 18 Aug 2003

arxiv:physics/ v2 [physics.geo-ph] 18 Aug 2003 Is Earthquake Triggering Driven by Small Earthquakes? arxiv:physics/0210056v2 [physics.geo-ph] 18 Aug 2003 Agnès Helmstetter Laboratoire de Géophysique Interne et Tectonophysique, Observatoire de Grenoble,

More information

High-resolution image of Calaveras Fault seismicity

High-resolution image of Calaveras Fault seismicity JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 107, NO. B9, 2186, doi:10.1029/2001jb000633, 2002 High-resolution image of Calaveras Fault seismicity David P. Schaff, 1 Götz H. R. Bokelmann, and Gregory C. Beroza

More information

SUPPLEMENTAL INFORMATION

SUPPLEMENTAL INFORMATION GSA DATA REPOSITORY 2013310 A.M. Thomas et al. MOMENT TENSOR SOLUTIONS SUPPLEMENTAL INFORMATION Earthquake records were acquired from the Northern California Earthquake Data Center. Waveforms are corrected

More information

SCEC Community Fault and Velocity Models and how they might contribute to the DCPP seismic hazard assessment. Andreas Plesch Harvard University

SCEC Community Fault and Velocity Models and how they might contribute to the DCPP seismic hazard assessment. Andreas Plesch Harvard University SCEC Community Fault and Velocity Models and how they might contribute to the DCPP seismic hazard assessment Andreas Plesch Harvard University SCEC Unified Structural Representation (USR) The SCEC USR

More information

A century of oil-field operations and earthquakes in the greater Los Angeles Basin, southern California

A century of oil-field operations and earthquakes in the greater Los Angeles Basin, southern California A century of oil-field operations and earthquakes in the greater Los Angeles Basin, southern California Egill Hauksson 1, Thomas Göbel 1, Jean-Paul Ampuero 1, and Elizabeth Cochran 2 Downloaded 06/03/15

More information

Testing the stress shadow hypothesis

Testing the stress shadow hypothesis JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 110,, doi:10.1029/2004jb003277, 2005 Testing the stress shadow hypothesis Karen R. Felzer and Emily E. Brodsky Department of Earth and Space Sciences, University of

More information

Stress modulation on the San Andreas fault by interseismic fault system interactions Jack Loveless and Brendan Meade, Geology, 2011

Stress modulation on the San Andreas fault by interseismic fault system interactions Jack Loveless and Brendan Meade, Geology, 2011 Stress modulation on the San Andreas fault by interseismic fault system interactions Jack Loveless and Brendan Meade, Geology, 2011 A three step process: 1 - Assimilate plate boundary wide GPS data into

More information

The Seismogenic Thickness of the Southern California Crust

The Seismogenic Thickness of the Southern California Crust Bulletin of the Seismological Society of America, Vol. 94, No. 3, pp. 940 960, June 2004 The Seismogenic Thickness of the Southern California Crust by Julie J. Nazareth and Egill Hauksson Abstract The

More information

Empirical Green s Function Analysis of the Wells, Nevada, Earthquake Source

Empirical Green s Function Analysis of the Wells, Nevada, Earthquake Source Nevada Bureau of Mines and Geology Special Publication 36 Empirical Green s Function Analysis of the Wells, Nevada, Earthquake Source by Mendoza, C. 1 and Hartzell S. 2 1 Centro de Geociencias, Universidad

More information

Stress orientations at intermediate angles to the San Andreas Fault, California

Stress orientations at intermediate angles to the San Andreas Fault, California JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 109,, doi:10.1029/2004jb003239, 2004 Stress orientations at intermediate angles to the San Andreas Fault, California Jeanne L. Hardebeck and Andrew J. Michael U.S.

More information

Part 2 - Engineering Characterization of Earthquakes and Seismic Hazard. Earthquake Environment

Part 2 - Engineering Characterization of Earthquakes and Seismic Hazard. Earthquake Environment Part 2 - Engineering Characterization of Earthquakes and Seismic Hazard Ultimately what we want is a seismic intensity measure that will allow us to quantify effect of an earthquake on a structure. S a

More information

Spatial clustering and repeating of seismic events observed along the 1976 Tangshan fault, north China

Spatial clustering and repeating of seismic events observed along the 1976 Tangshan fault, north China Click Here for Full Article GEOPHYSICAL RESEARCH LETTERS, VOL. 34, L23309, doi:10.1029/2007gl031594, 2007 Spatial clustering and repeating of seismic events observed along the 1976 Tangshan fault, north

More information

A TESTABLE FIVE-YEAR FORECAST OF MODERATE AND LARGE EARTHQUAKES. Yan Y. Kagan 1,David D. Jackson 1, and Yufang Rong 2

A TESTABLE FIVE-YEAR FORECAST OF MODERATE AND LARGE EARTHQUAKES. Yan Y. Kagan 1,David D. Jackson 1, and Yufang Rong 2 Printed: September 1, 2005 A TESTABLE FIVE-YEAR FORECAST OF MODERATE AND LARGE EARTHQUAKES IN SOUTHERN CALIFORNIA BASED ON SMOOTHED SEISMICITY Yan Y. Kagan 1,David D. Jackson 1, and Yufang Rong 2 1 Department

More information

Imaging sharp lateral velocity gradients using scattered waves on dense arrays: faults and basin edges

Imaging sharp lateral velocity gradients using scattered waves on dense arrays: faults and basin edges 2017 SCEC Proposal Report #17133 Imaging sharp lateral velocity gradients using scattered waves on dense arrays: faults and basin edges Principal Investigator Zhongwen Zhan Seismological Laboratory, California

More information

Location uncertainty for a microearhquake cluster

Location uncertainty for a microearhquake cluster Analysis of location uncertainty for a microearhquake cluster: A case study Gabriela Melo, Alison Malcolm, Oleg Poliannikov, and Michael Fehler Earth Resources Laboratory - Earth, Atmospheric, and Planetary

More information

Fault Specific, Dynamic Rupture Scenarios for Strong Ground Motion Prediction

Fault Specific, Dynamic Rupture Scenarios for Strong Ground Motion Prediction Fault Specific, Dynamic Rupture Scenarios for Strong Ground Motion Prediction H. Sekiguchi Disaster Prevention Research Institute, Kyoto University, Japan Blank Line 9 pt Y. Kase Active Fault and Earthquake

More information

Central Coast Seismicity Locations. Jeanne Hardebeck US Geological Survey Menlo Park, CA

Central Coast Seismicity Locations. Jeanne Hardebeck US Geological Survey Menlo Park, CA Central Coast Seismicity Locations Jeanne Hardebeck US Geological Survey Menlo Park, CA 11/29/2011 Seismic Network Issues for Earthquake Location: - Many stations are single-component, difficult to identify

More information

Fault Processes on the Anza section of the San Jacinto Fault

Fault Processes on the Anza section of the San Jacinto Fault Fault Processes on the nza section of the San Jacinto Fault J. Steidl (USB), M. Gladwin, R. Gwyther (SIRO ust.), F. Vernon (USD) SUMMRY We propose to instrument the nza slip gap and transition areas of

More information

The magnitude distribution of earthquakes near Southern California faults

The magnitude distribution of earthquakes near Southern California faults JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 116,, doi:10.1029/2010jb007933, 2011 The magnitude distribution of earthquakes near Southern California faults Morgan T. Page, 1 David Alderson, 2 and John Doyle 3

More information

A systematic measurement of shear wave anisotropy in southern California Report for SCEC Award #15081 Submitted March 15, 2016

A systematic measurement of shear wave anisotropy in southern California Report for SCEC Award #15081 Submitted March 15, 2016 A systematic measurement of shear wave anisotropy in southern California Report for SCEC Award #15081 Submitted March 15, 2016 Investigators: Zhigang Peng (Georgia Tech) I. Project Overview... i A. Abstract...

More information

High Resolution Imaging of Fault Zone Properties

High Resolution Imaging of Fault Zone Properties Annual Report on 1998-99 Studies, Southern California Earthquake Center High Resolution Imaging of Fault Zone Properties Yehuda Ben-Zion Department of Earth Sciences, University of Southern California

More information

California foreshock sequences suggest aseismic triggering process

California foreshock sequences suggest aseismic triggering process GEOPHYSICAL RESEARCH LETTERS, VOL. 0, 1, doi:10.100/grl.50, 013 California foreshock sequences suggest aseismic triggering process Xiaowei Chen 1 and Peter M. Shearer 1 Received January 013; revised 9

More information

A GLOBAL MODEL FOR AFTERSHOCK BEHAVIOUR

A GLOBAL MODEL FOR AFTERSHOCK BEHAVIOUR A GLOBAL MODEL FOR AFTERSHOCK BEHAVIOUR Annemarie CHRISTOPHERSEN 1 And Euan G C SMITH 2 SUMMARY This paper considers the distribution of aftershocks in space, abundance, magnitude and time. Investigations

More information

2008 Monitoring Research Review: Ground-Based Nuclear Explosion Monitoring Technologies

2008 Monitoring Research Review: Ground-Based Nuclear Explosion Monitoring Technologies STRUCTURE OF THE KOREAN PENINSULA FROM WAVEFORM TRAVEL-TIME ANALYSIS Roland Gritto 1, Jacob E. Siegel 1, and Winston W. Chan 2 Array Information Technology 1 and Harris Corporation 2 Sponsored by Air Force

More information

Negative repeating doublets in an aftershock sequence

Negative repeating doublets in an aftershock sequence LETTER Earth Planets Space, 65, 923 927, 2013 Negative repeating doublets in an aftershock sequence X. J. Ma and Z. L. Wu Institute of Geophysics, China Earthquake Administration, 100081 Beijing, China

More information

Once you have opened the website with the link provided choose a force: Earthquakes

Once you have opened the website with the link provided choose a force: Earthquakes Name: Once you have opened the website with the link provided choose a force: Earthquakes When do earthquakes happen? On the upper left menu, choose number 1. Read What is an Earthquake? Earthquakes happen

More information

Supporting Online Material for

Supporting Online Material for www.sciencemag.org/cgi/content/full/1131692/dc1 Supporting Online Material for Localized Temporal Change of the Earth s Inner Core Boundary This PDF file includes: Materials and Methods Figs. S1 to S3

More information

Geophysical Journal International

Geophysical Journal International Geophysical Journal International Geophys. J. Int. (2015) 202, 1949 1968 GJI Seismology doi: 10.1093/gji/ggv259 Artefacts of earthquake location errors and short-term incompleteness on seismicity clusters

More information

SUPPLEMENTARY INFORMATION

SUPPLEMENTARY INFORMATION SUPPLEMENTARY INFORMATION DOI: 10.1038/NGEO1158 Characteristic Slip for 5 great earthquakes along the Fuyun fault in China. Klinger 1 Y., Etchebes 1 M., Tapponnier 2 P., and Narteau 1 C. 1 Institut de

More information

Comprehensive analysis of earthquake source spectra and swarms in the Salton Trough, California

Comprehensive analysis of earthquake source spectra and swarms in the Salton Trough, California JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 116,, doi:10.1029/2011jb008263, 2011 Comprehensive analysis of earthquake source spectra and swarms in the Salton Trough, California X. Chen 1 and P. M. Shearer 1

More information

RELOCATION OF THE MACHAZE AND LACERDA EARTHQUAKES IN MOZAMBIQUE AND THE RUPTURE PROCESS OF THE 2006 Mw7.0 MACHAZE EARTHQUAKE

RELOCATION OF THE MACHAZE AND LACERDA EARTHQUAKES IN MOZAMBIQUE AND THE RUPTURE PROCESS OF THE 2006 Mw7.0 MACHAZE EARTHQUAKE RELOCATION OF THE MACHAZE AND LACERDA EARTHQUAKES IN MOZAMBIQUE AND THE RUPTURE PROCESS OF THE 2006 Mw7.0 MACHAZE EARTHQUAKE Paulino C. FEITIO* Supervisors: Nobuo HURUKAWA** MEE07165 Toshiaki YOKOI** ABSTRACT

More information

of other regional earthquakes (e.g. Zoback and Zoback, 1980). I also want to find out

of other regional earthquakes (e.g. Zoback and Zoback, 1980). I also want to find out 4. Focal Mechanism Solutions A way to investigate source properties of the 2001 sequence is to attempt finding well-constrained focal mechanism solutions to determine if they are consistent with those

More information

Comparison of Short-Term and Time-Independent Earthquake Forecast Models for Southern California

Comparison of Short-Term and Time-Independent Earthquake Forecast Models for Southern California Bulletin of the Seismological Society of America, Vol. 96, No. 1, pp. 90 106, February 2006, doi: 10.1785/0120050067 Comparison of Short-Term and Time-Independent Earthquake Forecast Models for Southern

More information

UNIVERSITY OF CALGARY. Nontrivial Decay of Aftershock Density With Distance in Southern California. Javad Moradpour Taleshi A THESIS

UNIVERSITY OF CALGARY. Nontrivial Decay of Aftershock Density With Distance in Southern California. Javad Moradpour Taleshi A THESIS UNIVERSITY OF CALGARY Nontrivial Decay of Aftershock Density With Distance in Southern California by Javad Moradpour Taleshi A THESIS SUBMITTED TO THE FACULTY OF GRADUATE STUDIES IN PARTIAL FULFILLMENT

More information

A Time-Dependent Probabilistic Seismic-Hazard Model for California

A Time-Dependent Probabilistic Seismic-Hazard Model for California Bulletin of the Seismological Society of America, 90, 1, pp. 1 21, February 2000 A Time-Dependent Probabilistic Seismic-Hazard Model for California by Chris H. Cramer,* Mark D. Petersen, Tianqing Cao,

More information

ALM: An Asperity-based Likelihood Model for California

ALM: An Asperity-based Likelihood Model for California ALM: An Asperity-based Likelihood Model for California Stefan Wiemer and Danijel Schorlemmer Stefan Wiemer and Danijel Schorlemmer 1 ETH Zürich, Switzerland INTRODUCTION In most earthquake hazard models,

More information

On the nucleation of creep and the interaction between creep and seismic slip on rate- and state-dependent faults

On the nucleation of creep and the interaction between creep and seismic slip on rate- and state-dependent faults Click Here for Full Article GEOPHYSICAL RESEARCH LETTERS, VOL. 34, L15303, doi:10.1029/2007gl030337, 2007 On the nucleation of creep and the interaction between creep and seismic slip on rate- and state-dependent

More information

Magnitude-Area Scaling of Strike-Slip Earthquakes. Paul Somerville, URS

Magnitude-Area Scaling of Strike-Slip Earthquakes. Paul Somerville, URS Magnitude-Area Scaling of Strike-Slip Earthquakes Paul Somerville, URS Scaling Models of Large Strike-slip Earthquakes L Model Scaling (Hanks & Bakun, 2002) Displacement grows with L for L > > Wmax M

More information

Earthquake Tectonics and Hazards on the Continents June Recognizing and characterizing strike-slip faults and earthquakes in USA

Earthquake Tectonics and Hazards on the Continents June Recognizing and characterizing strike-slip faults and earthquakes in USA 2464-4 Earthquake Tectonics and Hazards on the Continents 17-28 June 2013 Recognizing and characterizing strike-slip faults and earthquakes in USA S. G. Wesnousky Univ. of Nevada USA Strike-slip Faults

More information

Source Duration Scales with Magnitude Differently For Earthquakes on the San Andreas Fault and on Secondary Faults in Parkfield, CA

Source Duration Scales with Magnitude Differently For Earthquakes on the San Andreas Fault and on Secondary Faults in Parkfield, CA Source Duration Scales with Magnitude Differently For Earthquakes on the San Andreas Fault and on Secondary Faults in Parkfield, CA Rebecca M. Harrington University of California, Los Angeles Dept. of

More information

What allows seismic events to grow big?: Insights from fault roughness and b-value analysis in stick-slip experiments

What allows seismic events to grow big?: Insights from fault roughness and b-value analysis in stick-slip experiments What allows seismic events to grow big?: Insights from fault roughness and b-value analysis in stick-slip experiments T. H. W. Goebel 1, G. Kwiatek 2, T. W. Becker 3, G. Dresen 2, E.E. Brodsky 1 1 UC Santa

More information

Yesterday scaling laws. An important one for seismic hazard analysis

Yesterday scaling laws. An important one for seismic hazard analysis Yesterday scaling laws An important one for seismic hazard analysis Estimating the expected size of an earthquake expected on a fault for which the mapped length is known. Wells & Coppersmith, 1994 And

More information

Plate Boundary Observatory Working Group Plan for the San Andreas Fault System

Plate Boundary Observatory Working Group Plan for the San Andreas Fault System Introduction Plate Boundary Observatory Working Group Plan for the San Andreas Fault System This document puts forward a draft implementation plan for the San Andreas Fault component of the Plate Boundary

More information

SONGS SSC. Tom Freeman GeoPentech PRELIMINARY RESULTS

SONGS SSC. Tom Freeman GeoPentech PRELIMINARY RESULTS SONGS SSC Tom Freeman GeoPentech PRELIMINARY RESULTS Focused Questions Summarize the tectonic setting What is the seismogenic thickness? Are you including deep ruptures in the upper mantle (~30 km)? Do

More information

ETH Swiss Federal Institute of Technology Zürich

ETH Swiss Federal Institute of Technology Zürich Swiss Federal Institute of Technology Zürich Earthquake Statistics using ZMAP Recent Results Danijel Schorlemmer, Stefan Wiemer Zürich, Swiss Seismological Service, Switzerland Contributions by: Matt Gerstenberger

More information

Simulations in the Los Angeles Basin

Simulations in the Los Angeles Basin 83 Chapter 4 Simulations in the Los Angeles Basin This chapter reports the responses of steel moment-resisting frame (MRF) buildings to simulated earthquakes in the Los Angeles basin. I use broadband ground

More information

29th Monitoring Research Review: Ground-Based Nuclear Explosion Monitoring Technologies

29th Monitoring Research Review: Ground-Based Nuclear Explosion Monitoring Technologies IMPROVING MAGNITUDE DETECTION THRESHOLDS USING MULTI-STATION, MULTI-EVENT, AND MULTI-PHASE METHODS David Schaff and Felix Waldhauser Lamont-Doherty Earth Observatory, Columbia University Sponsored by Air

More information

Geophysical Research Letters

Geophysical Research Letters Geophysical Research Letters 16 JANUARY 2007 Volume 34 Number 1 American Geophysical Union Toward a more objective determination of seismic hazards Andaman Islands deformation associated with the 2004

More information

Probabilities for Jumping Fault Segment Stepovers

Probabilities for Jumping Fault Segment Stepovers GEOPHYSICAL RESEARCH LETTERS, VOL.???, XXXX, DOI:1.129/, Probabilities for Jumping Fault Segment Stepovers Bruce E. Shaw Lamont-Doherty Earth Observatory, Columbia University, New York James H. Dieterich

More information

Asymmetric distribution of early aftershocks on large faults in California

Asymmetric distribution of early aftershocks on large faults in California Asymmetric distribution of early aftershocks on large faults in California Ilya Zaliapin 1 and Yehuda Ben-Zion 2 1 Department of Mathematics and Statistics, University of Nevada, Reno, 89557 (zal@unr.edu)

More information

Short-Term Earthquake Forecasting Using Early Aftershock Statistics

Short-Term Earthquake Forecasting Using Early Aftershock Statistics Bulletin of the Seismological Society of America, Vol. 101, No. 1, pp. 297 312, February 2011, doi: 10.1785/0120100119 Short-Term Earthquake Forecasting Using Early Aftershock Statistics by Peter Shebalin,

More information

Mechanics of Earthquakes and Faulting

Mechanics of Earthquakes and Faulting Mechanics of Earthquakes and Faulting Lecture 20, 30 Nov. 2017 www.geosc.psu.edu/courses/geosc508 Seismic Spectra & Earthquake Scaling laws. Seismic Spectra & Earthquake Scaling laws. Aki, Scaling law

More information

GEOPHYSICAL RESEARCH LETTERS, VOL. 34, LXXXXX, doi: /2007gl031077, 2007

GEOPHYSICAL RESEARCH LETTERS, VOL. 34, LXXXXX, doi: /2007gl031077, 2007 Click Here for Full Article GEOPHYSICAL RESEARCH LETTERS, VOL. 34, LXXXXX, doi:10.1029/2007gl031077, 2007 2 Magnitude scaling relations from P-waves in southern California 3 Louisa L. H. Tsang, 1,2 Richard

More information

Rotation of the Principal Stress Directions Due to Earthquake Faulting and Its Seismological Implications

Rotation of the Principal Stress Directions Due to Earthquake Faulting and Its Seismological Implications Bulletin of the Seismological Society of America, Vol. 85, No. 5, pp. 1513-1517, October 1995 Rotation of the Principal Stress Directions Due to Earthquake Faulting and Its Seismological Implications by

More information

FULL MOMENT TENSOR ANALYSIS USING FIRST MOTION DATA AT THE GEYSERS GEOTHERMAL FIELD

FULL MOMENT TENSOR ANALYSIS USING FIRST MOTION DATA AT THE GEYSERS GEOTHERMAL FIELD PROCEEDINGS, Thirty-Eighth Workshop on Geothermal Reservoir Engineering Stanford University, Stanford, California, February 11-13, 2013 SGP-TR-198 FULL MOMENT TENSOR ANALYSIS USING FIRST MOTION DATA AT

More information

crustal volume of the swarm. This initiated interest and concern that another

crustal volume of the swarm. This initiated interest and concern that another 3. Earthquae locations Data from permanent seismic networs located the 2001 Enola mainshoc within the crustal volume of the 1982-84 swarm. This initiated interest and concern that another 1982-lie sequence

More information

Data Repository: Seismic and Geodetic Evidence For Extensive, Long-Lived Fault Damage Zones

Data Repository: Seismic and Geodetic Evidence For Extensive, Long-Lived Fault Damage Zones DR2009082 Data Repository: Seismic and Geodetic Evidence For Extensive, Long-Lived Fault Damage Zones Fault Zone Trapped Wave Data and Methods Fault zone trapped waves observed for 2 shots and 5 local

More information

EVALUATION OF TSUNAMI RISK FOR COASTAL SOUTHERN CALIFORNIA CITIES. Appendix A - Figures

EVALUATION OF TSUNAMI RISK FOR COASTAL SOUTHERN CALIFORNIA CITIES. Appendix A - Figures EVALUATION OF TSUNAMI RISK FOR COASTAL SOUTHERN CALIFORNIA CITIES Appendix A - Figures Figure 1.Map showing major faults of the southern California region. Northwest-trending faults are rightslip in character.

More information

Journal of Geophysical Research Letters Supporting Information for

Journal of Geophysical Research Letters Supporting Information for Journal of Geophysical Research Letters Supporting Information for InSAR observations of strain accumulation and fault creep along the Chaman Fault system, Pakistan and Afghanistan H. Fattahi 1, F. Amelung

More information

UCERF3 Task R2- Evaluate Magnitude-Scaling Relationships and Depth of Rupture: Proposed Solutions

UCERF3 Task R2- Evaluate Magnitude-Scaling Relationships and Depth of Rupture: Proposed Solutions UCERF3 Task R2- Evaluate Magnitude-Scaling Relationships and Depth of Rupture: Proposed Solutions Bruce E. Shaw Lamont Doherty Earth Observatory, Columbia University Statement of the Problem In UCERF2

More information

BEM Model of slip on the Channel Islands Thrust, CA

BEM Model of slip on the Channel Islands Thrust, CA BEM Model of slip on the Channel Islands Thrust, CA Credit Where Credit is Due: Michele Cooke Michele Cooke UMass Amherst Has been training students/postdocs to work with and remesh the CFM since at least

More information

Observation of shear-wave splitting from microseismicity induced by hydraulic fracturing: A non-vti story

Observation of shear-wave splitting from microseismicity induced by hydraulic fracturing: A non-vti story Observation of shear-wave splitting from microseismicity induced by hydraulic fracturing: A non-vti story Petr Kolinsky 1, Leo Eisner 1, Vladimir Grechka 2, Dana Jurick 3, Peter Duncan 1 Summary Shear

More information

Zhigang Peng. Curriculum Vitae. School of Earth and Atmospheric Sciences Phone: Georgia Institute of Technology Fax:

Zhigang Peng. Curriculum Vitae. School of Earth and Atmospheric Sciences Phone: Georgia Institute of Technology Fax: Zhigang Peng Curriculum Vitae School of Earth and Atmospheric Sciences Phone: 404-894-0231 Georgia Institute of Technology Fax: 404-894-5638 Office: ES&T Building, Rm. 2256 Email: zhigang.peng@eas.gatech.edu

More information

Aftershock From Wikipedia, the free encyclopedia

Aftershock From Wikipedia, the free encyclopedia Page 1 of 5 Aftershock From Wikipedia, the free encyclopedia An aftershock is a smaller earthquake that occurs after a previous large earthquake, in the same area of the main shock. If an aftershock is

More information

A RE-EXAMINATION OF HISTORIC EARTHQUAKES IN THE SAN JACINTO FAULT ZONE, CALIFORNIA

A RE-EXAMINATION OF HISTORIC EARTHQUAKES IN THE SAN JACINTO FAULT ZONE, CALIFORNIA Bulletin of the Seismological Society of America, Vol. 81, No. 6, pp. 2289-2309, December 1991 A RE-EXAMINATION OF HISTORIC EARTHQUAKES IN THE SAN JACINTO FAULT ZONE, CALIFORNIA BY ALLISON L. BENT* AND

More information

Compound earthquakes on a bimaterial interface and implications for rupture mechanics

Compound earthquakes on a bimaterial interface and implications for rupture mechanics Geophys. J. Int. (13) xxx, 1 14 Compound earthquakes on a bimaterial interface and implications for rupture mechanics Enning Wang 1, Allan M. Rubin 1, Jean-Paul Ampuero 1 Department of Geosciences, Princeton

More information

SUPPLEMENTARY INFORMATION

SUPPLEMENTARY INFORMATION doi:10.1038/nature11492 Figure S1 Short-period Seismic Energy Release Pattern Imaged by F-net. (a) Locations of broadband seismograph stations in Japanese F-net used for the 0.5-2.0 Hz P wave back-projection

More information

Scaling Relations for Seismic Cycles on Mid-Ocean Ridge

Scaling Relations for Seismic Cycles on Mid-Ocean Ridge Scaling Relations for Seismic Cycles on Mid-Ocean Ridge Transform Faults Margaret S. Boettcher University of New Hampshire, Dept. of Earth Sciences, Durham, NH 03824 Jeffrey J. McGuire Woods Hole Oceanographic

More information

A three-dimensional crustal seismic velocity model for southern California from a composite event method

A three-dimensional crustal seismic velocity model for southern California from a composite event method JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 112,, doi:10.1029/2007jb004977, 2007 A three-dimensional crustal seismic velocity model for southern California from a composite event method Guoqing Lin, 1,2 Peter

More information

Lab 6: Earthquake Focal Mechanisms (35 points)

Lab 6: Earthquake Focal Mechanisms (35 points) Lab 6: Earthquake Focal Mechanisms (35 points) Group Exercise 1. Drawing Nodal Planes (8 pts) The outline map below labeled Figure 4.60a shows the positions of two earthquakes recorded on the Mid-Atlantic

More information

High-resolution Time-independent Grid-based Forecast for M >= 5 Earthquakes in California

High-resolution Time-independent Grid-based Forecast for M >= 5 Earthquakes in California High-resolution Time-independent Grid-based Forecast for M >= 5 Earthquakes in California A. Helmstetter, Yan Kagan, David Jackson To cite this version: A. Helmstetter, Yan Kagan, David Jackson. High-resolution

More information