Download as a PDF

5 downloads 140 Views 5MB Size Report
The Rialto-Colton fault is interpreted southwest of ... compared eight state-of-the-art noninvasive seismic sources .... Arrows indicate vibration point locations. ...... Park et al. (1995) investigated the San Jacinto basin, about 40 km southeast of ...
Bulletin of the Seismological Society of America, Vol. 92, No. 6, pp. 2504–2520, August 2002

Delineation of Faulting and Basin Geometry along a Seismic Reflection Transect in Urbanized San Bernardino Valley, California by W. J. Stephenson, J. K. Odum, R. A. Williams, and M. L. Anderson

Abstract Fourteen kilometers of continuous, shallow seismic reflection data acquired through the urbanized San Bernardino Valley, California, have revealed numerous faults between the San Jacinto and San Andreas faults as well as a complex pattern of downdropped and uplifted blocks. These data also indicate that the Loma Linda fault continues northeastward at least 4.5 km beyond its last mapped location on the southern edge of the valley and to within at least 2 km of downtown San Bernardino. Previously undetected faults within the valley northeast of the San Jacinto fault are also imaged, including the inferred western extension of the Banning fault and several unnamed faults. The Rialto-Colton fault is interpreted southwest of the San Jacinto fault. The seismic data image the top of the crystalline basement complex across 70% of the profile length and show that the basement has an overall dip of roughly 10⬚ southwest between Perris Hill and the San Jacinto fault. Gravity and aeromagnetic data corroborate the interpreted location of the San Jacinto fault and better constrain the basin depth along the seismic profile to be as deep as 1.7 km. These data also corroborate other fault locations and the general dip of the basement surface. At least 1.2 km of apparent vertical displacement on the basement is observed across the San Jacinto fault at the profile location. The basin geometry delineated by these data was used to generate modeled ground motions that show peak horizontal amplifications of 2–3.5 above bedrock response in the 0.05- to 1.0-Hz frequency band, which is consistent with recorded earthquake data in the valley. Introduction Delineating the geometry of the basin sediment/bedrock interface is a critical step toward evaluating possible basin effects induced by earthquakes. Numerous recent modeling studies in southern California have demonstrated the effects basin geometry can have on ground motions (e.g., Frankel, 1993; Pitarka and Irikura, 1996; Alex and Olsen, 1998; Graves et al., 1998). Currently, one of the areas of concentrated focus for earthquake hazards is the San Bernardino Valley (Fig. 1). The valley is bounded between the San Jacinto fault on the southwest and San Andreas fault on the northeast and between the San Gabriel mountains on the northwest and the badlands and Crafton hills on the southwest. Estimating the hazard from the San Jacinto fault zone is critical to an overall assessment of hazard in southern California because of its historic seismicity rate, which is high even when compared to the adjacent San Andreas fault (Working Group on California Earthquake Probabilities [WGCEP], 1988). Additionally, the San Jacinto fault cuts across major infrastructure such as railroads, Interstate 10, and Interstate 215. The Riverside–San Bernardino metropolitan area is home to 3.2 million people, with about 180 thousand of those within the San Bernardino Valley (U.S.

Census Bureau 1999 estimates). It is thus an area that could be seriously impacted by even a moderate earthquake on the San Andreas, San Jacinto, or other faults in the area. Frankel (1993) simulated ground-motion effects across the San Bernardino Valley from a M 6.5 earthquake on the San Andreas Fault and noted that basin edge effects such as reflected surface waves were a major contributor to large ground motions. However, the basin (throughout this manuscript, “basin” refers to the subsurface nonbedrock region under the valley) velocity model for that study was not well constrained at depth by either geological or geophysical data. The current San Bernardino basin reference threedimensional velocity model proposed by the Southern California Earthquake Center (SCEC version 2) generally has gently dipping layers between the San Jacinto and San Andreas faults (Magistrale et al., 2000). However, variation in amplification across the San Bernardino Valley from the 1992 Landers and 1999 Hector Mine Earthquakes suggests more steeply dipping structure within the basin sedimentary fill (R. Graves, written commun., 2000; Graves and Wald, 2000). The seismic reflection method has been successfully

2504

Delineation of Faulting and Basin Geometry along a Seismic Reflection Transect in Urbanized San Bernardino Valley

2505

Figure 1. Geographic setting of the San Bernardino Valley, California. The seismic line of this study is shown by the heavy black line. Well locations are shown by triangles (34G4, 34Q1, 22L17, LV3, and CR5). Fault traces are at inferred or mapped locations prior to this study. Opposing gray arrows on seismic lines indicate the location for the San Jacinto fault interpreted from new seismic and gravity data. Double line labeled “R.P.” is the location of refraction profile by Hadley and Combs (1974). Map modified from Anderson et al. (2000) and Woolfenden and Kadhim (1997). used in the nearby Los Angeles basin to delineate basin geometry and active faulting in heavily urbanized areas (e.g., Pratt et al., 1998; Shaw and Shearer, 1999). High-resolution sources (typically 50–500 Hz) have proven very successful at imaging sediment–basement interfaces as deep as 500 m but generally lack energy needed to image deeper in noisy urban areas. Although oil industry active seismic sources such as explosives and vibroseis have sufficient energy to image the basement surface at many kilometers depth, their use in a modern urban environment can be both financially and legally difficult, if not impossible. Doll et al. (1998) compared eight state-of-the-art noninvasive seismic sources

including hammers, weight drops, Mini-Sosie, variously sized vibroseis sources, and a land air gun. They concluded the Mini-vib by Industrial Vehicles, (a smaller-scale vibroseis) provided the best subsurface image at their test site in Tennessee. Because the Mini-vib is noninvasive and portable and therefore appropriate for an urban setting, we tested the viability of using this source to image basement at 1 km or greater depth in the urbanized San Bernardino Valley. We acquired a single seismic reflection profile that begins and ends near bedrock outcrops (Perris Hill and La Loma Hills, respectively) (Fig. 1). This profile also traversed the deepest part of basin as inferred by gravity measurements

2506

W. J. Stephenson, J. K. Odum, R. A. Williams, and M. L. Anderson

(e.g., Willingham, 1968; Lambert, 1987; Anderson et al., 2000), and crossed the projected locations of numerous geologically mapped faults including the Loma Linda and San Jacinto faults. These seismic data, combined with results from gravity and aeromagnetic modeling, reveal a basin geometry that has implications for seismic hazard in the region.

Geologic Setting of San Bernardino Basin As defined by Morton and Matti (1993), the San Bernardino basin is a triangular-shaped region framed by the San Jacinto and San Andreas faults. Topographically, the basin narrows to the northwest between the San Gabriel and San Bernardino Mountains and is bounded on the south by the San Timoteo Badlands and on the southeast by the Crafton Hills. The tectonic history and structural development of the basin is a direct result of its location within the convergent region between the San Andreas and San Jacinto fault zones. This convergence is expressed as an area of compression and uplift (eastern San Gabriel Mountains) that contrast with the San Bernardino basin, which is an area of extension and pull-apart subsidence (Morton and Matti, 1993). The major structural blocks that compose the physiographic region around San Bernardino consist of the San Bernardino Mountains block to the north, the Peninsular Ranges block to the southwest, and the San Gabriel Mountains block to the west (Matti et al., 1985). Prior to basin development, this area was a topographic high consisting of exposed pre-Tertiary igneous and metamorphic rock. From approximately 3.8–4(?) Ma to about 0.8–1.3 Ma, the San Bernardino high shed coarse-grained clastic material to the west and southwest. These sediments form the extensive San Timoteo Badlands, (Dutcher and Garrett, 1963; Morton and Matti, 1993). The transformation of the San Bernardino high into a basin began approximately 1.5 Ma with the inception of the San Jacinto fault zone. The San Jacinto fault is one the most active in the region, with an average slip rate of 20 mm/yr over the past 1.5 m.y. (Morton and Matti, 1993). Angular convergence of the evolving San Jacinto fault zone with the main strand of the San Andreas fault resulted in an area of extension between the fault zones and the initiation of San Bernardino basin subsidence at approximately 1 Ma (Morton and Matti, 1993). The San Bernardino basin stratigraphic section reflects its youthful geologic age. In general, other than the oldest fill materials, basin fill stratigraphy consists of unconsolidated, coarser grained, alluvial-fan detritus derived from the surrounding mountains. These deposits interfinger with coarse to finer grained clastic fluvial deposits (Dutcher and Garrett, 1963). The underlying basement complex is composed of pre-Tertiary igneous and various grade metamorphic rocks including the Pelona schist. Examples of basement complex rocks can be found in remnant hills within the valley (e.g., Shandin Hills, Perris Hill, La Loma Hills, and Crafton Hills), (Fig. 1), which protrude above the basin fill.

There is little documented information on the oldest basin fill lithology. As the San Bernardino area transformed from a sediment source area to one of sediment accumulation, deposition of the Pliocene San Timoteo Formation in the nearby badlands ceased at approximately 1 Ma (Morton and Matti, 1993). A thin accumulation of continentally derived material that is pre–San Timoteo Formation unconformably overlies the basement complex and locally occupies low areas of the subsiding topography. Based upon water-well drilling logs, Dutcher and Garrett (1963) described discontinuous units as much as 150–200 m thick of continentally derived, calcareous indurated clays, sands, and conglomerates that may represent these oldest basin fill deposits. Deposited upon this unit, or more commonly, directly upon the basement complex, are Tertiary to Quaternary weakly to well-compacted units consisting of interfingered gravel, sand, silt, and clay. Although little information exists on the areal extend of these units, logs from sparse drill holes indicate that locally the section may be 200–300 m thick. Eckis (1934) suggested that these units are equivalent to the upper part of the San Timoteo in age and composition. A Pleistocene unit referred to as “Older Alluvium” by Dutcher and Garrett (1963) unconformably overlies these deposits. The Older Alluvium unit consists of unconsolidated coarse gravel, sand, silt, and clay deposited primarily by fluvial and alluvial fan processes and may be as thick as 150– 200 m. These units are in turn overlain by “Younger Alluvium,” which consist of up to 35 m of unconsolidated, coarse gravel, sand, silt, and clay. Whereas the distinction between coarse-grained, permeable, and fine-grained hydrogeologic stratigraphy is important for groundwater modeling within the San Bernardino basin, the majority of the Younger and Older Alluvium are believed to be seismically homogeneous based on their lithologic description. These hydrogeologic units should not be confused with the seismic–stratigraphic units defined in the following section.

Seismic-Reflection Data Analysis and Interpretation Acquisition and Processing Approximately 14 km of seismic-reflection data were acquired northeast to southwest through the city of San Bernardino, from Perris Hill to La Loma Hills. The continuous seismic profile crossed approximately the southernmost 75% of the San Bernardino basin including the projected traces of the San Jacinto, Rialto-Colton, and Loma Linda faults (Fig. 1). The San Andreas fault, approximately 3 km north of Perris Hill, was not traversed. Previous near-surface seismicreflection studies in urbanized areas of the Los Angeles region have typically imaged no deeper than about 500 m using earth impacting sources (e.g., Stephenson et al., 1995; Pratt et al., 1998). A small vibrator source, or Mini-vib, was utilized for this work because it provided the capability to image deeper than was done in these previous near-surface investigations. The Mini-vib was also selected because it

Delineation of Faulting and Basin Geometry along a Seismic Reflection Transect in Urbanized San Bernardino Valley

was more portable (about 50% smaller) than conventional industry vibroseis trucks and was capable of sweeping to frequencies as high as 500 Hz. Based on noise tests, however, a 10- to 100-Hz linear sweep was ultimately used for acquisition. We acquired 120 channels of data per shot point, using a 10-m shot and receiver-station interval. An asymmetric shooting geometry was used, where the source was walked through the first 30 stations of each receiver array before the array was moved forward. This approach was chosen because the recording system did not have 120-channel rollalong capability. A single, vertical-component, 8-Hz resonant frequency geophone was used at each receiver station. Data signal-to-noise ratio was low to moderate, as revealed on typical shot records that have only automatic gain correction (AGC) and applied bandpass filter of 10–20–100–200 Hz (Fig. 2). The low-cut ramp of the filter of 10–20 Hz (0% and 100% pass points, respectively) effectively eliminated

Figure 2.

2507

much of the low-frequency surface-wave noise, but numerous high-frequency linear and random noise events are still evident. Refracted and reflected wave amplitudes at the far offsets are often barely above background noise, even using a 400-msec AGC window. Seismic data processing was generally conventional by oil-industry standards, as shown in Table 1. Significant effort was dedicated to the removal of surface waves and linear noise events generated by high traffic volume on the I-10 and I-215 freeways and on nearby urban streets. Several processing steps, including surgical muting, dip filtering, and prestack eigenvector filtering were utilized for this purpose. Although the data had a moderate to high level of noise from the urban environment, maintaining a nominal 60 fold common-midpoint (CMP) stack coverage significantly improved the final interpreted stack. The final migrated and depth-converted stacked section is shown in Figure 3A. Depth conversion was performed

Six representative seismic-reflection shot records acquired in the San Bernardino Valley. Station 180 is near Perris Hill, and station 1419 is near La Loma Hills (see Fig. 1). Each record contains 120 channels with a 2-sec recording time. Data have only AGC and a bandpass filter applied. Arrows indicate vibration point locations. Gray bracket is the receiver array length for each shot. A high urban noise level is apparent in these records. The relatively high nominal 60-fold coverage was invaluable in successfully stacking in reflection signal as was stacking eight records per vibration point.

2508

W. J. Stephenson, J. K. Odum, R. A. Williams, and M. L. Anderson

Table 1 Generalized Processing Steps for San Bernardino Seismic Reflection Profile Processing Step

Data reformat Vibroseis correlation Geometry Trace edit AGC Bandpass filter (Hz) Top Mute Surgical mute Elevation statics Sort to CMP Adaptive deconvolution Dip scan filter NMO

Residual statics

Eigenvector filtering DMO Stack Migration

Eigenvector filtering Time-to-depth conversion

Description

Convert from SEG 2 to ProMAX威 internal Correlate on ground force signal, 12-sec sweep, 2-sec record length Assign coordinate information to data set Eliminate bad traces, correct polarity reversals Adjust amplitudes using 400 msec root mean square gain window 10–20–80–100 Hz Zero all data amplitude before and including first arrivals Zero amplitudes of residual surface waves Correct travel times for variation in station elevation Reorder data by common midpoint number L2-norm predictive, 200-msec operator length Velocity filter to remove coherent linear noise Correct for normal moveout (with best velocity function of four analyses); 100% stretch mute applied Surface consistent, based on maximum stack power; three solutions applied iteratively; 15-msec maximum static shift allowed 0%–30% eigenimage accepted Common-offset frequency–wavenumber dip moveout Zero-offset mean stack Steep-dip finite-difference time migration, using 70% smoothed stacking-velocity field 0%–30% eigenimage accepted Digital conversion using best smoothed stacking velocity function

with a proprietary software algorithm using the best-estimate smoothed two-dimensional stacking velocities. Typical vertical resolution of the final stacked section varied from 16 m at 200-m depth to 26 m at 1000-m depth, with a dominant frequency of 30 Hz through this range. Resolution below 1000 m is poor due to a high ambient noise level. Poor resolution is also evident in the interval velocity field (Fig. 3B) derived from the smoothed stacking velocities, where velocities below even 800 m are not well constrained by reflection velocity analysis. The interval velocity of 2800 m/sec is dashed to indicate it is the deepest interval reasonably estimated from the stacking velocities of these data.

ten comm. 2000). Two of the wells are located within a few hundred meters of the profile (wells LV3 and CR5; Fig. 1). Although caution is required when interpreting these borehole data (i.e., do the wells penetrate basement or a granitic clast?), they are interpreted to constrain the bedrock depth to the upper 250 m between stations 1100 and 1400. The seismic profile also passed by an exposed bedrock outcrop near station 1460 (La Loma Hills). We can trace a strong reflection from station 1150 to the line end at station 1470 that tie these three points together, and thus we have confidence in our pick of the bedrock surface through this region. The high amplitude and lateral coherency of this reflection is also indicative of a sediment–bedrock interface. Between stations 1050 and 1150, a strong reflection at 220- to 250-m depth is believed to be the northeastern extension of the bedrock surface, although the reflection is not as continuous as it is to the southwest (Fig. 4). If the reflection disruption is caused by a fault, then there is little apparent vertical displacement occurring across it because the reflection occurs at similar depths across the disrupted region. The Pelona schist, one of the rock types making up the basement-rock complex, is exposed in outcrop approximately 100 m east of the northeastern end of the seismic profile at Perris Hill (Fig. 1). Other hills within the valley are also composed of basement rocks of various types in outcrop, suggesting the basement composition and topography may be complex. Bedrock is first imaged on the seismic profile at about 100-m depth beneath the northeastern end of the profile (Fig. 3A, purple shaded region). The bedrock surface dips at about 13⬚ southwest for roughly 1.5 km before becoming subhorizontal for approximately 1.5 km, to station 420. This surface then apparently dips as much as 12⬚ southwest for about 4 km southwest of station 420 into the deepest part of the basin as inferred from gravity data (Anderson et al., 2000) between stations 825 and 1075. The basement surface is not well imaged in the seismic data between stations 650 and 1075. Between stations 650 and 850, a sequence of coherent but discontinuous reflection packages that get deeper to the southwest is picked as the basement surface. This correlation is further guided by analyzing shallower coherent reflections for possible faulting and projecting these faults to depth. Unfortunately, significant noise in the data, lack of energy from the seismic source, and/or structural complexity makes interpretation quite ambiguous between stations 900 and 1010. Based on a subhorizontal reflection sequence through this station range, we infer this interface is between 1.4 and 2 km depth, or roughly 1.7 km deep near the San Jacinto fault. Basin Sediment

Basement (Bedrock) Depth Although deep borehole information is sparse within the San Bernardino Valley, four water wells that encounter granitic bedrock constrain the interpreted depths near the southeastern end of the seismic profile (well information from W. Danskin, written comm. 2000, and L. Woolfenden, writ-

Several water wells near the seismic profile help constrain the interpretation of the sedimentary deposits in the upper 300 m over much of the profile length. Well data from Izbicki et al. (1998) showed that the middle water-bearing unit (MWB) within the basin tends to correlate on average with a 30%–40% increase in electrical resistivity (three of

2509

Figure 3. (Caption on next page)

2510

Figure 3. (A) Stacked seismic-reflection profile acquired across the San Bernardino Valley. The San Bernardino seismic-reflection profile covers over 14 km of the basin, from Perris Hill on the northeast to La Loma Hills on the southwest (see Fig. 1). The continuous profile crossed numerous busy streets through the city of San Bernardino. Several wells (vertical white bars on upper section) and outcrops on each end of the profile constrain the location of the sediment–bedrock interface. Bedrock (basement) is the purple shaded region. White dashed boxes outline areas shown in Figures 4 and 5. The sediment–basement interface is queried where interpretation is not well constrained. Shallow sediment region within the San Jacinto fault zone is also queried. The basin is asymmetric about its axis, with its northeastern flank dipping at about 10⬚ southwest. Numerous faults are interpreted along the seismic transect, including the San Jacinto, Loma Linda, possibly the Banning, and several unnamed faults (red lines, top section). The main strand of the San Jacinto fault zone is interpreted at the truncation of the bedrock and sedimentary reflections near station 1070. The basin sediments are divided into two seismic–stratigraphic units (gold region is younger sediment, orange brown region is older sediment) based on the strong, generally continuous reflection observed along the northeastern half of the profile. All of the sediments discussed here are undifferentiated Quaternary and Tertiary deposits. (B) Depth-converted final migrated stack overlain by smoothed interval velocity contours (in m/sec). These velocities were derived from the best-estimate stacking velocities. Dashed 2800 m/sec contour is the deepest interval velocity reasonably estimated from these data.

2511

Delineation of Faulting and Basin Geometry along a Seismic Reflection Transect in Urbanized San Bernardino Valley

Figure 4. Gray-scale image of seismic data from stations 1000 to 1300. Disruption of reflections, with little apparent vertical displacement, may be the location of the Rialto-Colton fault as it converges toward the San Jacinto fault to the south. Black lines represent interpreted faults.

Station 950

900

SW

Depth (m)

100

ult? a fa ind aL Lom

the well locations, 34G4, 34Q1, and 22L17, are shown in Fig. 1). Empirically, the seismic velocity of sediments, and thus impedance, is often proportional to resistivity (e.g., Rudman et al., 1975) (by this nonlinear relationship, a 30% increase in resistivity is roughly a 5% increase in velocity in the 100 ohm m range). It appears that the MWB correlates with the strong, continuous reflection observed over most of the northeastern side of the seismic profile, particularly in the regions between stations 200–440 and 640–810 (Fig. 3a). This unit has thus been interpreted as a seismic–stratigraphic boundary between shallow and deep basin sediments (in Fig. 3A, the orange brown shaded region represents sediments below this boundary). The velocity contrasts between the various deposits within the entire basin sediment package, even across this inferred seismic boundary, do not appear to be greater than 5%–10% based on reflection velocity analysis, which is less than what one might expect across the sediment–basement interface. By correlating the strong boundary reflection with available water well logs, the top of the deeper basin sediments is mapped from station 180 to at least station 1150. The deeper sediments generally increase in thickness southwest of station 150 toward the San Jacinto fault zone and accommodate faulting by upwarping, downwarping, and tilting. An apparent tilted block is imaged between stations 820 and 930 (Fig. 5). Reflection character within the block is not as continuous as it is to the northeast (station numbers less than 820). The strongest reflection, believed to be from the seismic– stratigraphic boundary and shown by the gray line with teeth in Figure 5, gets markedly shallower from stations 840 to 900. A borehole log from Izbicki et al. (1998) projects into the seismic profile at about station 940 and roughly corroborates this interpretation. Although the block is tilted at about 15⬚, there is no known clear evidence of deformation at the ground surface (e.g., Dutcher and Garrett, 1963). Thus, it appears that much of this block rotation occurred prior to deposition of the youngest basin deposits.

850

800

750 NE

200 300 400 500 600

200 m

Figure 5. Gray-scale image of seismic data from stations 740 to 950 showing interpreted faults and a tilted sedimentary block. Heavy gray lines with teeth represent seismic–stratigraphic contact. This block, tilted at about 15⬚ northeast, is probably caused by tectonic deformation between the San Jacinto and Loma Linda faults. The seismic–stratigraphic sedimentary boundary is not strongly imaged between stations 940 and 1070, in the projected vicinity of the San Jacinto fault (Fig. 3A). The boundary was thus picked primarily on water-well-log descriptions through this region rather than lateral correlation of the reflection. We speculate that the deeper seismic–stratigraphic unit terminates beneath station 1190. Overall, the shallower basin sediments are not imaged as continuously as the deeper deposits; however, this may be caused more by lower-fold coverage from the selected station interval than by a generally nonreflective nature of the sediments. Faulting/Basin Structure Faulting is evident at numerous locations along the seismic profile (Fig. 3A). However, only faults that we believe

2512

W. J. Stephenson, J. K. Odum, R. A. Williams, and M. L. Anderson

are the most prominent features are discussed here. On the southwestern end of the profile, we interpret a fault disrupting bedrock near station 1140 and infer that it is the southern extension of the Rialto-Colton fault (Fig. 4). This is based on the southern projection of the fault intersecting the seismic profile in the vicinity of the disrupted location. The region between I-10 and I-215 (approximately stations 950– 1050) has a very high cultural noise level, and it is unclear whether the region is nonreflective because of extensive faulting or poor signal quality. We believe it is caused by both and thus interpret this region from stations 950 to 1070 as the San Jacinto fault zone. This region may contain several additional fault strands within the sedimentary deposits that are not clearly imaged in the seismic data. The main strand of the San Jacinto fault zone is interpreted at the truncation of the bedrock and sedimentary reflections near station 1070. This location is more than 1 km southwest of the projection of the fault from its mapped location to the southeast in the Loma Linda Hills (Sharp, 1972), and places it west of the I-10 and I-215 interchange (Fig. 1). At least 1.2 km of apparent vertical displacement is inferred across the main strand of the San Jacinto fault zone. Several faults dipping at about 30⬚ to the southwest are evident within the tilted sedimentary block between stations 810 and 950 (Fig. 5). At the change in slope of the sedimentary reflections (about station 810), the Loma Linda fault, mapped south of the valley, projects across the seismic profile. If this change in slope is related to fault deformation, then these data may provide the first evidence of the Loma Linda fault continuing at depth into the San Bernardino Valley. This change in slope is also associated with 50 m of apparent vertical displacement at 300 m depth. Approximately 2 km to the northeast near station 625, another fault crosses the seismic profile that has over 70 m of apparent vertical displacement on the basement (because the sense of motion is different on the shallow horizon than at depth, significant out-of-plane motion must have occurred). Matti et al. (1985) suggested the western continuation of the Banning fault projects into the San Bernardino Valley (Fig. 1) in the general area of this imaged fault. We therefore speculate this structure may be related to the Banning fault beneath the San Bernardino Valley. Finally, another fault near station 450 accommodates apparent downwarping of the sedimentary deposits and basement rocks as well as about 30 m of apparent vertical displacement. Major deformation of the deeper basin sediments and the thickening of the stratigraphic section to the west begin across this fault.

Gravity and Aeromagnetic Data Anderson et al. (2000) documented data from 611 recently acquired gravity stations throughout the region in and around the San Bernardino Basin. The seismic profile traverses within a few hundred meters of 35 of the new stations (Fig. 6), indicating good overlap between the two data sets. Gravity data were processed using conventional routines

Figure 6. Map view of the seismic reflection profile, shown in gray in upper figure, overlain by 35 gravity stations acquired within several hundred meters of the seismic profile (black diamonds). Isostatic gravity values projected onto the seismic reflection profile are shown in the lower part of the figure. Every fifth point is correlated across the line of projection by a dashed vertical line. The basin gravity anomaly is over 30 mGal. These gravity data were gridded in conjunction with a large regional data set (Anderson et al., 2000), and the resulting data were used to extract two-dimensional gravity profiles for modeling.

outlined by Telford et al. (1990). Terrain corrections were carried out to 166.7 km, and isostatic corrections were made using an Airy-Heiskanen model of compensation. Isostatic gravity data for the nearest 35 stations projected onto the seismic profile in Figure 6 reveal that there is over 30 mGal variation along the transect. The entire newly acquired data set and previously acquired data were gridded (Anderson et al., 2000), and the resulting grids were used to extract twodimensional gravity profiles. Densities and magnetic susceptibility information for the potential-field modeling were derived from data of Anderson et al. (2000). Typical bulk densities for rock samples around the valley include 2520–2620 kg/m3 for Pelona schist, 2690–2850 kg/m3 for gneiss and tonalite, 2630–2790 kg/m3 for diorite, and 2440–2640 kg/m3 for Tertiary sandstone. Unconsolidated basin fill is assumed to have a bulk average density between 2000 and 2400 kg/m3, typical for Quaternary–Tertiary basin fill. Magnetic susceptibilities (SI units times 10ⳮ7) for these rock types range from 8 to 16 for Pelona schist, 16 to 190 for gneiss, 630 to 995 for tonalite, 72 to 613 for diorite, and 56 to 477 for Tertiary sandstone. Unconsolidated basin fill susceptibility is assumed to be zero. The gridded gravity data along three transects are generally parallel with the trend of the seismic profile and about 2.5 km apart (Fig. 7). These three two-dimensional transects

Delineation of Faulting and Basin Geometry along a Seismic Reflection Transect in Urbanized San Bernardino Valley

2513

Figure 7. Map view of isostatic gravity across San Bernardino Valley and vicinity. The data are shown with 1 mGal contour interval (from Anderson et al., 2000). The heavy black line is the seismic profile location and gray triangles are gravity station locations. Medium black lines are previously mapped faults as shown in Figure 1. Three gravity modeling transects were extracted from the gridded three-dimensional gravity data set along lines aa⬘, bb⬘, and cc⬘. The three gravity transects are generally parallel with the trend of the seismic profile and about 2.5 km apart. Profile bb⬘ approximately overlies much of the seismic profile location.

were then forward modeled in conjunction with aeromagnetic profiles along the same transects. The resulting models, to 2-km depth and to a few hundred meters northeast of the San Andreas fault, are shown in Figure 8. The aeromagnetic data provide a useful constraint on intrabedrock structure when used in conjunction with the gravity data during the two-dimensional modeling. These data further delimit the interpreted deformation and fault locations observed in the seismic data. Anderson et al. (written comm., 2001) presented a detailed analysis of these aeromagnetic data, as well as the complete gravity data set. Bedrock outcrops on opposite ends of the seismic transect serve as ties for the gravity and magnetic modeling, and

their variable composition suggests that basement gravity and magnetic properties should vary along the profile. Overall, the modeling results reflect this basement rock variability on each of the three transects, with each requiring denser bedrock material on the southwest side of the valley in the Peninsular Range block. Highly magnetic bodies are necessary to match the magnetic anomalies within the basin, even though their inclusion appears at first to be somewhat ad hoc. Possible error in the model basement depths can be up to several hundred meters over the deepest model regions. The best-fit model for each of the three profiles suggests an asymmetric basin, with the southwest flank dipping more steeply than the northeast flank (Fig. 8). On each of the pro-

2514

W. J. Stephenson, J. K. Odum, R. A. Williams, and M. L. Anderson

Figure 8. Three two-dimensional gravity and aeromagnetic forward-modeling results along transects shown in Figure 7. Densities and magnetic susceptibilities were derived from data of Anderson et al. (2000). SAF and SJF are San Andreas and San Jacinto faults, respectively. D, density, in kg/m3, of body polygons; S, magnetic susceptibility in SI units times 10ⳮ3. Basin sediments are modeled as a single body of 2120–2170 kg/m3. Vertical exaggeration is approximately 2:1. All three of the profiles indicate an asymmetric basin, with the southwest flank steeper than the northeast flank. On each of the profiles, a knob of bedrock protrudes toward the surface several kilometers northeast of the San Jacinto fault.

Delineation of Faulting and Basin Geometry along a Seismic Reflection Transect in Urbanized San Bernardino Valley

files, a knob of bedrock protrudes toward the surface several kilometers northeast of the San Jacinto fault. The knob appears to decrease in height along profile bb⬘ relative to the adjacent two profiles. Although the knob is reflected prominently in the magnetic data, it is only a subtle feature in the gravity, and there is a height uncertainty of Ⳳ100 m. Profile bb⬘ most closely overlies the seismic profile location, and these two profiles are directly compared in Figure 9. The heavy black line representing the best-guess sediment– bedrock interface on the seismic data, when overlain on the gravity profile, shows a close match between the two data sets. Although the seismic interface was not used to constrain the gravity and magnetic modeling beyond providing an initial first guess for the basin shape, the disagreement

Figure 9.

2515

between the model and interpretation is generally 200 m or less. The two results disagree most where the seismic interpretation is most weakly constrained, in the central part of the transect and northeast of the main San Jacinto fault. Disagreement between these results is in part not surprising because the transects are not exactly collocated (Fig. 7). Maximum depth is not well constrained on the seismic data, but the inferred depth nonetheless is consistent with the gravity result. The largest discrepancy occurs where the bedrock knob is modeled from the gravity and magnetic data (Fig. 9). The seismic data are very noisy at depth in this area, but it is interesting to note the upwarped sedimentary package between stations 850 and 940, southwest of the interpreted

Comparison of seismic reflection profile to gravity and aeromagnetic modeling profile bb⬘, shown in Figure 8. Bedrock is simplified into single body for comparison. Seismic interface between bedrock and basin sediment (heavy black line) is overlain on the gravity result. The two results disagree most where the seismic interpretation is weakest, in the central part of the transect and northeast of the San Jacinto fault; however, the disagreement is generally 200 m or less. Although depth is not well constrained on the seismic data, it is generally within 10% of the gravity result. The largest discrepancy occurs where the bedrock knob is modeled in the gravity/ aeromagnetic data. The seismic data are very noisy at depth in this area, but it is interesting to note that the upwarped sedimentary package southwest of the Loma Linda fault directly overlies the bedrock knob. It is possible this knob is an uplifted block of basement caused by faulting.

2516

W. J. Stephenson, J. K. Odum, R. A. Williams, and M. L. Anderson

Loma Linda fault, directly overlies the gravity knob. It is possible this knob is an uplifted block of basement formed by reverse or thrust faulting. Alternately, the knob may be a horst with normal faulting occurring on both flanks. In either case, the overlying sediments may be warped as result of the same faulting that created the basement block uplift.

Discussion Basin Geometry Hadley and Combs (1974) detected a high-velocity southwest-dipping layer from seismic refraction data at about 1.1-km depth near the intersection of the refraction profile and the inferred Loma Linda fault. Their velocity of 5.3 km/sec is indicative of metamorphic and igneous rocks of the basement complex. Results of Anderson et al. (2000) and of this study suggest the basin may be as deep as 1.7 km west of the refraction profile. The general southwest dip of the basement interface is consistent between these studies. Previous gravity studies suggested that the western part of the San Bernardino basin may be between 1 and 2.5 km deep (Willingham, 1968; Lambert, 1987), also consistent with this study. Given the low signal level in the seismic-reflection data over the deepest part of the basin, the gravity modeling results presented in this study are probably more accurate in this area. Both the gravity and seismic-reflection data reveal an asymmetric basin that deepens to the southwest. This differs from the more gently dipping and symmetric San Bernardino basin reference velocity model currently proposed by SCEC (e.g., Magistrale et al., 2000). The gravity contours in Figure 7 (from Anderson et al., 2000) suggest the basin geometry imaged in the seismic profile and modeled in the gravity or magnetic data may be extrapolated for 10–15 km northwest of and up to 10 km southeast of the profile location. Basin Ground-Motion Effects The two-dimensional finite-difference SH code of Frankel and Clayton (1986) was used to estimate the effect of the imaged asymmetric basin geometry on ground motion. Three separate simulations were performed: one using only the seismic interpretation to estimate basement (model I), the second incorporating the gravity-derived basement knob (Fig. 8) merged with the seismic result (model II), and the third using a version of the SCEC basin reference model (model III). For models I and II, the two-dimensional basin density and S-wave velocity structures used in the finitedifference modeling were reasonably well constrained by the seismic-reflection/refraction data, gravity data, and surface geology. These models were comprised of three layers, whose overall shapes were derived primarily for the San Bernardino seismic-reflection profile (Fig. 10A). Geologically, the layers represent (1) younger Quaternary/Tertiary basin deposits; (2) older Quaternary/Tertiary sediments as-

sociated with the lower seismic–stratigraphic unit; and (3) basement rock of metamorphic or granitic composition. The velocity structure for the model III was developed in a manner similar to Graves and Wald (2000) from the SCEC reference model (Magistrale et al., 2000) (Fig. 10B). The S-wave velocities for models I and II were derived from P-wave reflection/refraction velocities and assumed Vp /Vs ratios for representative rock types in the model layers. These velocities ranged from 400 m/sec at the surface to 2500 m/sec at depth. Layer 1 had a 400–800 m/sec vertical velocity gradient included. The velocity model was extended beyond the ends of the seismic line to minimize model edge effects across the region of interest. Density was varied within the layers using values for representative materials that were consistent with the densities used in gravity modeling. S-wave velocities and densities for model III were derived from SCEC version 2.2 code. Attenuation was not incorporated in the simulations. Because the goal was to determine the general effects of the basin geometry, no attempt was made to model geologic variation below about 2-km depth. A vertically incident SH-plane wave at 3-km depth was used as the source in all simulations. Nominal peak frequency for the simulation was 10 Hz using a grid spacing of 2.5 m. The resulting displacement seismograms were differentiated to velocity and bandpass filtered from 0.05 to 1.0 Hz for comparison with ground motions in an approximate frequency band where interpreted basin amplification was previously observed (Graves and Wald, 2000). Because the seismograms for both simulations II and I are visually quite similar, only those for the model I are presented in Figure 10C. Peak horizontal velocity for these simulations indicate the basement knob derived from gravity modeling has very little effect on the relative ground motion across the basin (Fig. 10D). Amplification of over 3 in velocity is observed over the deepest part of the basin in both simulations, northeast of the San Jacinto fault. The moderate dip of the basin toward the San Jacinto fault does not seem to cause dramatic variation in amplification once a factor of about 2.5 has been attained, with the exception of the deepest basin between the Loma Linda and San Jacinto faults. Likewise the variation in the interface between layers 1 and 2 does not appear to have much effect on amplification in the 0.05- 1.0-Hz frequency band (intuitively, this shallow interface should have more effect at higher frequency). A slight decrease in amplification at traces 8 and 9 seems to be related to the sliver of layer 2 sediment overlying bedrock southwest of the San Jacinto fault. Generally, the amplifications from these models are consistent with the amplification of 2–4 for recordings in the basin from the 1999 Hector Mine earthquake on stations within a few hundred meters of the seismic profile (e.g., Graves and Wald, 2000). Amplification from the model III simulation is about a factor of 2 above the bedrock reference ground motion (as observed in the model I simulation), but it is generally con-

Delineation of Faulting and Basin Geometry along a Seismic Reflection Transect in Urbanized San Bernardino Valley

2517

Figure 10. Results of two-dimensional finite-difference simulations along seismic reflection transect and comparison with seismograms of Graves and Wald (2000). Filtered transverse-horizontal velocity seismograms recorded at stations ehos, 5339, and hldr from the 1999 Hector Mine earthquake are shown in E. Peak horizontal velocities (cm/sec) for these seismograms are shown by the station name in parentheses. Peak velocity amplification for three simulations along the seismic reflection transect are shown in D. Amplification curves are predicted peak values caused by basin geometry as interpreted from seismic data only (model I, solid black line), as modified by gravity/aeromagnetic modeling results (model II, dashed black line), and by basin derived from the SCEC version-2 reference model (model III, dashed gray line). For all simulations, the source is a vertically incident SH-plane wave. Only the synthetic seismograms for model I are shown in C, after bandpass filtering from 0.05 to 1 Hz. Model III (shown in B) has a relatively flat response in amplification of about 2 across the seismic profile location. The shape of the basin generally causes a factor of 2–4 amplification in the simulations of models I and II (shown in A), relative to bedrock response at trace 1.

stant across the part of basin underlying the seismic transect. Seismic energy from the Hector mine event did not enter the basin as a vertically incident plane wave because it occurred at 5-km depth and 100 km to the northeast (e.g., Dreger and Kaverina, 2000). Amplification factors will change with angle of incidence of the wave field. Nonetheless, the variability seen across the basin in the tangential horizontal component (Fig. 10E; from Graves and Wald, 2000) is more consistent with amplification observed in the simulations for models I and II than with the roughly constant amplification seen in the model III simulation. Basin Fault Structures The Rialto-Colton and Banning faults are two of the more speculative faults interpreted in Figure 3a. Dutcher and

Garrett (1963) inferred a groundwater obstruction from water well data southwest of the San Jacinto fault they referred to as the Rialto-Colton barrier. This barrier has been interpreted as the Rialto-Colton fault in more recent groundwater studies (Woolfenden and Kadhim, 1997) and gravity studies (Anderson et al., 2000) 2–10 km west of the San Jacinto fault zone. Based on structural restoration arguments, the Rialto-Colton fault was interpreted as a strand of the San Jacinto fault zone by Anderson et al. (2000). The RialtoColton fault apparently converges southeastward toward the San Jacinto fault. While there is no apparent vertical offset observed at the reflector disruption on the seismic profile (Fig. 4; stations 1130 to 1170), we infer by its location that it may be the southward extension of the Rialto Colton fault or some ancillary fault linking it to the San Jacinto fault.

2518 As an alternate interpretation, the main strand of the San Jacinto fault may not be on the southwestern edge of the interpreted fault zone but rather within the zone or on its northeastern edge, at station 950 (Fig. 3a). The fault truncating bedrock at station 1070 might then be the RialtoColton or another associated fault. With this interpretation, the deepest part of the basin would presumably fall between the San Jacinto and Rialto-Colton faults. However, this scenario seems less likely given the arguments of Anderson et al. (2000), who interpret only 2 km of right-lateral offset on the Rialto-Colton fault based on magnetic anomalies and only 600 m of vertical offset based on two-dimensional modeling. The gravity data also suggest the fault at station 1070 is the main San Jacinto fault because of the size of the gravity anomaly associated with it. Another possible scenario is that the Rialto-Colton fault merges onto the San Jacinto fault near or northwest of the seismic transect. Resolving this question will require additional seismic reflection data. Matti et al. (1985) inferred that the Banning fault continues westward past the Crafton Hills into the San Bernardino basin. A fault can be inferred, in fact, along this projection from the gravity data (e.g., Anderson et al., 2000) (Fig. 7). Faulting and deformation of basin sediments also exist in the seismic data within a few kilometers of its projection; however, whether it is due to the Banning fault cannot be determined definitively from a single seismic reflection profile. The Banning fault would have to cut through, or be associated with, significant faulting observed in the Crafton Hills (Fig. 1), which is certainly possible at depth. It is equally plausible that the deformation observed in the seismic data is related to secondary faulting associated with the San Jacinto fault. Park et al. (1995) investigated the San Jacinto basin, about 40 km southeast of the San Bernardino basin, for intrabasinal fault structures. This basin was formed by a dilatational right step of the San Jacinto fault zone. They present evidence for a possible flower structure and multiple faults within the graben associated with the San Jacinto fault. The tilted sedimentary blocks between stations 810 and 950 (Fig. 3, 5) are reminiscent of the tilted block adjacent to the interpreted flower structure of Park et al. (1995) and could represent a similar kind of structure in the San Bernardino basin. We believe the southwest-dipping bedrock surface interpreted across the San Bernardino basin is the top of a tilted block of crystalline basement that has rotated about a horizontal axis in response to extensional stress accommodation between the San Jacinto and San Andreas faults. Figure 11 is a hypothetical structural diagram of the San Bernardino Valley between the San Andreas and San Jacinto faults from the seismic profile location northward. The subsurface gray dashed line represents a theoretical surface within crystalline basement that indicates relative sense of vertical displacement believed to have occurred across the San Jacinto fault. The three hachured patterns all represent undifferentiated crystalline basement, but they are subdivided into the sepa-

W. J. Stephenson, J. K. Odum, R. A. Williams, and M. L. Anderson

rate major blocks based on their location relative to the San Jacinto and San Andreas faults. The peninsular ranges block and San Bernardino Mountain blocks bound the block underlying the San Bernardino basin. We speculate that the Loma Linda fault is a strand of the San Jacinto fault that has accommodated some compression as the valley block has tilted under general strikeslip motion. This seems a reasonable assumption, given the complexity of the San Jacinto fault zone as previously discussed by Morton and Matti (1993). The San Jacinto fault is mapped as a very linear feature through the hills south of the San Bernardino basin, where it is observed as a sequence of small surface ruptures with stepovers of typically no more than a few hundred meters (Sharp, 1972). We believe this overall structural pattern probably continues into the valley. The two other faults interpreted within the San Bernardino basin northeast of the Loma Linda and San Jacinto faults offset shallow basin sediments but do not appear to have accommodated as much displacement. Because of their location within the basin under the urban area, however, they may pose an equally significant hazard to the region.

Conclusions Seismic reflection data acquired through the urbanized San Bernardino Valley reveal that the basin geometry is asymmetric with its northeastern flank dipping at roughly 8⬚ to 13⬚ southwest and its southwestern flank bounded by the nearly vertical San Jacinto fault zone. This is in contrast to the current SCEC version 2 reference velocity model (Magistrale et al., 2000) that consists of a more gently dipping and symmetric sediment–basement boundary across the imaged part of the basin. The seismic-reflection data show evidence of numerous faults cutting through the Quaternary– Tertiary basin sediments and basement surface and possibly provide the first evidence of the Loma Linda fault and the Banning fault cutting into the valley. At least one additional previously unidentified fault within the basin has also been observed north of the Banning fault. The main strand of the San Jacinto fault is interpreted at a location more than 1 km southwest of its previously inferred position, and it appears to be the southwestern limit of an over 1-km-wide zone of faulting. The Rialto-Colton fault is also interpreted from these data southeast of the San Jacinto fault zone. Basin sediments are deformed between the San Jacinto and Loma Linda faults, with the strata at 200-m depth tilted at 15⬚ to the northeast. The combined gravity/aeromagnetic modeling and seismic interpretation indicate the basin near the profile may be as deep as 1.7 km. Over 1.2 km of apparent vertical displacement on the basement is observed across the San Jacinto fault in the gravity/magnetic and seismic-reflection data. Finite-difference modeling suggests the overall basin geometry can contribute to a 2- to 3.5-fold amplification of earthquake waves in the 0.05–1.0 Hz frequency band, consistent with recorded earthquake data within the basin.

Delineation of Faulting and Basin Geometry along a Seismic Reflection Transect in Urbanized San Bernardino Valley

2519

Figure 11. Hypothetical block diagram showing how structural features delineated in seismic and gravity/aeromagnetic data might relate to surface features and at depth. The three hachured patterns represent undifferentiated basement in three blocks: the San Bernardino Mountain block, the Peninsular Ranges block, and the San Bernardino Valley block. SJF, LLF, BF, and SAF are San Jacinto, Loma Linda, Banning, and San Andreas faults, respectively. Heavy black lines represent hypothetical faults, dashed where inferred on surface. Subsurface solid gray line is sediment–bedrock boundary from seismic and gravity analyses. Subsurface dashed gray line represents theoretical surface within the crystalline basement that indicates a relative sense of vertical displacement across the San Jacinto fault.

Acknowledgments We greatly appreciate the efforts of Mike Seal, John LaRose, and Chris Smith of the County of San Bernardino Flood Control District, who made county property permitting and access go smoothly, and who allowed us to use County facilities. Deb Underwood helped greatly in permitting the southern portion of the seismic profile. We thank Richard Dart, Steve Harmsen, Carlos Mendoza, Nenna Okpara, and Susan Rhea for all their hard work during field acquisition. We especially thank our field observer David Worley for keeping the seismic acquisition equipment functional during some difficult moments. Rob Huggins and Craig Lippus of Geometrics Inc. assisted us immensely with our recording equipment. We very much appreciate the help of Elmo Christensen and O. B. Velez of Industrial Vehicles during field acquisition. Rufus Catchings generously lent acquisition equipment. The gravity models developed in this study were produced with GMSYS, interactive software produced by Northwest Geophysical Associates. Thanks to Wes Danskin, Linda Woolfenden, Jon Matti, Don Hough, and Gene McMeans for their assistance in obtaining well information. Discussions with Rob Graves and Art Frankel greatly improved the manuscript. The manuscript was also greatly improved through reviews by Vickie Langenheim, Harold Magistrale, Bill Savage, and an anonymous reviewer. Thanks to Lucy Jones for supplemental NEHRP funding, Wes Danskin for supplemental funding from the U.S. Geological Survey Water Resource Division, and Jon Matti and Doug Morton for additional funding from the Southern California Areal Mapping Project (SCAMP). This work was primarily supported by the U.S. Geological Survey National Earthquake Hazard Reduction Program. Use of brand names is for descriptive purposes only and does not represent a product endorsement by the U.S. Geological Survey.

References Alex, C. M., and K. B. Olsen (1998). Lens-effect in Santa Monica? Geophys. Res. Lett. 25, 3441–3444. Anderson, M., C. W. Roberts, and R. C. Jachens (2000). Principal facts for gravity stations in the vicinity of San Bernardino, Southern California, U.S. Geol. Surv. Open-File Rept. 00-193, 32 pp. Doll, W. E., R. D. Miller, and J. Xia (1998). A noninvasive shallow seismic source comparison of the Oak Ridge Reservation, Tennessee, Geophysics 63, no. 4, 1318–1331. Dreger, D. S., and A. Kaverina (2000). Seismic remote sensing for the earthquake source process and near-source strong shaking; a case

study of the October 16, 1999 Hector Mine earthquake, Geophys. Res. Lett. 27, no. 13, 1941–1944. Dutcher, L. C., and A. A. Garrett (1963). Geologic and hydrologic features of the San Bernardino area, California, U.S. Geol. Surv. Water-Supply Pap. 1419, 114 pp. Eckis, R. (1934). Geology and ground-water storage capacity of valley fill, south coastal basin investigation, California Dept. Public Works, Water Resources Bull. 45, 273 pp. Frankel, A. F. (1993). Three-dimensional simulations of ground motions in the San Bernardino Valley, California, for hypothetical earthquakes on the San Andreas fault, Bull. Seism. Soc. Am. 83, no. 4, 1020–1041. Frankel A. F., and R. W. Clayton (1986). Finite difference simulations of seismic scattering: implications for the propagation of short-period seismic waves in the crust and models of crustal heterogeneity, J. Geophys. Res. 91, 6465–6489. Graves, R. W., and D. J. Wald (2000). Observed basin response and 3D ground motion simulations for the 1999 Hector Mine earthquake, EOS 81, no. 48, F827. Graves, R. W., A. Pitarka, and P. G. Somerville (1998). Ground-motion amplification in the Santa Monica area: effects of shallow basin-edge structure, Bull. Seism. Soc. Am. 88, 1224–1242. Hadley, D., and J. Combs (1974). Microearthquake distribution and mechanisms of faulting in the Fontana-San Bernardino area of southern California, Bull. Seism. Soc. Am. 64, no. 5, 1477–1499. Izbicki, J. A., W. R. Danskin, and G. O. Mendez (1998). Chemistry and isotopic composition of ground water along a section near the Newmark area, San Bernardino County, California, U.S. Geol. Surv. Water Res. Inv. Rept. 97-4149, 27 pp. Lambert, D. W. (1987). A geophysical survey of a contaminated aquifer in Redlands, California, M. S. thesis, University of California, Riverside, 126 pp. Magistrale, H. S., S. Day, R. W. Clayton, and R. Graves (2000). The SCEC southern California reference 3D seismic velocity model version 2, Bull. Seism, Soc. Am. 90, no. 6B, S65–S76. Matti, J. C., D. M. Morton, and B. F. Cox (1985). Distribution and geologic relations of fault systems in the vicinity of the central Transverse Ranges, southern California, U.S. Geol. Surv. Open-File Rept. 85365, 27 pp. scale 1:250,000. Morton, D. M., and J. C. Matti (1993). Extension and contraction within an evolving divergent strike-slip fault complex: the San Andreas and San Jacinto fault zones at their convergence in southern California, Geol. Soc. Am. Memoir 178, 217–230.

2520 Park, S. K., D. Pendergraft, W. J. Stephenson, K. M. Shedlock, and T. C. Lee (1995). Delineation of intrabasin structure in a dilational jog of the San Jacinto fault zone, Southern California, J. Geophys. Res. 100, no. B1, 691–702. Pitarka, A., and K. Irikura (1996). Basin structure effects on long-period strong motions in the San Fernando Valley and the Los Angeles Basin from the 1994 Northridge earthquake and an aftershock, Bull. Seism. Soc. Am. 86, no. 1, (part B Suppl.), 126–137. Pratt, T. L., J. F. Dolan, J. K. Odum, W. J. Stephenson, R. A. Williams, and M. E. Templeton (1998). Multi-scale seismic imaging of active fault zones for hazard assessment: a case study of the Santa Monica fault zone, Los Angeles, California, Geophysics 63, no. 2, 479–489. Rudman, A. J., J. F. Whaley, R. F. Blakely, and M. E. Biggs (1975). Transformation of resistivity to pseudovelocity logs, Am. Assoc. Pet. Geol. Bull. 59, 1151–1165. Sharp, R. V. (1972). Map showing recently active breaks along the San Jacinto fault zone between the San Bernardino area and Borrego Valley, California, U.S. Geol. Surv. Misc. Geol. Inv. Map I-675, scale 1:24,000. Shaw, J. H., and P. M. Shearer (1999). An elusive blind-thrust fault beneath metropolitan Los Angeles, Science 283, 1516–1518. Stephenson, W. J., T. K. Rockwell, J. K. Odum, K. M. Shedlock, and D. A. Okaya (1995). Seismic reflection and geomorphic characterization of the onshore Palos Verdes fault zone, Los Angeles, California, Bull. Seism. Soc. Am. 85, no. 3, 943–950.

W. J. Stephenson, J. K. Odum, R. A. Williams, and M. L. Anderson Telford, W. M., L. P. Geldart, and R. E. Sheriff (1990). Applied Geophysics, Cambridge University Press, New York, 11–15. Willingham, C. R. (1968). A gravity survey of the San Bernardino valley, southern California, M. A. thesis, University of California, Riverside, 96 pp. Working Group on California Earthquake Probabilities (WGCEP) (1988). Probabilities of large earthquakes occurring in California along the San Andreas fault, U.S. Geol. Surv. Open-File Rept. 88-398, 62 pp. Woolfenden, L. R. and D. Kadhim (1997). Geohydrology and water chemistry in the Rialto-Colton Basin, San Bernardino County, California, U.S. Geol. Surv. Water-Res. Inv. Rep. 97-4012, 101 pp. U.S. Geological Survey P.O. Box 25046, MS 966 Denver, Colorado 80225 [email protected] (W.J.S., J.K.O., R.A.W.) Department of Geosciences University of Arizona Gould-Simpson Building Tucson, Arizona 85721-0077 (M.L.A.) Manuscript received 8 August 2001.