Earthquake Report: San Clemente Island

Well, yesterday was the start of a sequence of earthquakes offshore of San Clemente Island, about 100 km west of San Diego, California. The primary tectonic player in southern CA is the Pacific – North America plate boundary fault, the San Andreas (SAF).

    Here are the earthquakes in this sequence:

  • 2019.06.05 10:47:18 (UTC)M 4.3
  • 2019.06.05 14:32:09 (UTC)M 4.3
  • 2019.06.05 14:37:35 (UTC)M 4.3
  • 2019.06.05 16:13:43 (UTC)M 4.3
  • 2019.06.05 22:33:25 (UTC)M 3.3
  • 2019.06.06 01:44:33 (UTC)M 2.4
  • 2019.06.06 02:21:17 (UTC)M 2.3
  • 2019.06.06 11:18:09 (UTC)M 2.8
  • 2019.06.06 11:25:36 (UTC)M 3.5
  • 2019.06.06 17:19:10 (UTC)M 1.6

The region offshore where this ongoing sequence is called the California Continental Borderlands (CCB). There exists an excellent record of how the North America – Pacific plate margin boundary has evolved through time (remember, prior to about 29 million years ago, this plate boundary in southern CA was a subduction zone).

There was an earthquake offshore of Los Angeles last year. Check out my earthquake report and report update.

In places the SAF is a single thoroughgoing fault (e.g. in the southern San Joaquin Valley), in others it splays into multiple strands (in Orange County between the Santa Ana Mtns and Lake Elsinore), and in other places it bends to create regions of uplift (like in Ventura or the Santa Monica Mtns). The active faulting in the CCB is basically a series of right-lateral faults that step and bend to form uplifted islands and terraces, along with pull-apart sedimentary basins.

San Clemente Island is a region of uplifted non-marine Tertiary volcanic rocks (andesite and dacite) with ages ranging from 14.8 – 16.5 million years ago (Yeats, 1968; Merifield et al., 1971; Ward and Valenise, 1996). These rocks are overlain by Tertiary (Miocene) sediments (limestone, siltstone, shale, and diatomite; correlates to the Monterey Formation) and Plio-Pleistocene sediments (sandstones and conglomerates; correlates to the Fernando Formation found onshore; Stadum & Susuki, 1976; Ward and Valenise, 1996).

The bedrock is folded into a northwest trending anticline (rocks are folded upwards with the crest in the center of the island, forming a convex upward fold). Moore (1969) use regional compilations of seismic reflection data to show that this type of tectonic folding is ubiquitous throughout the CCB.

Ward and Velensise (1996) suggest that the San Clemente island formed via uplift during progressive slip on two, southeast striking, southwest dipping, blind thrust faults. These faults initiated movement between 3 and 5 Ma. There are a suite of Pleistocene marine terraces (2.56 Ma and younger) that provide evidence that uplift is continuing. Using fossil age determinations and correlation of marine terrace elevations with global eustatic sea level curves, the island is currently uplifting at rates between 0.2 and .5 mm/year. So, the underlying thrust faults are slipping at about 0.6-1.5 mm/yr (Ward and Velensise, 1996).

Muhs et al. (2014) used numerical ages (uranium-series analysis of corrals and amino acid geochronology of mollusks) to calculate marine terrace uplift rates in the CCB. When compared to uplift rates from different tectonic regimes, the terrace uplift rates in CCB is comparable to regions where strike-slip tectonics are dominant. These authors suggest that uplift like that found at the Big Bend (e.g. Ventura and Santa Monica Mtns) is not influencing terrace uplift rates in the CCB.

Along with this compression, there is a right-lateral (dextral) strike-slip fault on the east side of the island, the San Clemente fault, which has a slip rate of about 1 – 4 mm.yr (Ward and Valensise, 1996). The Southern California Earthquake Center suggests the slip rate is about 1.5 mm/yr for the SCF.

The ongoing sequence of earthquakes near the San Clemente Island are small in magnitude. If these were foreshocks to a larger earthquake, this would be felt across the southland, possibly cause damage on the island (where there is a U.S. Naval base), could possibly trigger submarine landslides or a small tsunami. Strike-slip earthquakes are not always considered a significant source for large tsunami, but there is abundant evidence that they do, though often much smaller than tsunami generated from thrust or subduction zone earthquakes. It is possible, if not probable, that this sequence will fizzle out.

Below is my interpretive poster for this earthquake

I plot the seismicity from the past month, with color representing depth and diameter representing magnitude (see legend). I include earthquake epicenters from 1918-2018 with magnitudes M ≥ 3.0 in one version.
I plot the USGS fault plane solutions (moment tensors in blue and focal mechanisms in orange), possibly in addition to some relevant historic earthquakes.

  • I placed a moment tensor / focal mechanism legend on the poster. There is more material from the USGS web sites about moment tensors and focal mechanisms (the beach ball symbols). Both moment tensors and focal mechanisms are solutions to seismologic data that reveal two possible interpretations for fault orientation and sense of motion. One must use other information, like the regional tectonics, to interpret which of the two possibilities is more likely.
  • I also include the shaking intensity contours on the map. These use the Modified Mercalli Intensity Scale (MMI; see the legend on the map). This is based upon a computer model estimate of ground motions, different from the “Did You Feel It?” estimate of ground motions that is actually based on real observations. The MMI is a qualitative measure of shaking intensity. More on the MMI scale can be found here and here. This is based upon a computer model estimate of ground motions, different from the “Did You Feel It?” estimate of ground motions that is actually based on real observations.

    I include some inset figures. Some of the same figures are located in different places on the larger scale map below.

  • In the lower right corner is a map from Wallace (1990) that shows the plate boundary and major faults overlain upon a topographic/bathymetric map. Check out the patterns made by the uplifted regions and the faulting (e.g. pull-apart basins). I placed a blue star in the general location of this ongoing sequence.
  • In the upper right corner there is a map that shows more detailed fault mapping and bathymetric contours (Chaytor et al., 2008).
  • In the upper left corner, there is a map from Legg et al. (2015) that shows how the strike-slip faults transect the CCB. Select earthquake mechanisms are shown (use legend at the top of the poster to help interpret these symbols) for some historic earthquakes. These authors collected and interpreted a number of seismic reflection profiles, including C-C.’
  • Below the Chaytor et al. (2008) map is seismic reflection profile C-C’ which shows how the basins are filled with sediment, the islands and terraces are also constructed of sedimentary rocks, and there are some steeply dipping faults. This profile is not travel time corrected, so depth is in two-way-travel-time (in seconds), not in depth. The faults probably dip more shallowly than is shown on the figure. The faults in this figure are aligned with the San Clemente fault system labeled on the map. Note that there are some faults that bound the Santa Nicolas Basin.
  • In the lower left corner is a figure that shows how a right-lateral strike-slip fault can create a geometry (e.g. in a step over) where there is extension that forms a pull-apart basin. This is one way to explain the formation of the Santa Cruz, Santa Nicolas, and Catalina basins shown on the maps.
  • Here is the map with a month’s seismicity plotted.

  • Here is the map with a century’s seismicity plotted.

  • Here is a map that shows detailed bathymetry data for the region (Dartnell et al., 2016, 2017) overlain on GEBCO coarser bathymetry data downloaded from GMRT. The land data are at 10 m resolution from The National Map (NED).
  • I plot USGS Quaternary Fault and Fold Database faults as faint white lines. Earthquakes include the past month for magnitudes M ≥ 0.5 and events since 1919 for M ≥ 4.0.
  • Look at the bathymetry surrounding the island. We can clearly see the SCF to the east of the island. There is evidence for a north-south striking fault to the west of the island. In the area just southeast of the earthquakes, there appears bedrock sticking up out of the continental shelf. This bedrock aligns with a ridge in the slop to the south of the island. This ridge may just be sediment, but it may also be tectonic in origin.

  • This map has the USGS MMI contours. The two M 4.3 temblors were felt across the southland.

  • Here is a larger scale map so that we can look at the bathymetry surrounding San Clemente Island in greater detail. I updated the USGS seismicity for 2019.06.06 at 20:00 Pacific time.

Other Report Pages

Some Relevant Discussion and Figures

  • Here is the figure showing the evolution of the SAF since its inception about 29 Ma. I include the USGS figure caption below as a blockquote.

  • EVOLUTION OF THE SAN ANDREAS FAULT.
    This series of block diagrams shows how the subduction zone along the west coast of North America transformed into the San Andreas Fault from 30 million years ago to the present. Starting at 30 million years ago, the westward- moving North American Plate began to override the spreading ridge between the Farallon Plate and the Pacific Plate. This action divided the Farallon Plate into two smaller plates, the northern Juan de Fuca Plate (JdFP) and the southern Cocos Plate (CP). By 20 million years ago, two triple junctions began to migrate north and south along the western margin of the West Coast. (Triple junctions are intersections between three tectonic plates; shown as red triangles in the diagrams.) The change in plate configuration as the North American Plate began to encounter the Pacific Plate resulted in the formation of the San Andreas Fault. The northern Mendicino Triple Junction (M) migrated through the San Francisco Bay region roughly 12 to 5 million years ago and is presently located off the coast of northern California, roughly midway between San Francisco (SF) and Seattle (S). The Mendicino Triple Junction represents the intersection of the North American, Pacific, and Juan de Fuca Plates. The southern Rivera Triple Junction (R) is presently located in the Pacific Ocean between Baja California (BC) and Manzanillo, Mexico (MZ). Evidence of the migration of the Mendicino Triple Junction northward through the San Francisco Bay region is preserved as a series of volcanic centers that grow progressively younger toward the north. Volcanic rocks in the Hollister region are roughly 12 million years old whereas the volcanic rocks in the Sonoma-Clear Lake region north of San Francisco Bay range from only few million to as little as 10,000 years old. Both of these volcanic areas and older volcanic rocks in the region are offset by the modern regional fault system. (Image modified after original illustration by Irwin, 1990 and Stoffer, 2006.)

  • Here is a map that shows the tectonic provides in this region (Legg et al. (2015). While the region inherits topography and geologic structures from past tectonic regimes, the dominant tectonic control here is currently the North America – Pacific plate boundary.

  • Map of the California Continental Borderland showing major tectonic features and moderate earthquake locations (M >5.5). The dashed box shows area of this study. The large arrows show relative plate motions for the Pacific-North America transform fault boundary (~N40° ± 2°W; RM2 and PA-1 [Plattner et al., 2007]). BP = Banning Pass, CH = Chino Hills, CP = Cajon Pass, LA = Los Angeles, PS = Palm Springs, V = Ventura, ESC = Santa Cruz Basin, ESCBZ = East Santa Cruz Basin fault zone, SCI = Santa Catalina Island, SCL = San Clemente Island, SMB = Santa Monica Basin, and SNI = San Nicolas Island. Base map from GeoMapApp/Global Multi-Resolution Topography (GMRT) [Ryan et al., 2009].

  • This map (Legg et al., 2007) shows an interpretation of the tectonics in this area. Note the location of the seismic reflection profile 116. San Clemente Island is on the southern edge of this map.

  • Shaded relief map of Santa Catalina Island and vicinity, where several restraining-bend pop-ups and releasing-bend basins exist along major fault zones. Epicentres for two moderate earthquakes (1981 Santa Barbara Island, M 6.0; 1986 Oceanside, M 5.8) and aftershocks bound the Santa Catalina Island restraining bend (locations by Astiz & Shearer 2000; focal mechanism from Corbett 1984). Other restraining-bend pop-ups include the Palos Verdes Hills (PVH) and Lasuen knoll along the Palos Verdes fault zone, and Signal Hill (SH) and possibly the San Joaquin Hills (SJH) along the Newport–Inglewood fault zone. Small pop-ups and pull-apart basins in the vicinity of Crespi knoll are shown in Figure 14. Total relief across the Catalina Fault is almost 2000 m, from Catalina Basin to Mt Orizaba. From 60 to 72 km of right-slip on San Clemente Fault is inferred from offset of Emery Knoll crater rim (Legg et al. 2004b).

  • Here is the USGS seismic reflection profile 116 (Legg et al., 2007). The San Clemente fault zone and the Catalina fault are shown. Check out the pull-apart basin.

  • Seismic-reflection profile USGS-116 across the Catalina basin (see Fig. 12 for profile location). Note the thin sediment cover over an irregular basement surface. A pull-apart basin exists where the San Clemente Fault steps to the NE to eventually merge with the Catalina Fault. The major faults have subvertical dips, typical of strike-slip faults. Convergence across the Catalina Fault has elevated Santa Catalina Island, and uplift occurs on both sides of the PDZ. Seismic data from USGS (J. Childs 2005, pers. comm.) FK migration at 4800 fps velocity was applied to 22-fold USGS stacked data.

  • Here is the figure with more details about the tectonic interpretation of the area (Legg et al., 2015)

  • Map showing bathymetry, Quaternary faults, and recent seismicity in the Outer Borderland. Fault locations are based on the high-resolution bathymetry, available high-resolution seismic reflection profiles, and published fault maps [cf. California Geological Survey (CGS), 2010]. The red symbols show magnitude-scaled (M>4) epicenters for seismicity recorded for the period of 1932 to 2013. Seismicity data and focal mechanisms are derived from the Southern California Seismograph Network catalogs, National Earthquake Information Center [2012–2013], and Legg [1980]. Focal mechanism event numbers correspond to Table S2 in the supporting information. The black rectangle shows location of Figure 10. The light blue lines show tracklines of multichannel seismic profiles—the labeled white profiles are shown in Figures 12 (124) and 13 (108 and 126).

  • Here is the summary figure from Legg et al. (2015). This helps us put these faults systems into context. Seismic reflection profiles from their publication are shown here (profile C-C’ is located in the rectangle labeled Fig 6 and plotted below).

  • Map showing major active tectonic elements of the northern part of the California Continental Borderland. Major active (Quaternary) faults are shown in red (SAF = San Andreas fault, ABF = Agua Blanca fault, SCF = San Clemente fault, and SCCR = Santa Cruz-Catalina Ridge, Ferrelo). Major strike-slip offsets are shown by shaded areas with estimated displacement (EK = Emery Knoll crater; Tanner Basin near DB = Dall Bank; and SDT = San Diego Trough, small pull-apart near Catalina). Other symbols show oblique fault character including transpressional restraining bends (CAT = Santa Catalina Island, CB = Cortes Bank, and TB = Tanner Bank), uplifts (SRI = Santa Rosa Island, SCz = Santa Cruz Island, SNI = San Nicolas Island, CB = Cortes Bank, TB = Tanner Bank, and SBM = San Bernardino Mountains), and transtensional pull-apart basins (SD = San Diego, ENS = Ensenada, SCB = San Clemente Basin, and SIB = San Isidro Basin). The large arrows show Pacific-North America relative plate motions with the blue dashed line (PAC-NAM) along a small circle for the RM2 [Minster and Jordan, 1978] plate motions model through San Clemente Island (SCL). Boundary between the Inner and Outer Borderland follows the East Santa Cruz Basin fault zone (dotted line; modified from Schindler [2010] and De Hoogh [2012]). Holocene volcanoes exist along the coast (SQ= San Quintín) and within the Gulf of California Rift (CP = Cerro Prieto and Obsidian Buttes, Salton Trough). Dates show year of earthquakes with mapped focal mechanisms (see Table S2 in the supporting information). SB = Santa Barbara, LA = Los Angeles, and PS = Palm Springs.

  • Here is the seismic reflection profile C-C’ shown on the poster (Legg et al., 2015).

  • High-resolution 24-channel 4 kJ sparker seismic profiles along the Santa Cruz-Catalina Ridge (C. Goldfinger, personal communications, 2001). The profiles are arranged from north to south along the southern part of the ridge (see Figure 6a for profile locations). Profile OSU-128 located near the bifurcation of the Santa Cruz-Catalina Ridge east of Santa Barbara Island. The colored horizons are correlated to stratigraphy mapped by Sorlien et al. [2013] (see Figure 6 for profile location). The East San Clemente fault zone merges with the Catalina Ridge fault zone within the eastern valley (Figure 5a). Reverse slip is evident on several steep faults that bound blocks of sedimentary rocks that were squeezed up in this transpressional fault system.

  • Here is a map from Maier et al. (2018) that shows how the faults are configured, as well as the sedimentary distribution systems (the focus of their paper). I grew up on the [concrete] banks of the San Gabriel River and this is where the submarine canyon and channels send their sediment loads.

  • Color-contoured slope-shaded multibeam bathymetry gridded at 10 meters. A) The Catalina Basin and the San Gabriel Canyon–Channel depositional system. Dashed line in the Catalina Basin indicates approximate extent of channels resolved on the seafloor.

  • Below are seismic reflection profiles plotted on the above map (Maier et al., 2018)

  • Northwest channels and lobes. See Figure 1A for profile locations. Gray lines represent profile intersections. A) Chirp profile across the northwestern Catalina Basin shows the stacking of lobes that do not reach the Kimki Fault (KF). B) An obliquely oriented chirp profile shows that the lobe deposits originate from the northwest channels, end before reaching the San Clemente Fault (SCF), and do not overlap in extent with lobe b.

  • This shows the timeline of what has controlled the tectonics in this region (Legg et al., 2015).

  • Chronology of major Cenozoic events in the Southern California region (after Wright [1991] and Legg and Kamerling [2012]). Intensity of tectonic deformation is represented by the curve. Local (Los Angeles Basin) biostratigraphic zonation is shown. The slanted labels for Neogene stages represent the time-transgressive nature of these boundaries.

Pleistocene Marine Terraces

    • Schematic cartoon illustrating the cutting and abandonment of marine terraces in an actively uplifting landscape in relation to sea level fluctuation. (a) Marine terrace cut during a relative sea level high stand. (b) Sea level drops and the marine terrace is uplifted. (c) During the next relative sea level high stand a new marine terrace is cut into the landscape below the older terrace. Modified after Nalin et al., (2007).

    • Here is a figure that shows the geomorphic features of a marine terrace (Wikipedia).

    • Here is a beautiful low angle oblique photo of the marine terraces on San Clemente Island (Yatsko, 2000). These authors studied the archaeological deposits on this island.

    • Emergent Pleistocene marine terraces on the west side of the island between Norton and Box canyons.

    • Here are some views of the terraces on San Clemente Island as photographed by Daniel Muhs (USGS).



    • Here is a map I prepared using the 2016 USGS Topobathy data (LiDAR and historic bathymetry mosaic).
    • I present these data as a shaded relief (hillshade) beneath an elevation raster with color representing height or depth. I also use a slopeshade raster to help highlight the changes in slope.
    • The 100 meter topographic contours are labeled. The inset shows the location of the main map in relation to the CCB with a pink polygon.

    • UPDATE: 2019.06.07
    • I prepared a couple maps that show the entire island. These are below, with 2 different color ramps.



    • Below is a fantastic summary showing the uplift rates for Pleistocene marine terraces along the North America – Pacific plate boundary system(Legg et al., 2015). Note the high uplift rates at the Big Bend and the Mendocino triple junction (another plate where there is a major change in SAF tectonics).

    • Map showing the plate tectonic setting of western North America (simplified from Drummond (1981) and Simkin et al. (2006)). SAF, San Andreas Fault; MTJ, Mendocino Triple Junction; CSZ, Cascadia subduction zone. Also shown are marine terrace localities with reliably dated ~120 ka, ~80 ka, or ~49 ka corals, or amino acid ratios in mollusks that permit correlation to ~120 ka, ~80 ka, or ~49 ka terrace localities, and elevation data that allow calculations of late Quaternary uplift rates. Paleo-sea levels, relative to present, used for uplift rate calculations are þ6 m (~120 ka), 11 m (~80 ka), and 62 m (~49 ka), derived from data in Muhs et al. (2012). Abbreviations and sources of data, south to north: CP, Cabo Pulmo (Muhs et al., 2002a); LP, La Paz (Sirkin et al., 1990); BH, Bahía Magdalena (Omura et al., 1979); IC, Isla Coronados and PC, Punta Chivato (Johnson et al., 2007; see also Table 2); MU, Mulege (Ashby et al., 1987); BT, Bahía de Tortugas (Emerson et al., 1981); PB, Punta Banda (Rockwell et al., 1989; Muhs et al., 2002a); PL, Point Loma (Kern, 1977; Muhs et al., 2002a); SCI, San Clemente Island (Muhs et al., 2002a, 2014); NB, Newport Bay (Grant et al., 1999); SNI, San Nicolas Island (Muhs et al., 2012); PV, Palos Verdes Hills (Muhs et al., 2006); NCI, Northern Channel Islands (this study); V, Ventura (Lajoie et al., 1979; Kennedy et al., 1982;Wehmiller, 1982); IV, Isla Vista (Gurrola et al., 2014; see also Table 2); SB, Shell Beach (Stein et al., 1991; Hanson et al., 1994); PSL, Point San Luis (Hanson et al., 1994; Muhs et al., 1994); C, Cayucos (Stein et al., 1991; Muhs et al., 2002a); AN, A~no Nuevo (Muhs et al., 2006); PA, Point Arena (Muhs et al., 2006); PD, Point Delgada (McLaughlin et al., 1983a, 1983b; Merritts and Bull, 1989); CC, Crescent City (Kennedy et al., 1982; Polenz and Kelsey, 1999); CB, Cape Blanco (Kelsey, 1990; Muhs et al., 1990); B, Bandon (McInelly and Kelsey, 1990; Muhs et al., 1990, 2006); YB, Yaquina Bay (Kennedy et al., 1982; Kelsey et al., 1996).

Geologic Fundamentals

  • For more on the graphical representation of moment tensors and focal mechanisms, check this IRIS video out:
  • Here is a fantastic infographic from Frisch et al. (2011). This figure shows some examples of earthquakes in different plate tectonic settings, and what their fault plane solutions are. There is a cross section showing these focal mechanisms for a thrust or reverse earthquake. The upper right corner includes my favorite figure of all time. This shows the first motion (up or down) for each of the four quadrants. This figure also shows how the amplitude of the seismic waves are greatest (generally) in the middle of the quadrant and decrease to zero at the nodal planes (the boundary of each quadrant).

  • Here is another way to look at these beach balls.
  • There are three types of earthquakes, strike-slip, compressional (reverse or thrust, depending upon the dip of the fault), and extensional (normal). Here is are some animations of these three types of earthquake faults. The following three animations are from IRIS.
  • Strike Slip:

    Compressional:

    Extensional:

  • This is an image from the USGS that shows how, when an oceanic plate moves over a hotspot, the volcanoes formed over the hotspot form a series of volcanoes that increase in age in the direction of plate motion. The presumption is that the hotspot is stable and stays in one location. Torsvik et al. (2017) use various methods to evaluate why this is a false presumption for the Hawaii Hotspot.

  • A cutaway view along the Hawaiian island chain showing the inferred mantle plume that has fed the Hawaiian hot spot on the overriding Pacific Plate. The geologic ages of the oldest volcano on each island (Ma = millions of years ago) are progressively older to the northwest, consistent with the hot spot model for the origin of the Hawaiian Ridge-Emperor Seamount Chain. (Modified from image of Joel E. Robinson, USGS, in “This Dynamic Planet” map of Simkin and others, 2006.)

  • Here is a map from Torsvik et al. (2017) that shows the age of volcanic rocks at different locations along the Hawaii-Emperor Seamount Chain.

  • Hawaiian-Emperor Chain. White dots are the locations of radiometrically dated seamounts, atolls and islands, based on compilations of Doubrovine et al. and O’Connor et al. Features encircled with larger white circles are discussed in the text and Fig. 2. Marine gravity anomaly map is from Sandwell and Smith.

  • Here is a great tweet that discusses the different parts of a seismogram and how the internal structures of the Earth help control seismic waves as they propagate in the Earth.

    Social Media

    References:

  • Chaytor, J.D., Goldfinger, C., Meiner, M.A., Huftile, G.J., Romsost, C.G., Legg, M.R., 2008. Measuring vertical tectonic motion at the intersection of the Santa Cruz–Catalina Ridge and Northern Channel Islands platform, California Continental Borderland, using submerged paleoshorelines in GSA Bulletin, v. 120, no. 7/8, p. 1053-1071, https://dx.doi.org/10.1130/B26316.1
  • Dartnell, P., Driscoll, N.W., Brothers, D., Conrad, J.E., Kluesner, J., Kent, G., and Andrews, B., 2015, Colored shaded-relief bathymetry, acoustic backscatter, and selected perspective views of the inner continental borderland, Southern California, U.S. Geological Survey Scientific Investigations Map 3324, 3 sheets, https://dx.doi.org/10.3133/sim3324.
  • Dartnell, P., Roland, E.C., Raineault, N.A., Castillo, C.M., Conrad, J.E., Kane, R.R., Brothers, D.S., Kluesner, J.W., Walton, M.A.L., 2017, Multibeam bathymetry and acoustic-backscatter data collected in 2016 in Catalina Basin, southern California and merged multibeam bathymetry datasets of the northern portion of the Southern California Continental Borderland: U.S. Geological Survey data release, https://doi.org/10.5066/F7DV1H3W.
  • Du, X., Hendy, I., Schimmelmann, 2018. A 9000-year flood history for Southern California: A revised stratigraphy of varved sediments in Santa Barbara Basin in Marine Geology, v. 397, p. 29-42, https://doi.org/10.1016/j.margeo.2017.11.014
  • Frisch, W., Meschede, M., Blakey, R., 2011. Plate Tectonics, Springer-Verlag, London, 213 pp.
  • Fuis, G.S., Ryberg, T., Godfrey, N.J., Okaya, D.A., Murphy, J.M., 2001. Crustal structure and tectonics from the Los Angeles basin to the Mojave Desert, southern California in Geology, v. 29, no. 1, p. 15-18
  • Hayes, G., 2018, Slab2 – A Comprehensive Subduction Zone Geometry Model: U.S. Geological Survey data release, https://doi.org/10.5066/F7PV6JNV.
  • Kreemer, C., G. Blewitt, E.C. Klein, 2014. A geodetic plate motion and Global Strain Rate Model in Geochemistry, Geophysics, Geosystems, v. 15, p. 3849-3889, https://doi.org/10.1002/2014GC005407.
  • Legg., <.R., Goldfinger, C., Kamerling, M.J., Chaytor, J.D., and Einstein, D.E., 2007. Morphology, structure and evolution of California Continental Borderland restraining bends in W. D. & Mann, P. (Eds) Tectonics of Strike-Slip Restraining And Releasing Bends. Geological Society, London, Special Publications, v. 290, p. 143–168
  • Legg, M. R., M. D. Kohler, N. Shintaku, and D. S. Weeraratne, 2015. Highresolution mapping of two large-scale transpressional fault zones in the California Continental Borderland: Santa Cruz-Catalina Ridge and Ferrelo faults, J. Geophys. Res. Earth Surf., 120, 915–942, doi:10.1002/2014JF003322.
  • Merifield, P.M., Lamar, D.L., and Stout, M.L., 1971. Geology of Central San Clemente Island, California in GSA Bulletin, v. 82, p. 1989-1994
  • Maier, K.L., Roland, E.C., Walton., A.L., Conrad,m J.E., Brothers, D.S., Bartnell, P., and Kleusner, J.W., 2018. The Tectonically Controlled San Gabriel Channel–Lobe Transition Zone, Catalina Basin, Southern California Borderland in Journal of Sedimentary Research, v. 88, p. 942-959, http://dx.doi.org/10.2110/jsr.2018.50
  • Meyer, B., Saltus, R., Chulliat, a., 2017. EMAG2: Earth Magnetic Anomaly Grid (2-arc-minute resolution) Version 3. National Centers for Environmental Information, NOAA. Model. https://doi.org/10.7289/V5H70CVX
  • Muhs, Daniel R.; Simmons, Kathleen R.; Schumann, R. Randall; Groves, Lindsey T.; DeVogel, Stephen B.; Minor, Scott A.; and Laurel, DeAnna, “Coastal tectonics on the eastern margin of the Pacific Rim: late Quaternary sea-level history and uplift rates, Channel Islands National Park, California, USA” (2014). USGS Staff — Published Research. 932.
    http://digitalcommons.unl.edu/usgsstaffpub/932
  • Müller, R.D., Sdrolias, M., Gaina, C. and Roest, W.R., 2008, Age spreading rates and spreading asymmetry of the world’s ocean crust in Geochemistry, Geophysics, Geosystems, 9, Q04006, https://doi.org/10.1029/2007GC001743
  • Nalin, R., Massari, F., and Zecchin, M., 2007, Superimposed Cycles of Composite Marine Terraces: The Example of Cutro Terrace (Calabria, Southern Italy): Journal of Sedimentary Research, v. 77, no. 4, p. 340-354.
  • Pinter, N., Lueddecke, S.B., Keller, E.A., Simmons, K.R., 1998. Late Quaternary slip on the Santa Cruz Island fault, California in GSA Bulletin, v. 110, no. 6, p. 711-722
  • Pinter, N., Johns, B., Little, B., Vestal, W.D., 2001. Fault-Related Folding in California’s Northern Channel Islands Documented by Rapid-Static GPS Positioning in GSA Today, May, 2001
  • Schindler, C.S., 2010. 3D Fault Geometry and Basin Evolution in the Northern Continental Borderland Offshore Southern California Catherine Sarah Schindler, B.S. A Thesis Submitted to the Department of Physics and Geology California State University Bakersfield In Partial Fulfillment for the Degree of Masters of Science in Geology
  • Shaw, J.H., Suppe, J., 1994. Active faulting and growth folding in the eastern Santa Barbara Channel, California in GSA Bulletin, v. 106, p. 607-626
  • Wallace, Robert E., ed., 1990, The San Andreas fault system, California: U.S. Geological Survey Professional Paper 1515, 283 p. [https://pubs.er.usgs.gov/publication/pp1515].
  • Yeats, R. S., 1968. Southern California structure, sea-floor spreading, and history of the Pacific Basin in Geol. Soc. America Bull., v. 79, p. 1693-1702

Return to the Earthquake Reports page.


Very interesting. I am unsure about the Ward and Valensise report as a reference for the origin of San Clemente Island – they comment that strike-slip hypotheses are unlikely, yet the recent multibeam data (2003 Navy survey) shows a very linear San Clemente Island fault that resembles the Carrizo Plain (your map is based on the USGS Catalina Basin survey, which seems to be missing a few other multibeam swaths in the area, but is still very impressive. Shepard and Emery, 1941, first suggested as much as 40-km of right-slip on the fault based on former juxtaposition of Fortymile Bank and San Clemente Island. The crater paper (Legg et al, 2004 Geophys Jour Int) suggests 60 km of right-slip based on offset rim of Emery Knoll crater – note that the west flank of the knoll has a major normal separation fault scarp associated with the transtensional section of the fault at the north end of the island. The south end is tranpressional with the “Carrizo Plain – Dragon’s Back Pressure Ridge” morphology.

Leave a Reply to Mark R. Legg Cancel reply

Your email address will not be published. Required fields are marked *