Geodetic strain across the San Andreas fault reflects elastic plate thickness variations (rather than fault slip rate)

Size: px
Start display at page:

Download "Geodetic strain across the San Andreas fault reflects elastic plate thickness variations (rather than fault slip rate)"

Transcription

1 Available online at Earth and Planetary Science Letters 269 (2008) Geodetic strain across the San Andreas fault reflects elastic plate thickness variations (rather than fault slip rate) Jean Chéry UNIVERSITE MONTPELLIER 2, Géosciences Montpellier, CNRS (ou CNRS/INSU), UMR 5243, Place Eugène Bataillon, CC Montpellier, France Received 11 January 2007; received in revised form 15 October 2007; accepted 27 January 2008 Available online 16 February 2008 Editor: C.P. Jaupart Abstract The interseismic velocity field provided by geodetic methods is generally interpreted in the framework of a thick elastic lithosphere with a slipping fault at depth. Because lateral variations of lithospheric rheology play a key role in determining the geological strain distribution, I examine the idea that interseismic strain rate variations also occur in response to lateral variations in the elastic thickness of the lithosphere. Using a stress balance principle and some simplifying assumptions, I show using a 1D model that elastic thickness is inversely proportional to strain rate for the simple case of pure strike-slip faulting. Elastic thickness computed on three profiles crossing the San Andreas fault system (SAFS) suggests that the distribution of interseismic strain rate is compatible with a thick elastic lithosphere in the Great Basin-Sierra Nevada province and on the Pacific plate. Conversely, a thin plate with a shallow asthenosphere is needed on the SAFS to explain its high strain rate. A 2.5D Finite Element model of interseismic strain in the Carrizo Plain region in Central California shows how known vertical and horizontal variations of elastic properties refine 1D model predictions. In a case of a multiple fault system, I point out that the interseismic velocity is not causally tied to faults slip rate. Therefore, analysing the velocity field across the SAFS cannot reliably provide faults slip rate distribution as previously claimed. Rather, the apparent correlation between geologic slip rate and interseismic strain may only indicate that the elastic thickness plays a dominant role in controlling fault strength. Finally, I suggest that interseismic geodetic strain could be a new way to infer effective elastic plate thickness on the continents Elsevier B.V. All rights reserved. Keywords: geodesy; interseismic strain; elastic thickness; GPS; lithosphere; rheology; fault; San Andreas fault; stress 1. Introduction Seismic hazard assessment strongly relies on the measurement of fault slip rates. For times longer than kyrs, repeated earthquakes offset the geomorphic features crossed by a fault (gullies and moraines). Offsets measurements and dated features determine an average fault slip rate (Sieh and Jahns, 1984). This geological approach is thought to be accurate due to its direct relation to fault slip observations. However, this method is not always straightforward as it requires unambiguously datable offset features. Another way to compute fault slip rate is to measure the interseismic strain by geodetic means such as the GPS technique. Due to the global GPS coverage, this method is potentially applicable worldwide on land and is address: jean@dstu.univ-montp2.fr. becoming a major tool to provide a global strain pattern at the Earth's surface (Kreemer et al., 2003). However, switching from interseismic strain to fault slip rate remains challenging. Because GPS observations are made on a small fraction of the seismic cycle, computing the fault slip results from a huge time extrapolation using a physical model of repeated earthquakes. Also, one must be confident that the surface strain are representative of the bulk deformation of the continental lithosphere, especially in the case of weak stratified plate. For strike-slip faulting, a widely used concept is based on a thick lithosphere with an embedded fault (Savage and Burford, 1973), referred in this paper as SB73 model or thick lithosphere model. During the interseismic phase, the fault is locked from the surface to a depth d (locking depth, see Fig. 1a). The fault below this depth slips at a constant rate s. This model corresponds to a X/$ - see front matter 2008 Elsevier B.V. All rights reserved. doi: /j.epsl

2 J. Chéry / Earth and Planetary Science Letters 269 (2008) Fig. 1. a) The thick lithosphere model in a vertical cross-section. The vertical fault is infinite perpendicular to the cross-section. During interseismic period, the fault above d is locked while the deep part slips at a rate s; b) adjustment of the GPS velocity data (solid dots) parallel to the SAF in the Carrizo segment in Central California (see Fig. 2 for precise location) using the screw dislocation model (solid curve and Eq. (1)). Values of 12 km and 34 mm/yr are used for d and s respectively. screw dislocation and the velocity variation at the surface is given by: v ¼ ðs=pþ arctan ðx=dþ ð1þ implying that interseismic velocity v reaches ~90% to the geological fault slip rate s when the distance to the fault x is larger than ~2πd. Such a formulation is highly attractive as an entire set of velocities may be fitted by adjusting the locking depth and the fault slip rate. Velocity profile across the central segment of the San Andreas fault (SAF) illustrates well this aspect (Fig. 1b). Using a slip rate of 34 mm/yr and a locking depth of 12 km leads to a RMS misfit between the SCEC 3.0 velocity field (Shen et al., 2003) and the model of 2.25 mm/yr, about twice as higher as the 1-σ formal data uncertainty of ~1 mm/yr (Schmalzle et al., 2006). It is remarkable that the geodetic slip rate matches well with the geological slip rate of 34 mm/yr (Sieh and Jahns, 1984; Brown, 1990) and that the locking depth corresponds to the maximum seismicity depth in this zone (Miller and Furlong, 1988). The arctangent shape of the velocity profiles across many large strikeslip faults encouraged scientists to extend the concept developed by Savage and Burford for a single strike-slip fault to multiple fault settings in which vertical faults delimit elastic blocks. Again, interseismic velocities are well fit by least square inversion for fault slip rate and locking depth. Based on its conceptual simplicity and on successful slip rate predictions of the SB73 model, its extension to the block model is about to become a routine tool to estimate fault slip rates in Northern California (Freymueller et al., 1999; Savage et al., 2004; D'Alessio et al., 2005) Southern California (Lisowski et al., 1991; Bennett et al., 1996; Becker et al., 2005; Fay et al., 2005; Meade and Hager, 2005) and other continental areas (McClusky et al., 2000; Wallace et al., 2004). To summarize, the use of the thick lithosphere model is mostly due 1) its conceptual simplicity 2) its ability to model the geodetic strain field 3) the determination of desirable parameters such as the long term fault slip rate. Because the thick lithosphere model provides a long term slip rate prediction, the link between short term and long term time scales is mandatory to assess this prediction reliability. Therefore, the motivation of this paper is to examine how this model is compatible with a long term mechanical model of lithospheric strain. Indeed, an interseismic strain model should be viewed in principle as a time fraction of a seismic cycle model. Also, a seismic cycle model spans another time fraction embedded in a long term evolution of the geological strain. Using the example of the SAFS, I start with describing the mechanical and rheological aspects of a long term model for parallel strike-slip faults. Then, I attempt to extract the model behaviour for the different phases of the seismic cycle (coseismic, postseismic, interseismic). This analysis leads me to propose that the interseismic strain pattern may be so much influenced by variations in lithosphere elastic thickness that it becomes difficult to obtain unambiguously an estimation of the slip rate on this sole basis. I test this hypothesis using a mechanical analysis of the geodetic interseismic strain on the SAF using a variable thickness mechanical model. 2. Geological strain, lithosphere rheology and the seismic cycle on the San Andreas fault Plate reconstruction and geological analysis have shown that most of the Quaternary strain between the Pacific plate and the Sierra Nevada is concentrated on a few faults spaced by a few tenths of km (Brown, 1990). In zones where the SAF orientation is aligned with the Pacific plate motion, these faults are nearly purely strike-slip. Therefore, the sum of the geological slip rate of these faults is believed to be (at 10% uncertainty) equal to the Sierra Nevada Pacific plate differential motion (Dixon et al., 2000). In this context, two kinds of active fault settings occur (Fig. 2). First, the slip rate can be distributed over a few faults. In northern California at the latitude N, the slip occurs from west to east on the San Andreas fault (20 25 mm/yr) the Maacama Rodger Creek faults (6 10 mm/yr) and the Bartlett Spring Green Valley fault (~5 mm/yr). A similar situation occurs in southern California at a latitude of 33 N but with an inverse distribution with respect to the northern California setting. Indeed, the slip rate is small on the Elsinore fault on the coast (~3 mm/yr),

3 354 J. Chéry / Earth and Planetary Science Letters 269 (2008) Fig. 2. Surface velocity (blue arrows) across the SAF system given by permanent and campaign mode GPS data (USGS and SCEC public data) in a North American reference frame. This velocity field (black arrows) mostly represents interseismic strain accumulation. Major faults are given in red (SAF = San Andreas fault; RC = Rodgers Creek fault; GV = Green Valley fault; SJ = San Jacinto fault; ELS = Elsinore fault). The three profiles across the fault system are drawn with a black line. The surface velocity field used in the computations and in Figs. 6, 7 and 8 are marked with blue dots (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.). while it is larger on the San Jacinto fault (8 mm/yr) and maximum on the SAF to the east (22 mm/yr). In contrast, in central California between the San Francisco Bay and the Big bend, most of the strain occurs on the SAF only at a rate of 34 mm/yr. Geophysical evidence may explain this pronounced strain localization in Central California. First, a high heat flow of ~80 mw/m 2 has been measured in a zone of 100 km around the SAF (Lachenbruch and Sass, 1980; Williams et al., 2004), suggesting high temperature in the lower crust and the uppermost mantle induces low viscosities. Also, stress measurements in deep boreholes indicate that the SAF supports a shear stress as low as MPa in the seismogenic zone (Zoback et al., 1987; Rice, 1992) contrasting with high values of ~ 150 MPa expected from laboratory measurements of fault friction (Byerlee, 1967). The combination of high heat flow in a 100 km wide zone around the SAF and the low resolved shear stress on the fault itself provides a physical explanation for slip localization on the SAF in a thermally weakened lithosphere (Furlong, 1993). Furthermore, the low compliance of the lithosphere seems to occur at a smaller scale. For example, magnetotelluric experiments at Parkfield show a low resistivity area beneath the seismogenic part of the SAF (Unsworth et al., 1997), probably due to a high fluid concentration related to intense shear at depth. Geomechanical modelling supports this view in requiring an effective friction coefficient of for the Central SAF (Bird and Kong, 1994) (Chéry et al., 2001). In zones of multiple parallel faults such as northern and southern California, the measured fault slip rates can be explained by a combination of lateral heat flow variations with low effective friction coefficients on faults (Provost and Chéry, 2006). In general, the contribution of the basal stress in mechanical modelling is neglected, although some authors have provided arguments that it may significantly affect the stress balance of the SAFS (Lachenbruch and Sass, 1973). Geological strain across active strike-slip faults results from a combination of the rheological stratification of the lithosphere versus depth and low resisting stress in fault zones (Gilbert et al., 1994). Depth stratification is mainly temperature dependent, and the depth of the 350 C isotherm represents the transition between frictional faulting and thermally activated viscous strain (Sibson, 1982). Another transition occurs at the crust mantle transition that marks a strength increase due to olivine rheology (Brace and Kohlstedt, 1980). As both crustal and mantle viscous laws are temperature dependent, an easy way to estimate differential stress variation with depth is to assume a constant strain rate through the Fig. 3. Typical stress envelops with depth for a) cold continental lithosphere, b) hot continental lithosphere c) weak fault zone embedded in a hot continental lithosphere (the dashed curve represents the frictional stress associated to a high friction as in a) and b). τ represents the magnitude of deviatoric stress and : e ¼ ct means that the strain rate does not vary with depth. Horizontal and vertical scales are indicative. The crust mantle transition occurs at ~30 km depth.

4 J. Chéry / Earth and Planetary Science Letters 269 (2008) lithosphere for different geotherms. Although the constant strain rate assumption is generally incorrect in actively deforming zones as demonstrated by mechanical modelling (Chéry et al., 2001), three end member models emerge for the lithospheric strength (Fig. 3). At low heat flow (40 mw/m 2 ), the 350 C isotherm is close to Moho depth, causing the crust to be mostly brittle (Fig. 3a). Low temperature in the subcrustal mantle should imply very high viscosities (Strehlau and Meissner, 1987). However, stress controlled plastic behaviour is likely to limit the maximum sustainable stress to about 600 MPa (Tsenn and Carter, 1987). At high heat flow (80 mw/m 2 ), the 350 C isotherm is shallow (7 15 km) and the Moho temperature is high (N700 C), leading to a strength profile mostly controlled by upper crustal friction and middle crust viscosity (Fig. 3b). However, this model has to be modified for a fault zone, with the constrain of low effective friction within the seismogenic layer (Zoback et al., 1987; Wang et al., 1995; Hassani et al., 1997)(Fig. 3c). In addition, strain rate weakening and metamorphic reactions may alter the deformation processes in the middle crust, and some authors have proposed that a drastic strength reduction occurs at the brittle ductile transition (Gueydan et al., 2001). Lithospheric strength profiles are useful to study the link between geophysical variables such as P and T and the effective rheology of the lithosphere. However, they represent the maximum sustainable stress (yield stress) of the lithosphere for a given strain rate, not the actual lithospheric stress. As an example, I consider the behaviour of the northern SAF as modelled by Provost and Chéry (Fig. 4a,b). In their study, the authors account with both strike-slip and shortening between the Pacific plate and the Sierra Nevada. The small shortening strain, which is accommodated by plastic strain inside the crust and also by dip-slip fault motion, is not considered in the present paper to keep a simple mechanical analysis. In the case of pure strike-slip motion, the stress magnitude along faults depends on the effective friction coefficient of the seismogenic zone and the viscosity of the middle crust. Because both frictional and viscous parameters in fault zones have low values, this limits the stress of the surrounding lithosphere that cannot Fig. 4. a) Strain rate invariant associated with parallel strike-slip faults of northern California (adapted from Provost and Chéry, 2006); major faults are modelled with intrinsically weak material and display high strain rate (Maa = Maacama fault; BS = Bartlett Springs fault); b) long term velocity in the fault-parallel direction at the surface (solid line) and at 25 km depth (dashed line); c) fault-parallel stress profiles for the fault zone located on the SAF and for the crust. The profile to the left corresponds to the maximum fault strength for this slip rate. The profile to the right inside the crust remains much below the maximum sustainable stress corresponding to the dashed curve. Stress integrals with depth must be equal on the two profile to ensure stress equilibrium.

5 356 J. Chéry / Earth and Planetary Science Letters 269 (2008) reach its maximum value. Therefore, the lithosphere does not deform and the viscous stress is zero due to the lack of strain. In this case, the lithospheric stress of this zone corresponds to elastic strain accumulation without involving frictional processes (right profile on Fig. 4c). Let us now consider the behaviour of such a long term model during the seismic cycle. In contrast with the SB73 model which is driven directly by imposing fault motion on the fault plane, this long term model is kinematically driven from the sides of the block at the prescribed side plate velocities. In this context, coseismic motion occurs in response to a step in effective fault friction, initially proposed by Brace and Byerlee (1966). Fault rupture during large continental earthquakes takes place between the surface and km depth and produces elastic strain in the fault vicinity. Postseismic motion occurs minutes to years following the earthquake as a result of a variety of processes like afterslip of the deep fault plane, viscoelastic processes in the middle crust and the mantle and poroelastic effects in the crust. If I ignore the poroelastic effect for this discussion, both afterslip and viscoelastic strain act to release stress buildup at depth and reload the fault plane above. Due to the thermal stratification of the lithosphere, a large viscosity spectrum [ Pa s] is likely to control the strain following large earthquakes as shown by postseismic modelling studies (Freed et al., 2006). Once this stress transfer is complete, a steady interseismic strain build up occurs in response to plate motion. A chief difference of this phase with respect to the long term behaviour is that the fault is locked, implying that the whole seismogenic layer behaves elastically or viscoelastically with viscosities higher than Pa s. At depth, a low viscosity stress occurs in response to interseismic strain beneath the seismogenic zone. If one attempts to interpret the interseismic strain across the SAFS with the model of Fig. 4c, it becomes clear that the surface strain has to be influenced by the thickness of the elastic and viscoelastic layers. As suggested by Fig. 5, thin parts of the layer should display strain accumulation while thick parts should not accumulate much strain. Although this two-layer model has been invoked for decades to explain postseismic strain (Nur and Mavko, 1974; Pollitz et al., 2000) (Kenner and Segall, 1999), it is less used to explain the interseismic strain distribution across fault systems (Bourne et al., 1998; Cohen and Darby, 2003; Schmalzle et al., 2006). In such a case, elastic strength is likely to depend on the product of the thickness of the seismogenic zone and the average shear modulus of this layer as detailed later. The viscoelastic strength determination follows the same scheme and depends on the product of the thickness of the viscous layer and its average viscosity. Two lines of evidence suggest that the viscous contribution to the lithospheric strength is relatively small. First, a large stress reduction with depth occurs in only a few km only due to the high sensitivity of power law flow to temperature. For example, a typical granite-type power law (Kirby, 1985) loaded at a strain rate of s 1 implies a deviatoric stress of 100 MPa at 350 C and only 1 10 MPa at 450 C. Assuming a surface thermal gradient of 20 C/km, this suggests that the thickness of the layer hosting significant viscous stress should not exceed 5 10 km. Second, effective viscosities determined in the crust and Fig. 5. Interseismic stress accumulation and corresponding velocity. a) Crosssection of a lithosphere with lateral variation along x-axis of its geodetic elastic thickness T g. The position of the 350 C isotherm is given by the dashed line and marks the upper limit of the viscoelastic zone. A low deviatoric stress is assumed to occur below corresponding to effective viscosities lower than Pa s. A fault-parallel stress increase Δτ xy occurring a time interval Δt (see Eq. (2)) is shown on two profiles. For each profile, the stress increase is assumed to be constant with depth accordingly and is materialized with a solid rectangle. Due to stress equilibrium the stress increase integrals on the two profiles are equal. b) Corresponding interseismic velocity field (y-axis component) across the lithosphere computed using Eq. (3). the mantle by postseismic modelling are on the order of Pa s, suggesting that these zones produce a stress contribution of ~1 MPa if an interseismic strain rate of s 1 is applied. To summarize this analysis, interseismic strain measured by geodetic tools at the Earth's surface may reflect the elastic strain accumulation of the seismogenic zone (i.e., grossly the part of the crust above 350 C and possibly the cold uppermost mantle) and the viscoelastic behaviour of a thin zone below the seismogenic layer in the crust and possibly in the mantle. To significantly contribute to interseismic stress accumulation, the corresponding viscosities must be higher than Pa s. This way to interpret interseismic strain has been proposed for the SAFS and for the Alpine fault of New Zealand (Bourne et al., 1998) but assuming that viscous strain in the lower crust and in the mantle drives and controls the entire fault system. Rather, I propose that interseismic strain reflects lateral variations in the thickness and elastic modulus of a lithospheric stress guide. This view is obviously close to the elastic plate model used to explain the flexural behaviour of the lithosphere (Watts, 2001). The main difference comes from the kind of forces applied to the plate. Classical plate theory aims to explain the relationship between vertical motion (plate bending) and plate thickness, while the concept I propose here considers the relation between horizontal strain and plate thickness. Because this thickness is evaluated by analysing horizontal geodetic strain, I name it geodetic elastic

6 J. Chéry / Earth and Planetary Science Letters 269 (2008) thickness (GET) or T g in order to differentiate it to the flexural elastic thickness (T e ) provided by plate bending analysis. 3. A simple model between strain rate and elastic thickness To build a simple model explaining the relation between the interseismic strain and the lithospheric strength, I assume here that the crust and the mantle are elastic for temperature lower than respectively 350 C and 750 C. Also, I neglect viscoelastic effects that is to say that the stress at temperature greater than 350 C and 750 C for the crust and the mantle respectively is negligible. The contribution of interseismic strain corresponds therefore to the deformation of a plate having an effective thickness T g. as presented in Fig. 5a. Because the mechanical system obeys stress equilibrium and assuming that only the stress component τ xy contributes to the total force, stress increase Δτ xy during a time Δt on two profiles of thickness T 1 g and T 2 g satisfies the relation: space derivative of the velocity field, I compute the strain rate using a least square adjustment of the interseismic velocity over a moving window of 20 km minimum half-width. Therefore, most of the long and short wavelength features of the profiles are preserved (Fig. 6a) and interseismic strain rate is given by the slope of the least square adjustment (Fig. 6b). The given formal error on the strain rate corresponds to the slope uncertainty associated to the least square adjustment. This leads to obtain a small error when all the points are aligned in the moving window, therefore corresponding to a constant strain rate model. A more complete formulation that should also include individual RMS associated to the data has not been developed here. The computation of the elastic thickness using Eq. (3) requires a priori information on T g because a trivial solution is given by T g =0, corresponding to the lack of information on the shear force change of Eq. (2). Because earthquake seismicity near the SAF does not occur below km, I assume a minimum Z T 1 g 0 Z T 2 Ds 1 xy ðþdz z ¼ g 0 Ds 2 xyðþdz z ¼ DF ð2þ where ΔF represents the shear force change applied to the lithosphere. I assume a linear shear stress shear strain relation given by Δτ xy =G Δε xy where G is the average shear modulus on the layer. Assuming that Δε xy is constant along a vertical profile and dividing the previous equation by Δt leads to: P Gx ðþt g ðþ x e : xy ðþ¼c x where P G is the average shear modulus on the profile and C is a constant, meaning that fault-parallel strain rate : e xy is inversely proportional to the integrated elastic strength P G T g. With the simplifying assumptions above, knowledge of the interseismic strain rate should teach us how elastic thickness may vary. I test this hypothesis on three profiles across the SAFS. 4. Analysis of velocity profiles crossing the San Andreas fault ð3þ GPS profiles shown on Fig. 2 cross the SAFS north of San Francisco Bay (North Bay profile, see Fig. 6), in the Carrizo plain south of Parkfield (Carrizo profile, see Fig. 7) and close to the Salton Sea south of Los Angeles (Salton profile, see Fig. 8). Faults are straight around these three locations so the deformation is nearly two-dimensional. Also, no large earthquakes have affected the SAF since the 1906 San Francisco earthquake for the North Bay profile and the 1857 Ft. Tejon earthquake for the Carrizo profile. Also, there is no evidence that the 1857 rupture extended down into the Salton Trough region. Therefore, the postseismic signal on the three profiles is expected to be small compared to the interseismic strain. I use the public data available on USGS and SCEC web sites corresponding to GPS surface velocities with respect to the North American plate. Because the RMS associated to the data are of the order 1 mm/yr, the direct derivative of the linear velocity field interpolation with respect to the profile direction is meaningless and not of a practical use in our case. In order to obtain a smooth Fig. 6. Interpretation of the GPS velocity field on a profile perpendicular to the northern SAFS. a) Fault-parallel GPS interseismic velocity fields (see Fig. 2 for location). A continuous version of the velocity measurements is given by the red curve. The RMS (mm/yr) of the adjustment between the curve and the data is given by the blue solid line with the scale to the right; b) fault-parallel horizontal strain rate corresponding to the continuous velocity field slope c) elastic thickness across the SAFS based on Eq. (3) (black curve) and on the assumptions given in the text. Minimum and maximum thickness associated to the strain rate formal error are represented by red lines (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.).

7 358 J. Chéry / Earth and Planetary Science Letters 269 (2008) elastic thickness on each profile of 13 km. Elastic thickness distribution is computed according to this value, also assuming that the average shear modulus does not vary along the profile. If computations were done with a different minimum elastic thickness, it would have been affected elastic thickness on the profile only by a multiplicative constant (see Eq. (3)). From north to south, interseismic strain rate patterns and their corresponding elastic thickness are markedly different. Along the North Bay profile, strain rate displays a marked asymmetry with a maximum along the Pacific plate to the west and a slow decrease towards the east up to the Central Valley and the Sierra Nevada. Consequently, the elastic thickness is higher close to the Pacific plate (40 70 km) and gently decreases (13 to 20 km) across the Coast Range from the SAF to the Green Valley fault. Because of the small strain rate of the Central Valley, the inverse relation between strain rate and thickness leads to an elastic thickness larger than 100 km (Fig. 6c). Interseismic velocity in central California leads to a different interseismic strain variation (Fig. 7b). Strain rate gently increases from southwest (the Pacific coast) to northeast up to a maximum 10 km east to the SAF trace as noticed by previous work (Schmalzle et al., 2006). Despite a limited GPS data set in the Central Valley and Sierra Nevada, the strain rate 20 km east to the SAF is virtually zero according to other geodetic studies of the Sierra Nevada block (Dixon et al., 2000). According to this strain Fig. 8. Same as for Fig. 6 for the southern SAFS (see Fig. 2 for location). variation, elastic thickness gradually decreases from values higher than 80 km to 13 km across the western Coast Ranges (Fig. 7c), and jumps back to large values east to the SAF in the Central Valley. The southern California profile near the Salton Sea reveals a reversed strain pattern compared to the one obtained in the North Bay area. Elastic thickness progressively decreases from the Pacific coast when crossing the Elsinore fault, reaches its minimum value between the San Jacinto and SAF and increases towards ~80 km on the north American plate to the east (Fig. 8c). 5. Finite Element model of the interseismic strain Fig. 7. Same as for Fig. 6 for the central SAFS (see Fig. 2 for location). The simple formulation used to compute the elastic thickness of the SAFS provides a first order relation between interseismic strain and plate thickness. However, stress components are likely to vary vertically within the plate, implying that an accurate strain solution requires solving a more complete stress balance equation. Also, the assumption of constant elastic modulus through the entire lithosphere is questionable, as vertical and lateral elastic properties variations of the lithosphere are known to be significant (Meissner, 1986). Involving these effects with solving stress balance equation with complex geometry needs to be done numerically. I design one experiment to determine if the simple concept developed above about the strain rate elastic thickness relation still holds for a finite thickness lithosphere. Because errors induced by the 1D model

8 J. Chéry / Earth and Planetary Science Letters 269 (2008) Fig. 9. a) Horizontal and vertical Young's modulus variation used in the Finite Element model; b) Geometry of the FEM and stress rate accumulation : s xy corresponding to a 34 mm/yr loading velocity between the lateral sides (at 0 and 350 km). Zone of low stress rate to the right corresponds to the Great Valley and the Sierra Nevada. Zone 20 km east to the SAF has low elastic modulus based on seismological evidence and loads at a high stress rate of 0.01 MPa/yr. Note that this value predicts an interseismic stress loading of 2.5 MPa for a recurrence time of 250 yr, which is compatible with the average static stress drop for a large earthquake (Hanks, 1977). Zone west to the SAF corresponds to a progressive plate thickening increase. are likely to be higher when large geometrical variations occur, I use the example of the central SAF for which high thickness gradients are expected. The geometry of the model represents a cross-section of the elastic part of lithosphere perpendicular to the fault direction. The loading corresponds to a motion of the Pacific plate with respect to a fixed Sierra Nevada at a rate of 34 mm/yr parallel to the fault direction. Assuming no velocity variation in the direction parallel to the fault direction (y), non-zero stress components are τ xy and τ yz. Both horizontal and vertical bulk elastic variations are taken into account (Fig. 9a). I incorporate Young's modulus increase with depth as observed worldwide in continents (Meissner, 1986). Based on a seismic velocity model of the Parkfield area (Eberhart- Philips and Michael, 1993) I also account for a lateral decrease of Young's modulus in a 20 km width zone east of the SAF. I also adjust elastic thickness along the Carrizo profile. Elastic thickness of the Great Valley Sierra Nevada is set to 200 km in order to account for both its low geodetic deformation (Dixon et al., 2000) and its high flexural rigidity (Kennelly and Chase, 1989). Elastic thickness of the Pacific plate west to the SAF is set to 40 km to make it compatible with currently estimated elastic thickness (Watts, 2001). The interseismic profile velocity for the model of Fig. 9b displays a data-model RMS of 1.77 mm/yr (Fig. 10). Due to intrinsic velocity errors of about 1 mm/yr on the Carrizo profile, the smooth interseismic curve (1D model of Fig. 6) fits velocity data with a RMS of 1.34 mm/yr. The data fit of the FEM model is not as good as the one provided by the 1D model. This is understandable as the 1D strain model is obtained by direct smoothing of the discrete velocity data, therefore representing the best possible data fit. Despite its lower accuracy, I argue that the data adjustment provided by the FEM has a greater physical meaning than the 1D model presented before. This opinion is based on three factors. First, it corresponds to a finite thickness lithosphere, implying that the hypothesis of having : s xy constant Fig. 10. Interseismic velocity field provided by the Finite Element model (thick black line) compared with the smooth data fit (dashed thin line) and with the thick lithosphere model (purple curve). Discrete GPS velocity values across the central SAFS are given by the solid black circles.

9 360 J. Chéry / Earth and Planetary Science Letters 269 (2008) on a vertical profile is no longer needed. Clearly, this 1D assumption is violated when high thickness gradients are present as shown by Fig. 9b. Second, the FEM allows us to include our best a priori knowledge of the elastic properties of the lithosphere based on seismology. Interestingly, a reliable estimate of bulk elastic properties requires the thickness only to be adjusted. Third, interseismic strain provided by the FEM results from a trial-and-error procedure. It is likely that a suitable inversion technique based on a grid search for an optimal elastic thickness would lead to a fit as good as the one provided by the 1D smooth fit. In comparison to the 1D and the FEM data fit, the adjustment of fault slip rate and locking depth of the SB73 model leads to a data fit of 2.25 mm/yr. This is mostly because the asymmetry of the elastic profile resulting from the variable elastic thickness model fits the GPS data better compared to the symmetric solution provided by the SB73 model. However, the fit of the thick lithosphere model is quite acceptable given the small number of free parameters and may be still improved if elastic modulus contrast across the SAF is taken into account. 6. Discussion Choosing among different mechanical models of interseismic strain on the sole basis of geodetic data fitting has been shown to be meaningless because of the non-uniqueness of the problem even in the case of a single fault (Savage, 1990). In other words, a bad data fit can allow us to discard a model but a good data fit is not proof of a model's relevance. Therefore, in chasing among possible models I need to consider the problem from a broader point of view than the one of finding the best data. Rather, it requires consideration of each model's relevance from a geophysical, rheological and geodynamical point of view. Having in mind the current view of the thermomechanic state of the lithosphere beneath the SAF (Lachenbruch and Sass, 1980; Furlong, 1993; Chéry et al., 2001), the use of a thick elastic model for the SAFS seems at odds to this knowledge. For example, the northern SAFS is very juvenile as it results from the northward progression of the Mendocino triple junction (Furlong, 1984). There, the disappearance of the subducting plate is thought to have opened an asthenospheric window in contact with the upper plate. As a consequence, a sharp contrast is likely to exists between a thin lithosphere (10 20 km) surrounding the SAFS and a thick lithosphere ( km) to the east in the Great Valley and in the Sierra Nevada. In other words, an interpretation of the whole fault system using a thick elastic lithosphere is therefore unlikely because of the presence of a hot low viscosity zone in the upper mantle in the region of the slab window. By contrast, the variable elastic thickness model predict a low rigidity of the SAFS, which is compatible with a shallow low viscosity zone. As already noted (Le Pichon et al., 2005; Schmalzle et al., 2006), the data fit provided by the Savage and Burford model is limited by the asymmetrical character of the velocity field. In the simple case of a single vertical strike-slip fault, one expects a perfect symmetry with respect to the fault axis. This is not always the case, as shown by Le Pichon et al. for different strike-slip faults. A plausible explanation is that lateral variations of elastic properties of a thick lithosphere cause a symmetry break (Rybicki and Kasahara, 1977). Indeed, large lateral variations of bulk rigidity are likely to occur and could explain the asymmetry of the deformation. However, a bulk rigidity contrast of 10 that is required to explain the strain asymmetry around the Sumatra fault (Le Pichon et al., 2005) probably exceeds known variations of shear modulus within the crust. An alternative explanation for strain asymmetry is a variation of the elastic thickness in conjunction with a bulk rigidity contrast (Melbourne and Helmberger, 2001; Cohen and Darby, 2003; Schmalzle et al., 2006) Interseismic strain, elastic plate thickness and fault slip rate As shown by the analysis of the central SAF profiles, SB73 and variable thickness models both provide a good fit to the interseismic velocity field. In contrast, different behaviours can be expected between these models when fault slip rates are searched. In the case of the SB73 model, the long term fault slip rate is equal to the far-field velocity (the differential plate velocity). The long term slip rate for the variable thickness model is less straightforward to define because of the remote fault drive. As discussed in the introduction, the effective fault friction should be considered here. If the fault friction is low, the fault will slip at the differential plate velocity because the lithosphere does not deform anelastically. If the fault friction is high, two cases have to be considered. If one assumes that the lithosphere always behaves elastically (meaning that it can sustain arbitrarily high strain) the fault slip rate must still be equal to the differential plate velocity. However, both in-situ and laboratory rock strength estimates indicate that the differential stress of the seismogenic lithosphere is limited by a friction coefficient of [ ] (Townend and Zoback, 2000). Therefore, a high fault friction is likely to pervasively deform the lithosphere around, especially where the plate is the thinnest. For this case only, the fault slip rate is expected to be different to the differential plate velocity. Considering that large faults separating thick lithospheric plates are often thought to sustain low shear stress, the fault slip rate determined by the variable thickness model must therefore be in fair agreement with the best fit SB73 model Slip rate computation of a parallel fault system From a mechanical viewpoint (see Appendix for a discussion), the SB73 model mimics the mechanical behaviour of two thick lithospheric blocks separated by a low strength fault zone. Interestingly, most fault slip rates inferred from SB73 and block models analysis around continental plate boundaries are in good agreement with the corresponding geologic slip rates (Reilinger et al., 2006). Therefore, it is logical to conclude that the geodetic fault slip determination using the SB73 model is valid only if it corresponds to a setting of a weak fault between two strong plates. However, many other fault settings occur in nature as multiple fault system, faults embedded in wide orogens or high plateaus. To discuss this point in the framework of the SAFS, I attempt here to explain how the variable thickness model behaves when two faults or more are embedded (Fig. 11). This model is directly adapted from the 1-fault model of Fig. A1c but

10 J. Chéry / Earth and Planetary Science Letters 269 (2008) Fig. 11. Mechanical model of a variable thickness lithosphere model with two embedded fault. Each fault strength i is the sum of a frictional stress occurring on a width h f i and of a viscous stress occurring on a width h v i. adding another frictional discontinuity and a viscoelastic zone according to Fig. 4c. As shown by the previous mechanical modeling, frictional and viscous fault strength both control slip rate distribution in a multiple fault system (Bird and Baumgardner, 1984; Provost and Chéry, 2006). Using the example of Fig. 11 and using a 1D balance assumption similar of Eq. (2), it can easily be shown that the effective strength of both faults are equal. Also, the strength of the fault i is equal to the sum of the frictional strength F f i and of the viscous strength F v i. For each fault, the frictional strength is equal to the stress integral along the frictional thickness h f i. Assuming a frictional shear stress constant with depth leads the frictional strength to be equal to: F i f ¼ s i h i f The viscous strength is equal to the integral of the viscous stress along the viscous layer which has a thickness h v i. This integral depends on the average strain rate inside this layer around the fault plane and is also affected by the large viscosity variations that likely occur with depth. Assuming that the viscous strain occurs on an average width w, the viscous strength is therefore defined by the relation ð4þ This relation is represented on Fig. 12 for different values of h 1 s 1 in the case or h 2 Nh 1. As the respective contribution of frictional and viscous fault strength is still a matter of debate, this formulation is well suited as it does not require an a-priori choice. In order to see how southern and northern SAFS behave with respect to Eq. (8), let us assume that viscous to frictional ratio is the same for each fault. Thus the ratio between frictional thicknesses is equal to the ratio given by geodetic elastic thicknesses. Slip rate and frictional thickness ratios are given in Table 1 and plotted in Fig. 12. Using preferred values for slip rates, only pairs including the Elsinore fault display θ 1 s 1 values larger than 3. For other pairs, elastic thickness and slip rate ratios match curves computed with between 1/3 and 3, suggesting that viscous and frictional strengths may have an equal influence on the SAFS. In order to see how this parameter compares with an a priori calculation, I use the following set of parameter: a viscosity of Pa s; a fault zone width of 1 km; a ratio h i vh i f=0.25; a frictional shear stress of 10 MPa. In such a F i v ¼ gi w hi v si ð5þ where is the fault viscosity η i divided by w and s i the fault slip rate. The total fault strength is equal to F i ¼ s i 1 þ h i s i h i f ð6þ with h i ¼ gi w hi m s i h i f ð7þ Therefore, θ i represents the viscous component of the fault strength. Assuming for purposes of discussion that both θ i and τ i are equal for both faults, the slip rate ratio is given by: " # 1 h 1 s 1 s 2 1 h 1 f ¼ max 0; 1 þ s1 h 1 s 1 h 2 f ð8þ Fig. 12. Normalized relation between slip rate and mechanical thickness using Eq. (8) for different values of the viscosity parameter (red lines). Black circles give the relation between slip rate and thickness as computed in Table 1. Black lines represents slip rate uncertainties.

11 362 J. Chéry / Earth and Planetary Science Letters 269 (2008) Table 1 Geodetic elastic thickness (provided by this study) and known fault slip rates for northern (lines 1 to 3) and southern (lines 4 to 6) SAFS (Shen-Tu, Holt et al., 1999) Fault1 fault2 T g 1 (km) T g 2 (km) s 1 (mm/yr) s 2 (mm/yr) h f 1 /h f 2 s 2 /s 1 SAF RC (a) [12 30] 9 [7 10] [ ] SAF GV (b) [12 30] 6 [4 8] [ ] RC GV (c) [7 10] 6 [4 8] [ ] SAF SJ (d) [10 35] 12 [8 24] [ ] SAF ELS (e) [10 35] 6 [2 9] [ ] SJ ELS (f) [8 24] 6 [2 9] [ ] Numbers in columns 5, 6, 8 represent preferred values while brackets indicate min/max estimates. SAF = San Andreas fault; RC = Rodgers Creek fault; GV = Green Valley fault; SJ = San Jacinto fault; ELS = Elsinore fault. case, θ i = s. Using a velocity of 25 mm/yr for the SAF ( ms 1 ), is close to 0.2. Given the large uncertainty in rheological parameters such as the fault zone viscosity and width, the agreement between this a-priori computation and the value deduced from the elastic thickness and slip rate plot is encouraging. This analysis suggests that several geometrical and rheological parameters control fault strength and therefore geological fault slip rates. Frictional strength is controlled by the fault friction and the seismogenic thickness, while viscous strength is controlled by the viscosity and the size of the viscous domain. Any variation of these rheological parameters on a fault will cause the slip rate to vary as already demonstrated by some studies (Roy and Royden, 2000; Provost and Chéry, 2006). The present analysis suggests a complex link between geologic slip rate and interseismic strain if more than one fault are active. Despite that the SB73 model is based on a direct and causal relation between interseismic strain and fault slip rate, these two parameters may only be positively correlated in nature. In order to properly interpret their relation one must understand their causal relation. Indeed, fault slip rate chiefly depends on effective fault friction, fault zone viscosity and the mechanical thickness of all these process zones. Because a greater thickness induces a strength increase if other rheological parameters remains unchanged, such an increase in thickness causes both the slip rate and the interseismic strain to decrease. But a change of effective fault friction modifies fault slip rate without changing the interseismic strain. Therefore, the apparent correlation between geologic slip rate and interseismic strain may only indicate that the elastic thickness plays a dominant role in controlling fault strength A global relation between elastic thickness and interseismic strain? The goal of this paper is to study the interseismic strain rate distribution across the SAFS and its relation to lateral rigidity variations. If such a relation holds for this plate boundary, perhaps a similar relation can be expected for other deformation areas such as subduction zones, mountain belts, continental or oceanic rifts. If a relation between interseismic strain and elastic thickness is assumed, then plate boundaries that displays high strain rate concentration would need to be interpreted like thin elastic zones. Is this mechanically understandable? Current knowledge of the SAFS suggests that a thin elastic lithosphere indicates a locally weakened plate. This kind of weakness can be thermally induced, as high temperature gradients reduce the thickness of both brittle and viscous layers. This is likely to occur in mountain belts, rifts or extending plateaus as shown by geophysical measurements and mechanical modelling (Gaudemer et al., 1988; Buck 1991; Cattin et al., 2001). Also, a large contribution to the lithospheric weakness could be related to low strength faults that decouple (in a stress meaning) adjacent plates. Subduction faults and large intracontinental faults are probably a good example of this kind of weakness (Wang et al., 1995; Hassani et al., 1997; Cattin et al., 2001). Up to now, elastic plate thickness has been computed for continental and oceanic plates using the idea that vertical loads applied to the lithosphere produce vertical motions by plate bending (Watts, 2001). Using stress equilibrium of the plate and isostatic assumption, knowledge of the horizontal distribution of the loading function (topography, internal loads, mantle buoyancy, glaciations, etc) permits determination of the flexural rigidity that best explains topographic and gravimetric signals. Knowledge of the elastic parameters of the lithosphere permits conversion of flexural rigidity to equivalent plate thickness T e. However, one needs to be careful when plate thickness are obtained with different kinds of loads. Because the model assumes an elastic plate over an inviscid fluid, the determination of the flexural rigidity in nature is correctly done if the load is applied for a time scale long enough to allow a complete stress relaxation at depth. Short time scale loading does not allow such stress relaxation, therefore leading to larger values of equivalent plate thickness. For example, mechanical modelling of the rapid filling of the lake Mead in the Basin and Range leads to a plate thickness estimate of km (Kaufmann and Amelung, 1995) much higher than the 5 15 km thickness found using the surface topography as a loading function (Lowrie and Smith, 1995). On both oceanic and continental plates, a clear correlation is found between elastic thickness and the thermal state of the lithosphere if a long term loading is considered. The C isotherm matches the lower limit of the elastic plate in oceans, while no strong correlation has been found between effective elastic thickness and some particular isotherm for continental crust (Watts, 2001). The current interpretation for oceanic lithosphere is that most of the elastic stress is stored between the surface and the C isotherm. In the case where elastic stresses reach the frictional strength of the lithosphere (generally due to high plate curvature as it occurs in subduction zones), an elastoplastic plate bending model has to be used (Judge and Mc Nutt, 1991) to account for a reduced elastic thickness. The situation is more complex in a moderately cold continental lithosphere for which a large part of the strength is probably stored in the uppermost mantle. As I show using the example of the SAF, the use of horizontal GPS velocity gradients at the Earth's surface through a stress equilibrium principle is another way to estimate effective plate thickness using its transverse rigidity as a parameter to invert. However, such a use of GPS velocities relies on two key assumptions. First, the strain must really represent the interseismic stage during which faults are locked. The second assumption is that the plate strain is the result

12 J. Chéry / Earth and Planetary Science Letters 269 (2008) of stress equilibrium inside the plate without coupling with the mantle other than hydrostatic forces. For example, horizontal forces applied at the base of the elastic plate such as deep convection or slab traction can induce strain on the surface, therefore leading to an incorrect plate thickness estimate. Keeping in mind that the relation between interseismic strain and plate thickness may therefore breakdown, the applicability of this theory on continents is worldwide. Indeed, the rapid growth of GPS and InSAR mapping is already providing a dense and accurate velocity fields in many active region. Suitable interpolation of this velocity field makes possible the strain rate computation (Kreemer et al., 2003). Using an inversion procedure, this would potentially allow computation of rigidity maps on continents. 7. Conclusion The idea that interseismic strain is linked to variable elastic plate thickness is markedly different than the usual geodetic data interpretation using the block model driven by fault slip at depth. I have argued that the former model is rheologically more plausible as it is directly linked to brittle and ductile lithosphere properties. In such a model, faults do not play a direct role during the interseismic phase because they are locked. Interseismic strain is therefore chiefly controlled by elastic plate properties (mostly bulk elastic modulus and plate thickness). Two main implications can be drawn: Appendix A The variable elastic thickness model has markedly different seismic hazard implications compared to those deduced from SB73 and block models. For the latter, fault slip rates are a free parameter to adjust. For the former, elastic thickness can be computed, but no direct link can be made with the long term fault slip rate. This lack of information is clearly related to the model's assumption: interseismic strain is interpreted on the time scale during which geodetic measurements are made. By contrast with the SB73 model, no hypothesis is made about the relation between short term interseismic strain and long term slip rate. Does this mean that interseismic strain cannot be used to predict fault slip rate? To discuss this important issue, let us consider again the formal differences between these two models. The most significant is the way to apply the boundary conditions to the model. In the case of SB73 model, the differential velocity is applied on the deep part of the fault (near field boundary condition, see Fig. A1a). In such a case, the fault slip rate is a kinematical parameter and has no relation to the fault strength. In the case of the variable elastic thickness model, the velocity boundary condition is remotely applied according to differential plate motion. Because the source of plate motion is thought to be controlled by large-scale body forces, this way to setup the boundary condition is probably wiser than a local drive. Let us the use of the block model to infer fault slip rate must be restricted to single faults embedded in high rigidity domains like for example the central north Anatolian fault or the central SAF. Incorrect slip rates are likely to be inferred if this condition is not fulfilled. This case occurs when multiple faults are present as in southern and northern California and suggests that such geodetically based fault slip rates could be unreliable. In a simple tectonic setting such as pure strike-slip faulting, high interseismic strain can be interpreted in the same way as a low rigidity (small elastic thickness) zone, while low strain would correspond to a higher rigidity (large thickness). If this 1D analysis can be extended to 2D based on stress equilibrium principle, the variable plate rigidity model could be applied to compute elastic thickness on continents with a suitable inversion procedure using interseismic strain as input data. Acknowledgements This paper benefited from several discussions with my colleagues. I particularly thank Xavier Le Pichon who forced me to clarify my interpretations. Philippe Vernant read an early version of the paper. Careful reviews of Wayne Thatcher and of two anonymous reviewers allowed to significantly improve the manuscript. I thank also Jessica Murray for helping me in gathering USGS geodetic data. SCEC and USGS are acknowledged for providing geodetic data of high quality. Generic Mapping Tool (GMT) software has been used to prepare most of the figures. Fig. A1. Differences and similarities between the thick lithosphere model and the variable thickness lithosphere model; a) thick elastic lithosphere model (SB73 model); b) modified SB73 model with remote boundary conditions c) variable thickness elastic lithosphere.

A mechanical model of the San Andreas fault and SAFOD Pilot Hole stress measurements

A mechanical model of the San Andreas fault and SAFOD Pilot Hole stress measurements GEOPHYSICAL RESEARCH LETTERS, VOL. 31, L15S13, doi:10.1029/2004gl019521, 2004 A mechanical model of the San Andreas fault and SAFOD Pilot Hole stress measurements Jean Chéry Laboratoire Dynamique de la

More information

Kinematics of the Southern California Fault System Constrained by GPS Measurements

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

More information

Elizabeth H. Hearn modified from W. Behr

Elizabeth H. Hearn modified from W. Behr Reconciling postseismic and interseismic surface deformation around strike-slip faults: Earthquake-cycle models with finite ruptures and viscous shear zones Elizabeth H. Hearn hearn.liz@gmail.com modified

More information

GPS Strain & Earthquakes Unit 5: 2014 South Napa earthquake GPS strain analysis student exercise

GPS Strain & Earthquakes Unit 5: 2014 South Napa earthquake GPS strain analysis student exercise GPS Strain & Earthquakes Unit 5: 2014 South Napa earthquake GPS strain analysis student exercise Strain Analysis Introduction Name: The earthquake cycle can be viewed as a process of slow strain accumulation

More information

Surface changes caused by erosion and sedimentation were treated by solving: (2)

Surface changes caused by erosion and sedimentation were treated by solving: (2) GSA DATA REPOSITORY 214279 GUY SIMPSON Model with dynamic faulting and surface processes The model used for the simulations reported in Figures 1-3 of the main text is based on two dimensional (plane strain)

More information

Jack Loveless Department of Geosciences Smith College

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

More information

Seismotectonics of intraplate oceanic regions. Thermal model Strength envelopes Plate forces Seismicity distributions

Seismotectonics of intraplate oceanic regions. Thermal model Strength envelopes Plate forces Seismicity distributions Seismotectonics of intraplate oceanic regions Thermal model Strength envelopes Plate forces Seismicity distributions Cooling of oceanic lithosphere also increases rock strength and seismic velocity. Thus

More information

Lecture 2: Deformation in the crust and the mantle. Read KK&V chapter 2.10

Lecture 2: Deformation in the crust and the mantle. Read KK&V chapter 2.10 Lecture 2: Deformation in the crust and the mantle Read KK&V chapter 2.10 Tectonic plates What are the structure and composi1on of tectonic plates? Crust, mantle, and lithosphere Crust relatively light

More information

SUPPLEMENTARY INFORMATION

SUPPLEMENTARY INFORMATION SUPPLEMENTARY INFORMATION doi: 10.1038/ngeo739 Supplementary Information to variability and distributed deformation in the Marmara Sea fault system Tobias Hergert 1 and Oliver Heidbach 1,* 1 Geophysical

More information

Regional Geodesy. Shimon Wdowinski. MARGINS-RCL Workshop Lithospheric Rupture in the Gulf of California Salton Trough Region. University of Miami

Regional Geodesy. Shimon Wdowinski. MARGINS-RCL Workshop Lithospheric Rupture in the Gulf of California Salton Trough Region. University of Miami MARGINS-RCL Workshop Lithospheric Rupture in the Gulf of California Salton Trough Region Regional Geodesy Shimon Wdowinski University of Miami Rowena Lohman, Kim Outerbridge, Tom Rockwell, and Gina Schmalze

More information

Rheology III. Ideal materials Laboratory tests Power-law creep The strength of the lithosphere The role of micromechanical defects in power-law creep

Rheology III. Ideal materials Laboratory tests Power-law creep The strength of the lithosphere The role of micromechanical defects in power-law creep Rheology III Ideal materials Laboratory tests Power-law creep The strength of the lithosphere The role of micromechanical defects in power-law creep Ideal materials fall into one of the following categories:

More information

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

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

More information

Global Tectonics. Kearey, Philip. Table of Contents ISBN-13: Historical perspective. 2. The interior of the Earth.

Global Tectonics. Kearey, Philip. Table of Contents ISBN-13: Historical perspective. 2. The interior of the Earth. Global Tectonics Kearey, Philip ISBN-13: 9781405107778 Table of Contents Preface. Acknowledgments. 1. Historical perspective. 1.1 Continental drift. 1.2 Sea floor spreading and the birth of plate tectonics.

More information

Material is perfectly elastic until it undergoes brittle fracture when applied stress reaches σ f

Material is perfectly elastic until it undergoes brittle fracture when applied stress reaches σ f Material is perfectly elastic until it undergoes brittle fracture when applied stress reaches σ f Material undergoes plastic deformation when stress exceeds yield stress σ 0 Permanent strain results from

More information

Estimating fault slip rates, locking distribution, elastic/viscous properites of lithosphere/asthenosphere. Kaj M. Johnson Indiana University

Estimating fault slip rates, locking distribution, elastic/viscous properites of lithosphere/asthenosphere. Kaj M. Johnson Indiana University 3D Viscoelastic Earthquake Cycle Models Estimating fault slip rates, locking distribution, elastic/viscous properites of lithosphere/asthenosphere Kaj M. Johnson Indiana University In collaboration with:

More information

Measurements in the Creeping Section of the Central San Andreas Fault

Measurements in the Creeping Section of the Central San Andreas Fault Measurements in the Creeping Section of the Central San Andreas Fault Introduction Duncan Agnew, Andy Michael We propose the PBO instrument, with GPS and borehole strainmeters, the creeping section of

More information

DETAILS ABOUT THE TECHNIQUE. We use a global mantle convection model (Bunge et al., 1997) in conjunction with a

DETAILS ABOUT THE TECHNIQUE. We use a global mantle convection model (Bunge et al., 1997) in conjunction with a DETAILS ABOUT THE TECHNIQUE We use a global mantle convection model (Bunge et al., 1997) in conjunction with a global model of the lithosphere (Kong and Bird, 1995) to compute plate motions consistent

More information

Comparison of Strain Rate Maps

Comparison of Strain Rate Maps Comparison of Strain Rate Maps David T. Sandwell UNAVCO March 8, 2010 why strain rate matters comparison of 10 strain rate models new data required interseismic model velocity v(x) = V π tan 1 x D strain

More information

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

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

More information

Description of faults

Description of faults GLG310 Structural Geology Description of faults Horizontal stretch Crustal thickness Regional elevation Regional character Issues Normal Thrust/reverse Strike-slip >1 1 in one direction and < 1 in

More information

Regional deformation and kinematics from GPS data

Regional deformation and kinematics from GPS data Regional deformation and kinematics from GPS data Jessica Murray, Jerry Svarc, Elizabeth Hearn, and Wayne Thatcher U. S. Geological Survey Acknowledgements: Rob McCaffrey, Portland State University UCERF3

More information

The distributions of slip rate and ductile deformation in a strikeslip

The distributions of slip rate and ductile deformation in a strikeslip Geophys. J. Int. (22) 48, 79 92 The distributions of slip rate and ductile deformation in a strikeslip shear zone Frédérique Rolandone and Claude Jaupart Institut de Physique du Globe de Paris, 4, Place

More information

Power-law distribution of fault slip-rates in southern California

Power-law distribution of fault slip-rates in southern California Click Here for Full Article GEOPHYSICAL RESEARCH LETTERS, VOL. 34, L23307, doi:10.1029/2007gl031454, 2007 Power-law distribution of fault slip-rates in southern California Brendan J. Meade 1 Received 31

More information

Initiation of the San Jacinto Fault and its Interaction with the San Andreas Fault: Insights from Geodynamic Modeling

Initiation of the San Jacinto Fault and its Interaction with the San Andreas Fault: Insights from Geodynamic Modeling Pure appl. geophys. (2007) DOI 10.1007/s00024-007-0262-z Ó Birkhäuser Verlag, Basel, 2007 Pure and Applied Geophysics Initiation of the San Jacinto Fault and its Interaction with the San Andreas Fault:

More information

Ground displacement in a fault zone in the presence of asperities

Ground displacement in a fault zone in the presence of asperities BOLLETTINO DI GEOFISICA TEORICA ED APPLICATA VOL. 40, N. 2, pp. 95-110; JUNE 2000 Ground displacement in a fault zone in the presence of asperities S. SANTINI (1),A.PIOMBO (2) and M. DRAGONI (2) (1) Istituto

More information

M 7.0 earthquake recurrence on the San Andreas fault from a stress renewal model

M 7.0 earthquake recurrence on the San Andreas fault from a stress renewal model Click Here for Full Article JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 111,, doi:10.1029/2006jb004415, 2006 M 7.0 earthquake recurrence on the San Andreas fault from a stress renewal model Tom Parsons 1 Received

More information

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

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

More information

The influence of short wavelength variations in viscosity on subduction dynamics

The influence of short wavelength variations in viscosity on subduction dynamics 1 Introduction Deformation within the earth, driven by mantle convection due primarily to cooling and subduction of oceanic lithosphere, is expressed at every length scale in various geophysical observations.

More information

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

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

More information

Plate Boundary Observatory Working Group for the Central and Northern San Andreas Fault System PBO-WG-CNSA

Plate Boundary Observatory Working Group for the Central and Northern San Andreas Fault System PBO-WG-CNSA Plate Boundary Observatory Working Group for the Central and Northern San Andreas Fault System PBO-WG-CNSA Introduction Our proposal focuses on the San Andreas fault system in central and northern California.

More information

Geodynamics Lecture 5 Basics of elasticity

Geodynamics Lecture 5 Basics of elasticity Geodynamics Lecture 5 Basics of elasticity Lecturer: David Whipp david.whipp@helsinki.fi! 16.9.2014 Geodynamics www.helsinki.fi/yliopisto 1 Goals of this lecture Introduce linear elasticity! Look at the

More information

Elastic Rebound Theory

Elastic Rebound Theory Earthquakes Elastic Rebound Theory Earthquakes occur when strain exceeds the strength of the rock and the rock fractures. The arrival of earthquakes waves is recorded by a seismograph. The amplitude of

More information

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

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

More information

Can geodetic strain rate be useful in seismic hazard studies?

Can geodetic strain rate be useful in seismic hazard studies? Can geodetic strain rate be useful in seismic hazard studies? F. Riguzzi 1, R. Devoti 1, G. Pietrantonio 1, M. Crespi 2, C. Doglioni 2, A.R. Pisani 1 Istituto Nazionale di Geofisica e Vulcanologia 2 Università

More information

Integrating Geologic and Geodetic Estimates of Slip Rate on the San Andreas Fault System

Integrating Geologic and Geodetic Estimates of Slip Rate on the San Andreas Fault System International Geology Review, Vol. 44, 2002, p. 62 82. Copyright 2002 by V. H. Winston & Son, Inc. All rights reserved. Integrating Geologic and Geodetic Estimates of Slip Rate on the San Andreas Fault

More information

Predicted reversal and recovery of surface creep on the Hayward fault following the 1906 San Francisco earthquake

Predicted reversal and recovery of surface creep on the Hayward fault following the 1906 San Francisco earthquake GEOPHYSICAL RESEARCH LETTERS, VOL. 35, L19305, doi:10.1029/2008gl035270, 2008 Predicted reversal and recovery of surface creep on the Hayward fault following the 1906 San Francisco earthquake D. A. Schmidt

More information

Slip rates and off-fault deformation in Southern California inferred from GPS data and models

Slip rates and off-fault deformation in Southern California inferred from GPS data and models JOURNAL OF GEOPHYSICAL RESEARCH: SOLID EARTH, VOL. 8, 6 66, doi:./jgrb.6, Slip rates and off-fault deformation in Southern California inferred from GPS data and models K. M. Johnson Received 9 December

More information

Summary so far. Geological structures Earthquakes and their mechanisms Continuous versus block-like behavior Link with dynamics?

Summary so far. Geological structures Earthquakes and their mechanisms Continuous versus block-like behavior Link with dynamics? Summary so far Geodetic measurements velocities velocity gradient tensor (spatial derivatives of velocity) Velocity gradient tensor = strain rate (sym.) + rotation rate (antisym.) Strain rate tensor can

More information

Data Repository Hampel et al., page 1/5

Data Repository Hampel et al., page 1/5 GSA DATA REPOSITORY 2138 Data Repositor Hampel et al., page 1/5 SETUP OF THE FINITE-ELEMENT MODEL The finite-element models were created with the software ABAQUS and consist of a 1-km-thick lithosphere,

More information

GPS Strain & Earthquakes Unit 4: GPS strain analysis examples Student exercise

GPS Strain & Earthquakes Unit 4: GPS strain analysis examples Student exercise GPS Strain & Earthquakes Unit 4: GPS strain analysis examples Student exercise Example 1: Olympic Peninsula Name: Please complete the following worksheet to estimate, calculate, and interpret the strain

More information

Seismic and aseismic processes in elastodynamic simulations of spontaneous fault slip

Seismic and aseismic processes in elastodynamic simulations of spontaneous fault slip Seismic and aseismic processes in elastodynamic simulations of spontaneous fault slip Most earthquake simulations study either one large seismic event with full inertial effects or long-term slip history

More information

The Earthquake Cycle Chapter :: n/a

The Earthquake Cycle Chapter :: n/a The Earthquake Cycle Chapter :: n/a A German seismogram of the 1906 SF EQ Image courtesy of San Francisco Public Library Stages of the Earthquake Cycle The Earthquake cycle is split into several distinct

More information

Journal of Geophysical Research Letters Supporting Information for

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

More information

Stress equilibrium in southern California from Maxwell stress function models fit to both earthquake data and a quasi-static dynamic simulation

Stress equilibrium in southern California from Maxwell stress function models fit to both earthquake data and a quasi-static dynamic simulation Stress equilibrium in southern California from Maxwell stress function models fit to both earthquake data and a quasi-static dynamic simulation Peter Bird Dept. of Earth, Planetary, and Space Sciences

More information

Earthquake and Volcano Deformation

Earthquake and Volcano Deformation Earthquake and Volcano Deformation Paul Segall Stanford University Draft Copy September, 2005 Last Updated Sept, 2008 COPYRIGHT NOTICE: To be published by Princeton University Press and copyrighted, c

More information

Whole Earth Structure and Plate Tectonics

Whole Earth Structure and Plate Tectonics Whole Earth Structure and Plate Tectonics Processes in Structural Geology & Tectonics Ben van der Pluijm WW Norton+Authors, unless noted otherwise 4/5/2017 14:45 We Discuss Whole Earth Structure and Plate

More information

Lab 1: Plate Tectonics April 2, 2009

Lab 1: Plate Tectonics April 2, 2009 Name: Lab 1: Plate Tectonics April 2, 2009 Objective: Students will be introduced to the theory of plate tectonics and different styles of plate margins and interactions. Introduction The planet can be

More information

Gravity Tectonics Volcanism Atmosphere Water Winds Chemistry. Planetary Surfaces

Gravity Tectonics Volcanism Atmosphere Water Winds Chemistry. Planetary Surfaces Gravity Tectonics Volcanism Atmosphere Water Winds Chemistry Planetary Surfaces Gravity & Rotation Polar flattening caused by rotation is the largest deviation from a sphere for a planet sized object (as

More information

Dynamic models of interseismic deformation and stress transfer from plate motion to continental transform faults

Dynamic models of interseismic deformation and stress transfer from plate motion to continental transform faults JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 117,, doi:10.1029/2011jb009056, 2012 Dynamic models of interseismic deformation and stress transfer from plate motion to continental transform faults Christopher S.

More information

MAR110 Lecture #5 Plate Tectonics-Earthquakes

MAR110 Lecture #5 Plate Tectonics-Earthquakes 1 MAR110 Lecture #5 Plate Tectonics-Earthquakes Figure 5.0 Plate Formation & Subduction Destruction The formation of the ocean crust from magma that is upwelled into a pair of spreading centers. Pairs

More information

Can Lateral Viscosity Contrasts Explain Asymmetric Interseismic Deformation around Strike-Slip faults?

Can Lateral Viscosity Contrasts Explain Asymmetric Interseismic Deformation around Strike-Slip faults? Can Lateral Viscosity Contrasts Explain Asymmetric Interseismic Deformation around Strike-Slip faults? Ali Vaghri 1 and Elizabeth H. Hearn 1 1 Department of Earth and Ocean Sciences University of British

More information

When you are standing on a flat surface, what is the normal stress you exert on the ground? What is the shear stress?

When you are standing on a flat surface, what is the normal stress you exert on the ground? What is the shear stress? When you are standing on a flat surface, what is the normal stress you exert on the ground? What is the shear stress? How could you exert a non-zero shear stress on the ground? Hydrostatic Pressure (fluids)

More information

Effect of an outer-rise earthquake on seismic cycle of large interplate earthquakes estimated from an instability model based on friction mechanics

Effect of an outer-rise earthquake on seismic cycle of large interplate earthquakes estimated from an instability model based on friction mechanics Effect of an outer-rise earthquake on seismic cycle of large interplate earthquakes estimated from an instability model based on friction mechanics Naoyuki Kato (1) and Tomowo Hirasawa (2) (1) Geological

More information

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

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

More information

Strain-dependent strength profiles Implication of planetary tectonics

Strain-dependent strength profiles Implication of planetary tectonics Strain-dependent strength profiles Implication of planetary tectonics Laurent G.J. Montési 1 Frederic Gueydan 2, Jacques Précigout 3 1 University of Maryland 2 Université de Montpellier 2, 3 Université

More information

What is the LAB Dynamically: Lithosphere and Asthenosphere Rheology from Post-loading Deformation

What is the LAB Dynamically: Lithosphere and Asthenosphere Rheology from Post-loading Deformation What is the LAB Dynamically: Lithosphere and Asthenosphere Rheology from Post-loading Deformation Roland Bürgmann, UC Berkeley With contributions by Pascal Audet, Daula Chandrasekhar, Georg Dresen, Andy

More information

Analytic and Numeric Tests of Fourier Deformation Model (Copyright 2003, Bridget R. Smith and David T. Sandwell)

Analytic and Numeric Tests of Fourier Deformation Model (Copyright 2003, Bridget R. Smith and David T. Sandwell) Analytic and Numeric Tests of Fourier Deformation Model (Copyright 2003, Bridget R. Smith and David T. Sandwell) Although the solutions of our Fourier deformation model have been checked using computer

More information

} based on composition

} based on composition Learning goals: Predict types of earthquakes that will happen at different plate boundaries based on relative plate motion vector vs. strike (vector subtraction) Understand interseismic and coseismic deformation,

More information

Rheology. What is rheology? From the root work rheo- Current: flow. Greek: rhein, to flow (river) Like rheostat flow of current

Rheology. What is rheology? From the root work rheo- Current: flow. Greek: rhein, to flow (river) Like rheostat flow of current Rheology What is rheology? From the root work rheo- Current: flow Greek: rhein, to flow (river) Like rheostat flow of current Rheology What physical properties control deformation? - Rock type - Temperature

More information

Far-reaching transient motions after Mojave earthquakes require broad mantle flow beneath a strong crust

Far-reaching transient motions after Mojave earthquakes require broad mantle flow beneath a strong crust Click Here for Full Article GEOPHYSICAL RESEARCH LETTERS, VOL. 34, L19302, doi:10.1029/2007gl030959, 2007 Far-reaching transient motions after Mojave earthquakes require broad mantle flow beneath a strong

More information

Geology for Engineers Rock Mechanics and Deformation of Earth Materials

Geology for Engineers Rock Mechanics and Deformation of Earth Materials 89.325 Geology for Engineers Rock Mechanics and Deformation of Earth Materials Why do rocks break? Rock mechanics experiments a first order understanding. Faults and Fractures Triaxial load machine. a)

More information

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 110, B09401, doi: /2004jb003548, 2005

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 110, B09401, doi: /2004jb003548, 2005 JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 110,, doi:10.1029/2004jb003548, 2005 Fault slip rates, effects of elastic heterogeneity on geodetic data, and the strength of the lower crust in the Salton Trough

More information

Source parameters II. Stress drop determination Energy balance Seismic energy and seismic efficiency The heat flow paradox Apparent stress drop

Source parameters II. Stress drop determination Energy balance Seismic energy and seismic efficiency The heat flow paradox Apparent stress drop Source parameters II Stress drop determination Energy balance Seismic energy and seismic efficiency The heat flow paradox Apparent stress drop Source parameters II: use of empirical Green function for

More information

Directed Reading. Section: The Theory of Plate Tectonics. to the development of plate tectonics, developed? HOW CONTINENTS MOVE

Directed Reading. Section: The Theory of Plate Tectonics. to the development of plate tectonics, developed? HOW CONTINENTS MOVE Skills Worksheet Directed Reading Section: The Theory of Plate Tectonics 1. The theory that explains why and how continents move is called. 2. By what time period was evidence supporting continental drift,

More information

The Theory of Plate Tectonics

The Theory of Plate Tectonics Plate Tectonics Objectives Describe how plates move. Explain the features of plate tectonics. Describe the types of plate boundaries and the features that can form and events that can occur at each. The

More information

Deformation cycles of great subduction earthquakes in a viscoelastic Earth

Deformation cycles of great subduction earthquakes in a viscoelastic Earth Deformation cycles of great subduction earthquakes in a viscoelastic Earth Kelin Wang Pacific Geoscience Centre, Geological Survey of Canada School of Earth and Ocean Science, University of Victoria????

More information

Physics and Chemistry of the Earth and Terrestrial Planets

Physics and Chemistry of the Earth and Terrestrial Planets MIT OpenCourseWare http://ocw.mit.edu 12.002 Physics and Chemistry of the Earth and Terrestrial Planets Fall 2008 For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.

More information

High-Harmonic Geoid Signatures due to Glacial Isostatic Adjustment, Subduction and Seismic Deformation

High-Harmonic Geoid Signatures due to Glacial Isostatic Adjustment, Subduction and Seismic Deformation High-Harmonic Geoid Signatures due to Glacial Isostatic Adjustment, Subduction and Seismic Deformation L.L.A. Vermeersen (1), H. Schotman (1), M.-W. Jansen (1), R. Riva (1) and R. Sabadini (2) (1) DEOS,

More information

Gravitational constraints

Gravitational constraints Gravitational constraints Reading: Fowler p172 187 Gravity anomalies Free-air anomaly: g F = g g( λ ) + δg obs F Corrected for expected variations due to the spheroid elevation above the spheroid Bouguer

More information

Evolution of stress in Southern California for the past 200 years from coseismic, postseismic and interseismic stress changes

Evolution of stress in Southern California for the past 200 years from coseismic, postseismic and interseismic stress changes Geophys. J. Int. (2007) 169, 1164 1179 doi: 10.1111/j.1365-246X.2007.03391.x Evolution of stress in Southern California for the past 200 years from coseismic, postseismic and interseismic stress changes

More information

Lecture 20: Slow Slip Events and Stress Transfer. GEOS 655 Tectonic Geodesy Jeff Freymueller

Lecture 20: Slow Slip Events and Stress Transfer. GEOS 655 Tectonic Geodesy Jeff Freymueller Lecture 20: Slow Slip Events and Stress Transfer GEOS 655 Tectonic Geodesy Jeff Freymueller Slow Slip Events From Kristine Larson What is a Slow Slip Event? Slip on a fault, like in an earthquake, BUT

More information

Mid-Continent Earthquakes As A Complex System

Mid-Continent Earthquakes As A Complex System SRL complex earthquakes 5/22/09 1 Mid-Continent Earthquakes As A Complex System Niels Bohr once observed How wonderful that we have met with a paradox. Now we have some hope of making progress. This situation

More information

12. The diagram below shows the collision of an oceanic plate and a continental plate.

12. The diagram below shows the collision of an oceanic plate and a continental plate. Review 1. Base your answer to the following question on the cross section below, which shows the boundary between two lithospheric plates. Point X is a location in the continental lithosphere. The depth

More information

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

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

More information

Announcements. Manganese nodule distribution

Announcements. Manganese nodule distribution Announcements Lithospheric plates not as brittle as previously thought ESCI 322 Meet in Env. Studies Bldg Rm 60 at 1 PM on Tuesday One week (Thursday): Quiz on Booth 1994 and discussion. (Lots of odd terms

More information

State of Stress in Seismic Gaps Along the SanJacinto Fault

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

More information

Toward a SCEC Community Rheology Model: TAG Kickoff and Workshop SCEC Workshop Proposal Final Report

Toward a SCEC Community Rheology Model: TAG Kickoff and Workshop SCEC Workshop Proposal Final Report Toward a SCEC Community Rheology Model: TAG Kickoff and Workshop SCEC Workshop Proposal 17206 Final Report A one-day Community Rheology Model workshop was held at the Palm Springs Hilton on the Saturday

More information

Coulomb stress accumulation along the San Andreas Fault system

Coulomb stress accumulation along the San Andreas Fault system JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 108, NO. B6, 2296, doi:10.1029/2002jb002136, 2003 Coulomb stress accumulation along the San Andreas Fault system Bridget Smith and David Sandwell Institute for Geophysics

More information

Supplementary Material

Supplementary Material 1 Supplementary Material 2 3 4 Interseismic, megathrust earthquakes and seismic swarms along the Chilean subduction zone (38-18 S) 5 6 7 8 9 11 12 13 14 1 GPS data set We combined in a single data set

More information

Development of a Predictive Simulation System for Crustal Activities in and around Japan - II

Development of a Predictive Simulation System for Crustal Activities in and around Japan - II Development of a Predictive Simulation System for Crustal Activities in and around Japan - II Project Representative Mitsuhiro Matsu'ura Graduate School of Science, The University of Tokyo Authors Mitsuhiro

More information

EARTHQUAKE LOCATIONS INDICATE PLATE BOUNDARIES EARTHQUAKE MECHANISMS SHOW MOTION

EARTHQUAKE LOCATIONS INDICATE PLATE BOUNDARIES EARTHQUAKE MECHANISMS SHOW MOTION 6-1 6: EARTHQUAKE FOCAL MECHANISMS AND PLATE MOTIONS Hebgen Lake, Montana 1959 Ms 7.5 1 Stein & Wysession, 2003 Owens Valley, California 1872 Mw ~7.5 EARTHQUAKE LOCATIONS INDICATE PLATE BOUNDARIES EARTHQUAKE

More information

Activity Pacific Northwest Tectonic Block Model

Activity Pacific Northwest Tectonic Block Model Activity Pacific Northwest Tectonic Block Model The Cascadia tectonic margin is caught between several tectonic forces, during the relentless motions of the giant Pacific Plate, the smaller subducting

More information

Regional tectonic stress near the San Andreas fault in central and southern California

Regional tectonic stress near the San Andreas fault in central and southern California GEOPHYSICAL RESEARCH LETTERS, VOL. 31, L15S11, doi:10.1029/2003gl018918, 2004 Regional tectonic stress near the San Andreas fault in central and southern California J. Townend School of Earth Sciences,

More information

MAR110 LECTURE #6 West Coast Earthquakes & Hot Spots

MAR110 LECTURE #6 West Coast Earthquakes & Hot Spots 17 September 2007 Lecture 6 West Coast Earthquakes & Hot Spots 1 MAR110 LECTURE #6 West Coast Earthquakes & Hot Spots Figure 6.1 Plate Formation & Subduction Destruction The formation of the ocean crust

More information

Geo736: Seismicity and California s Active Faults Introduction

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

More information

Plate Tectonics and the cycling of Earth materials

Plate Tectonics and the cycling of Earth materials Plate Tectonics and the cycling of Earth materials Plate tectonics drives the rock cycle: the movement of rocks (and the minerals that comprise them, and the chemical elements that comprise them) from

More information

Captain s Tryouts 2017

Captain s Tryouts 2017 Captain s Tryouts 2017 Dynamic Planet Test Written by: Araneesh Pratap (Chattahoochee High School) Name: Date: Answer all questions on the answer sheet. Point values are given next to each question or

More information

GPS strain analysis examples Instructor notes

GPS strain analysis examples Instructor notes GPS strain analysis examples Instructor notes Compiled by Phil Resor (Wesleyan University) This document presents several examples of GPS station triplets for different tectonic environments. These examples

More information

Deformation of Rocks. Orientation of Deformed Rocks

Deformation of Rocks. Orientation of Deformed Rocks Deformation of Rocks Folds and faults are geologic structures caused by deformation. Structural geology is the study of the deformation of rocks and its effects. Fig. 7.1 Orientation of Deformed Rocks

More information

Basics of the modelling of the ground deformations produced by an earthquake. EO Summer School 2014 Frascati August 13 Pierre Briole

Basics of the modelling of the ground deformations produced by an earthquake. EO Summer School 2014 Frascati August 13 Pierre Briole Basics of the modelling of the ground deformations produced by an earthquake EO Summer School 2014 Frascati August 13 Pierre Briole Content Earthquakes and faults Examples of SAR interferograms of earthquakes

More information

Depth (Km) + u ( ξ,t) u = v pl. η= Pa s. Distance from Nankai Trough (Km) u(ξ,τ) dξdτ. w(x,t) = G L (x,t τ;ξ,0) t + u(ξ,t) u(ξ,t) = v pl

Depth (Km) + u ( ξ,t) u = v pl. η= Pa s. Distance from Nankai Trough (Km) u(ξ,τ) dξdτ. w(x,t) = G L (x,t τ;ξ,0) t + u(ξ,t) u(ξ,t) = v pl Slip history during one earthquake cycle at the Nankai subduction zone, inferred from the inversion analysis of levelling data with a viscoelastic slip response function Mitsuhiro Matsu'ura, Akira Nishitani

More information

The Mechanics of Earthquakes and Faulting

The Mechanics of Earthquakes and Faulting The Mechanics of Earthquakes and Faulting Christopher H. Scholz Lamont-Doherty Geological Observatory and Department of Earth and Environmental Sciences, Columbia University 2nd edition CAMBRIDGE UNIVERSITY

More information

Azimuth with RH rule. Quadrant. S 180 Quadrant Azimuth. Azimuth with RH rule N 45 W. Quadrant Azimuth

Azimuth with RH rule. Quadrant. S 180 Quadrant Azimuth. Azimuth with RH rule N 45 W. Quadrant Azimuth 30 45 30 45 Strike and dip notation (a) N30 E, 45 SE ("Quadrant"): the bearing of the strike direction is 30 degrees east of north and the dip is 45 degrees in a southeast (SE) direction. For a given strike,

More information

Plate Tectonics - Demonstration

Plate Tectonics - Demonstration Name: Reference: Prof. Larry Braile - Educational Resources Copyright 2000. L. Braile. Permission granted for reproduction for non-commercial uses. http://web.ics.purdue.edu/~braile/indexlinks/educ.htm

More information

Geologic Structures. Changes in the shape and/or orientation of rocks in response to applied stress

Geologic Structures. Changes in the shape and/or orientation of rocks in response to applied stress Geologic Structures Changes in the shape and/or orientation of rocks in response to applied stress Figure 15.19 Can be as big as a breadbox Or much bigger than a breadbox Three basic types Fractures >>>

More information

The numerical method used for experiments is based on an explicit finite element

The numerical method used for experiments is based on an explicit finite element Bialas 1 Model Supplementary Data The numerical method used for experiments is based on an explicit finite element technique similar to the Fast Lagrangian Analysis of Continua (FLAC) method (Cundall,

More information

Fault Slip Rates, Effects of Sediments and the Strength of the Lower Crust in the Salton Trough Region, Southern California

Fault Slip Rates, Effects of Sediments and the Strength of the Lower Crust in the Salton Trough Region, Southern California Fault Slip Rates, Effects of Sediments and the Strength of the Lower Crust in the Salton Trough Region, Southern California Noah P. Fay and Eugene D. Humphreys University of Oregon Noah Fay Department

More information

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 113, B12415, doi: /2008jb005809, 2008

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 113, B12415, doi: /2008jb005809, 2008 JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 113,, doi:10.1029/2008jb005809, 2008 Forces acting on the Sierra Nevada block and implications for the strength of the San Andreas fault system and the dynamics of

More information

4 Deforming the Earth s Crust

4 Deforming the Earth s Crust CHAPTER 7 4 Deforming the Earth s Crust SECTION Plate Tectonics BEFORE YOU READ After you read this section, you should be able to answer these questions: What happens when rock is placed under stress?

More information

OCN 201: Seafloor Spreading and Plate Tectonics I

OCN 201: Seafloor Spreading and Plate Tectonics I OCN 201: Seafloor Spreading and Plate Tectonics I Revival of Continental Drift Theory Kiyoo Wadati (1935) speculated that earthquakes and volcanoes may be associated with continental drift. Hugo Benioff

More information