Izmit earthquake postseismic deformation and dynamics of the North Anatolian Fault Zone

Size: px
Start display at page:

Download "Izmit earthquake postseismic deformation and dynamics of the North Anatolian Fault Zone"

Transcription

1 JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 114,, doi: /2008jb006026, 2009 Izmit earthquake postseismic deformation and dynamics of the North Anatolian Fault Zone E. H. Hearn, 1 S. McClusky, 2 S. Ergintav, 3 and R. E. Reilinger 2 Received 21 August 2008; revised 13 March 2009; accepted 4 May 2009; published 25 August [1] We have modeled postseismic deformation from 1999 to 2003 in the region surrounding the 1999 Izmit and Düzce earthquake ruptures, using a three-dimensional viscoelastic finite element method. Our models confirm earlier findings that surface deformation within the first few months of the Izmit earthquake is principally due to stable frictional afterslip on and below the Izmit earthquake rupture. A second deformation process is required, however, to fit the surface deformation after several months. Viscoelastic relaxation of lower crust and/or upper mantle with a viscosity of the order of 2to Pa s improves the models fit to later GPS site velocities. However, for a linear viscous rheology, this range of values is inconsistent with highly localized interseismic deformation around the North Anatolian Fault Zone (NAFZ) that was well observed prior to the earthquake sequence. The simplest solution to this problem is to assume that the effective viscosity of the relaxing material increases with time after large earthquakes, that is, that it has a power law or Burger s body (transient) rheology. A Burger s body rheology with two characteristic viscosities (2 to Pa s and at least Pa s) in the mantle is consistent with deformation around the NAFZ throughout the earthquake cycle. Citation: Hearn, E. H., S. McClusky, S. Ergintav, and R. E. Reilinger (2009), Izmit earthquake postseismic deformation and dynamics of the North Anatolian Fault Zone, J. Geophys. Res., 114,, doi: /2008jb Introduction [2] Deformation around the North Anatolian Fault Zone (NAFZ) has been monitored with high-precision space geodetic techniques since 1988 [e.g., Reilinger et al., 2006; Ergintav et al., 2002; Wright et al., 2001; McClusky et al., 2000; Kahle et al., 1998; Straub et al., 1997]. Data from these studies illustrate how the Earth s surface deforms around a plate boundary fault, throughout the earthquake cycle. Previously, we have modeled interseismic deformation around the 1940s rupture segment of the NAFZ, as well as coseismic and early postseismic deformation from the 1999 Izmit earthquake [Hearn et al., 2002a, 2002b; Reilinger et al., 2000; Hearn and Bürgmann, 2005]. We have concluded from these studies that the upper mantle viscosity must exceed Pa s (assuming linear viscoelasticity), that the NAFZ has an aseismic extension penetrating most or all of the crust, and that the aseismic creep along this shear zone (in the upper to middle crust) has a weakly velocity-strengthening rheology. [3] Here, we model postseismic deformation following the Izmit-Düzce earthquake sequence over a longer time period 1 Department of Earth and Ocean Sciences, University of British Columbia, Vancouver, British Columbia, Canada. 2 Department of Earth, Atmospheric, and Planetary Sciences, Massachusetts Institute of Technology, Cambridge, Massachusetts, USA. 3 TUBITAK, Marmara Research Center, Earth and Marine Sciences Institute, Gebze, Turkey. Copyright 2009 by the American Geophysical Union /09/2008JB006026$09.00 (from 1999 to 2003). We develop plate boundary models incorporating viscous fault zone creep, stable frictional afterslip, viscoelastic relaxation of nonlinearly or linearly viscoelastic lower crust, and combinations of afterslip and viscoelastic relaxation. We confirm that the earliest Izmit postseismic deformation is consistent with velocitystrengthening frictional afterslip, though viscous shear zone creep is probably occurring at depth. After several months, none of the afterslip models produces high enough slip rates to explain the GPS postseismic velocities. To address this, we explore models with viscoelastic relaxation of the lower crust and upper mantle, in addition to afterslip, and we investigate circumstances under which such models can be consistent with the localized interseismic deformation around the NAFZ. The GPS data analysis and velocity fields from the first 6 years after Izmit postseismic deformation, and kinematic afterslip inversions based on these data, are described by the companion paper by Ergintav et al. [2009]. 2. Methods [4] We use GAEA [Saucier and Humphreys, 1993], a 3-D viscoelastic finite element (FE) code, to model time-dependent postseismic deformation. The FE code solves for nodal displacements resulting from elastic deformation of the modeled volume in response to applied displacements or velocities. For the models presented here, kinematically imposed coseismic displacements across the Izmit and Düzce earthquake ruptures induce elastic stresses, which drive subsequent, distributed viscoelastic creep or fault zone slip. 1of21

2 Figure 1. Model mesh, showing the locations of the Izmit and Düzce ruptures, epicenters (stars), and GPS sites (small circles). Dashed lines show faults comprising the NAF system in the Marmara Sea, including the Main Marmara Fault. The Main Marmara Fault geometry is based on that of Demirbağ et al. [2003]. Inset shows the location of the modeled region. To calculate nodal displacements, GAEA uses a Galerkin weighted-residual method, incorporating 20-node block elements with quadratic shape and weighting functions. This allows smoothly varying fault geometries and slip distributions, and minimizes discontinuities in stresses between model elements. Distributed viscoelastic deformation is modeled by calculating the viscous strain that would occur over one time step (given the element stresses and viscosity), and using this information to update the element force vector. The same approach is employed in other FE codes, such as TECTON [Williams and Richardson, 1991] and PyLITH [Williams et al., 2005]. Postseismic fault zone creep is modeled by calculating the shear stress along a fault and using the constitutive relationship to calculate the slip per time step, and adding this to the slip at each fault node. Fault slip is imposed using the split node technique [Melosh and Raefsky, 1981]. Modeled surface displacements from GAEA are comparable to the elastic solutions of Okada [1985] and Wang et al. [2006], and the Savage [1990] viscoelastic earthquake cycle solution [Hearn, 2003; Hearn et al., 2005; Hearn and Bürgmann, 2005]. [5] The finite element mesh covers a 1000 km by 900 km area, extends to a depth of 330 km, and is centered on the Izmit rupture (Figure 1). The side and bottom model boundaries are fixed and the surface is modeled as a stressfree boundary. The mesh is unstructured, with nodes at 1 2 km intervals along the Izmit rupture. The elasticity structure LAY2 from Hearn and Bürgmann [2005] is assumed, as is the Izmit earthquake slip distribution obtained for this elastic model. In the LAY2 model, the Poisson s ratio is 0.25 and the Young s modulus is 30, 45, and 70 GPa at depth intervals of 0 16, 16 32, and 32+ km. In afterslip models, the crust viscosity is Pa s and the mantle viscosity is Pa s. The upper crust viscosity (0 to 16 km depth) is held at Pa s in all models. For the Izmit earthquake, the fault geometry is based on the geometry of the surface rupture, but represented as a single, continuous surface and extended downward. This surface is somewhat less smooth than that of Feigl et al. [2002], which is also shown for reference on Figures 5, 7a, 7b, 9, and 11. For the Düzce earthquake, we use the slip distribution from Bürgmann et al. [2002a]. Because there are few GPS stations with postseismic velocities this area, we model this surface as vertical (which simplifies meshing). Coseismic slip is interpolated from the centers of kinematic slip model patches to the fault node coordinates in the FE mesh, and is imposed using the split 2of21

3 We distinguish WRSS from WRSS t, which is the measure of model misfit to GPS velocity at an individual time epoch: WRSS t ¼ X55 s¼1 X 2 d¼1 D 2 t;s;d ð2þ s t;s;d Figure 2. Schematics of the lithosphere models. The black sections of fault zones slip coseismically. Dark gray regions are high-viscosity material (h =10 23 Pa s in the upper crust and Pa s below a depth of 16 km). The models are (a) frictional afterslip; (b) viscoelastic, finite width shear zone; (c) nonlinear viscoelastic lower crust (between 22 and 32 km depth); and (d) dual-process model, combining frictional afterslip with linear viscoelastic relaxation of lower crust and upper mantle. In Figures 2a and 2d, the dashed fault zone represents intervals where the model may choose between viscous creep with h/w = Pa s m 1 and frictional afterslip, depending on which yields faster slip. node technique [Melosh and Raefsky, 1981]. The Izmit earthquake slip is imposed at t =0dayandtheDüzce earthquake slip at t = 87 days. Viscoelastic structure and fault rheology are varied in the simulations, as described in sections and section GPS Velocities and Measures of Model Misfit [6] We compare our model results with surface velocities at 55 GPS sites from the region surrounding the Izmit and Düzce ruptures (small circles on Figure 1). The GPS postseismic data analysis is described in detail by Ergintav et al. [2009], who provide weekly velocities and errors at all continuous GPS sites in their analysis (which covers a broader region and a longer time span than we discuss here) in their auxiliary material. Ergintav et al. fit position measurements at survey GPS sites to a three-term logarithmic function, and they provide coefficients to this function which enable us to compute velocities and errors at any time [Ergintav et al., 2009, auxiliary material Data Set S1]. Secular GPS site velocities, computed with a block model [Reilinger et al., 2006], were subtracted from their GPS velocities to isolate the postseismic contribution. [7] To measure the performance of our finite element models, we calculate the weighted-residual sum of squares (WRSS), which is a measure of model misfit to the GPS velocity data. A weighted residual is found by subtracting the finite element modeled velocity from the GPS site velocity, and dividing this residual by the one-sigma GPS velocity error. This quantity is squared and summed for both horizontal velocity components (east and north), for each GPS site, for 900 daily intervals (up to 900 days after the Izmit earthquake). Hence, this is a measure of how well the model fits the GPS velocity data throughout the entire 900-day time interval, not just a measure of misfit to a snapshot of the displacement over 900 days: WRSS ¼ X900 t¼1 X 55 s¼1 X 2 d¼1 D 2 t;s;d ð1þ s t;s;d D t,s,d is the residual (GPS minus modeled) velocity, and s t,s,d is the GPS velocity error (that is, the 68% confidence interval). Subscripts t, s, and d refer to time, GPS site, and degree of freedom (east or north). In some cases, GPS sites were occupied for just part of the 900-day interval. The WRSS and WRSS t include weighted-residual contributions only from days within the time span covered by GPS observations. Vertical GPS site position data are provided by Ergintav et al. [2009] in their auxiliary material Data Set S2, but these data are not converted to velocities. Using vertical velocities from a previous version of the Ergintav et al. analysis, we found that the vertical component contributed less than 10% to the WRSS and WRSS t values, and this contribution was insensitive to changes in model parameters. In the WRSS and WRSS t values presented here, only the horizontal velocity components are included. 4. Models 4.1. Velocity-Strengthening Frictional Afterslip [8] Velocity-strengthening frictional afterslip (FAS) [Marone et al., 1991] is implemented using the hot friction parameterization of Linker and Rice [1997], described also by Hearn et al. [2002a], Hearn [2003], and Johnson et al. [2006]. The equation for afterslip velocity is dt V ¼ V o exp ð3þ ða b V o is the preearthquake slip rate, (a b) is the empirical constant relating fault friction change to change in slip velocity, sn 0 is the effective normal stress, and dt is the timedependent, earthquake-induced shear stress resolved onto the fault surface. dt is t t o, where t is the shear stress on the fault and t o is the shear stress on the fault prior to the earthquake, when the fault is assumed to have been slipping at a rate of V o. The product of a b and sn, 0 called A-B, controls the increase in slip velocity due to a stress change. For positive A-B, aseismic afterslip may occur. For large, positive (A-B), a coseismic stress change has just a minimal effect on the rate of aseismic slip. In our models, velocitystrengthening afterslip is allowed at depths of 0 2 km and below 10 km, and values of parameter A-B for these intervals are sampled independently. We assume velocityweakening friction (and no afterslip) at depths between 2 and 10 km the upper crust [e.g., Scholz, 1998]. In all of the models, A-B is doubled in value (relative to the lower crustal value) below the Moho. Viscous shear zone creep takes over at depths greater than 25 km if the rate of frictional afterslip falls below the viscous fault zone creep rate (assuming h/w =10 15 Pa s m 1 ). Figure 2a schematically illustrates the model Viscous Fault Zone Creep [9] For the second class of afterslip models, we treat the fault zone at depths exceeding 10 km as a finite width zone Þs 0 n 3of21

4 Figure 3a. Slip velocity as a function of time on the Izmit and Düzce ruptures for frictional afterslip model FAS1. Time epochs are (from top to bottom) 60 days, 180 days, 1 year, and 2.5 years. Yellow dashed lines show high-slip patches from the Izmit and Düzce earthquakes [Reilinger et al., 2000]. Stars indicate the Izmit and Düzce earthquake hypocenters. The fault depth interval shown here is 0 to 36 km. with a Newtonian viscosity. The creep velocity V (i.e., the velocity of one side of the finite shear zone relative to the other) is V ¼ wdt h where h is viscosity and w is the horizontal width of the shear zone. Parameter h/w is varied in our models. V is relative to the rate at which the fault zone was creeping prior to the earthquake; h/w may increase or decrease with depth between 10 and 32 km (crust). The maximum crustal value is used at depths of 32+ km. The equation assumed for the variation of h/w with depth (z) in kilometers is ð4þ h w ¼ h exp½ð10 zþ=z char Š ð5þ w o where (h/w) o is (h/w) at10kmdepth Viscoelastic Relaxation of the Lower Crust [10] Models incorporating several configurations of Newtonian viscoelastic layers demonstrate that early postseismic deformation characteristic of large, strike-slip earthquakes is most compatible with relaxation of a thin, lower crustal layer [Hearn, 2003]. Hearn [2003] also find that a nonlinear rheology fits the temporal evolution of postseismic deformation better than a Newtonian rheology but that unusual parameters (not consistent with available laboratory-derived flow laws) are required. [11] Here, we model linear and nonlinear viscoelastic relaxation of a layer between 22 and 32 km depth (Figure 2c). In the nonlinear model, the effective viscosity is calculated for each model element as described by Hearn [2003]. The equation for effective viscosity of a model element is s n 1 pre ð Þ ¼ h pre ð6þ s pre þ s ðt;elemþ h eff t;elem 4of21

5 Figure 3b. Slip velocity as a function of time on the Izmit and Düzce ruptures from kinematic slip inversions. For fault geometry and inversion details, see Ergintav et al. [2009]. Time epochs are (from top to bottom) 60 days, 180 days, 1 year, and 2.5 years. Inverted yellow triangles show the ends of rupture segments; Y is Yalova, G is Golcuk, W and E are west and east Sapanca, K is Karadare, and D is Düzce. Inverted white triangles show the limits of the fault for Figures 3a, 3c, and 3d. Slip along a depth interval of 0 to 40 km is shown. For other details, see Figure 3a caption. where h eff (t,elem) and s (t,elem) are the effective viscosity and the differential stress (coseismic plus postseismic) in the model element at time t after the earthquake and s pre is the preearthquake differential stress. Since s (t,elem) is calculated by the FE code prior to each time step and n = 3, the only free parameters are h pre and s pre. We can compare h pre with predictions of flow laws at lower crustal temperatures and s pre values to assess whether our required parameters are generally consistent with the rheology of crustal rocks. [12] This approach does not account for any strain perturbation resulting from the background tectonic stress acting on coseismically weakened viscoelastic material. We are modeling deformation assuming d _e ¼ ds ðt;elemþ ð7þ h eff ðt;elemþ d _e ds ðt;elem Þ ¼ 1 ð8þ h eff ðt;elemþ However, if we take the derivative of _e with respect to s, that is, s pre + s (t,elem), substituting equation (6) for h eff,we get d _e ds ¼ n ð9þ h eff ðt;elemþ If the coseismic stress change tensor and the preearthquake stress tensor have the same principal stress orientations, their differential stresses may be summed (so ds (t,elem) equals the total ds in equation (9)). In this case, d _e ¼ nds ðt;elemþ ð10þ h eff ðt;elemþ Equation (10) is just n times equation (7). If the principal stresses do not have the same orientation (as is usually the 5of21

6 Figure 3c. Slip velocity as a function of time on the Izmit and Düzce ruptures for viscous shear zone model VSZ1. Time epochs are (from top to bottom) 60 days, 180 days, 1 year, and 2.5 years. For other details, see Figure 3a caption. case), we cannot assume that ds (t,elem) = ds, and the difference between equations (7) and (10) is smaller. We provide ranges of inferred effective viscosities (assuming n = 3)in subsequent discussions of model results Dual-Process Models [13] Using our finite element model, we may model afterslip and viscoelastic relaxation with various rheologies simultaneously. We limit our explorations to models with velocity-strengthening frictional afterslip in the crust (and viscous creep on a fairly stiff lower crustal shear zone with h/w = Pa s m 1 ). Figure 2d shows a schematic of a dual-process model in which both afterslip and viscoelastic relaxation are modeled. Afterslip is modeled as described in section 4.1, and the lower crust and upper mantle are modeled as Newtonian viscoelastic layers. In the lower crust and upper mantle, afterslip and viscoelastic relaxation may occur simultaneously. We use friction parameters from the best FAS model (FAS1), and we vary viscosities in the lower crust and mantle layers. 5. Results 5.1. Velocity-Strengthening Frictional Afterslip Models Pattern and Rate of Afterslip [14] Figure 3a shows afterslip along and below the sections of the NAFZ that ruptured in the Izmit and Düzce earthquakes, due to FAS driven by coseismic stress. The maximum slip rate two months after the Izmit earthquake is about 1.5 m a 1. Afterslip on the rupture surface falls between areas of high coseismic slip, as expected from the initial slip distribution we use, which loads these sections of the fault. The total potency of afterslip as a function of time for the best FAS model (FAS1) at depths less than 2 km is similar to kinematic slip inversion results (Figure 3b). (The afterslip rate for 180 days shown on Figure 3b is from Ergintav et al. [2009]. The slip solutions for other time epochs were done by Hearn for Ergintav et al. [2009] but 6of21

7 Figure 3d. Slip velocity as a function of time on the Izmit and Düzce ruptures for viscous shear zone model VSZ2. Time epochs are (from top to bottom) 60 days, 180 days, 1 year, and 2.5 years. For other details, see Figure 3a caption. were not ultimately included.) However, FAS1 yields more localized patches of shallow slip in different locations. This could be due to the coarse representation of the coseismic slip distribution in our FE model, and by smoothing of slip (and the use of 4-km fault patches) in our kinematic inversion. Two areas of deeper slip are indicated by the FAS and the kinematic model results. The high-slip areas are shifted to the east in the kinematic slip inversion relative to the FAS models, possibly due to the inversion routine s effort to fit velocities at GPS sites southeast of the Marmara Sea, which have anomalously large southward velocities (e.g., FIS1 and CINA). [15] Model FAS1 cannot reproduce deep slip at the rates required by the kinematic inversions. At 180 days, the kinematic slip inversion requires a maximum afterslip rate of about 800 mm a 1, whereas FAS1 can only yield about 500 mm a 1 (Figures 3a and 3b). By 900 days after the Izmit earthquake, the FAS1 model yields perhaps one third the kinematically required deep afterslip rate. Deep afterslip triggered by the Düzce event is greater in FAS1 and the viscous fault zone creep (VSZ) models than indicated by the kinematic slip inversions. This may be because we used a coseismic slip distribution for a dipping surface [Bürgmann et al., 2002a] but modeled the fault as vertical Parameter Values and Fit to GPS Velocity Data [16] Figure 4 shows the sensitivity of WRSS to parameter variation for the FAS models. (See section 3 and equations (1) and (2) for definitions of WRSS and WRSS t.) In Figure 4, the WRSS reflects just an 80-day postseismic interval, and the minimum misfit is for a model (FAS1) in which velocitystrengthening parameter A-B (defined in section 4.1) for both depth intervals is about 0.5 to 0.7 MPa. This is comparable to, to somewhat larger than, our previous estimate (0.43 MPa, [Hearn et al., 2002a]). If we use data from the first 900 days, the best fit to the GPS velocities is obtained with (A-B) = 0.5 to 1 MPa in the shallow upper crust and significantly lower (up to 0.3 MPa) at 10 to 30 km depth (shaded region on Figure 4). For the 900-day time span, the sensitivity of WRSS to variations in A-B is low; WRSS values vary by less than 10% over the A-B range shown on Figure 4, and the minimum WRSS is about All of the models reduce the WRSS by about 65% relative to a model with no site motion, probably because owing to the form of equation (3), all of the FAS models yield very slow afterslip and surface velocities after the first few months. The FAS models (and all of the models presented in this paper) explain east west (fault-parallel) motion far better than north south motion. The largest reduction to east-wrss by any of the frictional 7of21

8 Figure 4. Weighted-residual sum of squares for the FAS models. The quantity shown, WRSS, is the sum of the squared weighted residuals at all sites, for horizontal velocity components, for 80 1-day intervals after the Izmit earthquake (see equation (1) and section 3). This value has been scaled by The shaded area shows where the minimum WRSS total is located on a similar plot (WRSS < ), for a postseismic time interval of 900 days. Dots indicate sets of parameter values for which models were run. afterslip models (for the first 80 days) is 72%, but none of the models fits the north south velocities better than a model with no site motion (that is, the north velocity residuals are larger than the north velocities). For the 900-day cumulative WRSS, model FAS1 fits the north component slightly better, reducing the north component of the WRSS by up to 36%. [17] Figure 5 shows snapshots of modeled and GPS velocities, as well as their difference (residuals) 2.5 years after the Izmit earthquake. Similar plots for 1 year after the earthquake are shown in auxiliary material Figure S1. 1 Table 1 summarizes how well the FE models fit GPS data at three postseismic time epochs. In all cases, most of the contribution to the WRSS t is from continuous GPS sites, which have very small velocity errors, especially after the first few weeks. Forty-five days after the Izmit earthquake, the best FAS model (FAS1) yields a 63% reduction in WRSS t relative to a model with fixed GPS sites. After 1 year and 2.5 years, the WRSS t value is reduced by 75% and 62% relative to a model with fixed sites. For FAS1, the mean velocity is 1/2 to 2/3 of the mean GPS velocity. Smaller velocities in the optimum FAS model are an indication that the model vectors are misoriented relative to the GPS vectors. The WRSS minimization selects models with smaller velocity amplitudes to minimize the WRSS. Because of this, WRSS during the first 80 days is somewhat larger than in our earlier model [Hearn et al., 2002a]. If we use the (A-B) values from Hearn et al. [2002a] to model afterslip, we do slightly better for the first 80 days but the WRSS is larger over the full 900-day interval Viscous Shear Zone Creep Models Pattern and Rate of Afterslip [18] Dynamically modeled afterslip from two viscous shear zone (VSZ) models is shown on Figures 3c and 3d. 1 Auxiliary materials are available in the HTML. doi: / 2008JB As is the case for model FAS1, these models produce two main patches of afterslip, and both of these patches are west of where the kinematic model places them. In the best fitting solution where h/w increases with depth (VSZ2, Figure 3d), the slip rate patterns are similar to those from the frictional aferslip model, except that no slip is allowed in the top two kilometers of the crust. The decay in slip velocities with time is slower, however, and there is a progressive deepening of the afterslip which is less apparent in model FAS1. This difference is not enough to be resolved by the GPS velocities. In the best fitting solution where h/w decreases with depth (VSZ1, Figure 3c), afterslip is evenly distributed over a broad depth interval and is faster than for the FAS and VSZ2 models (and comparable to the kinematic slip velocities). The slipping zone narrows late in the simulation, and becomes shallower (especially below the Düzce hypocenter). The western slip patch also disappears between years 1 and Parameter Values and Fit to GPS Velocity Data [19] Figure 6 shows the sensitivity of WRSS to parameter variation for the VSZ models. The best fit to the GPS data is obtained in models where h/w either decreases or increases with depth (i.e., models VSZ1 and VSZ2). Most of the modest difference in fit between these models, and one in which h/w is constant with depth, arises from their superior performance in the later postseismic interval. A plot similar to Figure 6, covering just the first 180 days after the Izmit earthquake, shows a WRSS minimum in the same region, but with less of a preference for the models with depthvarying h/w. The east component of WRSS is reduced by up to 76%, and north component by up to 39%, relative to a model with no site motion. In model VSZ1, h/w increases (following equation (5)) from Pa s m 1 at 10 km depth to Pa s m 1 at 30 km depth. In model VSZ2, these values are Pa s m 1 and Pa s m 1, respectively. The best model with constant h/w requires a value of Pa s m 1. [20] Modeled and GPS velocities, and residuals 2.5 years after the Izmit earthquake, are shown on Figures 7a and 7b. Similar plots for 1 year after the earthquake are shown in auxiliary material Figures S2 and S3. Like model FAS1, neither of the VSZ models fits the rapid, early postseismic velocities at near-field GPS sites. However, given the larger GPS velocity errors immediately after the Izmit earthquake, the WRSS penalty for this misfit is small. At 45 days, model VSZ1 reduces WRSS t by a mere 25% relative to a model with fixed GPS sites. After 1 and 2.5 years, the model reduces WRSS t by 73% and 67%, respectively, relative to a model with fixed GPS sites. VSZ2 reduces WRSS t by 61%, 85%, and 76% 45 days, 1 year, and 2.5 years after the Izmit earthquake. The better performance at 45 days is due to inhibited deep slip (and smaller modeled velocities). Both VSZ models yield velocity amplitudes which are closer to the GPS velocities than those produced by model FAS Nonlinear Viscoelastic Relaxation Models [21] We have also run models to assess whether relaxation of nonlinear viscoelastic lower crust (with stress exponent n = 3) could explain Izmit postseismic deformation. Figure 8 shows sensitivity of WRSS to variations in the preearthquake differential stress and viscosity. We find that for a nonlinear lower crust model to fit the GPS data as well as 8of21

9 Figure 5. (top) Modeled and GPS velocity vectors 2.5 years after the Izmit earthquake and (bottom) residuals for frictional afterslip model FAS1. In Figure 5 (top), dark arrows are modeled velocities and light arrows are GPS velocities (with error ellipses showing one-sigma errors (68% uncertainties)). Vectors for 1 year after the earthquake are shown in auxiliary material Figure S1. Site ANKR (inset) is about 50 km east of the location shown on all of the surface velocity plots. Site KDER is very close to the NAF trace and its velocity is not modeled accurately when shallow afterslip is allowed (here and on Figure 11). the best afterslip models, h pre and s pre are about 2 to Pa s and 0.05 to 0.1 MPa. Previously [Hearn et al., 2002a], we made a rough estimate of s pre, based on equation (6), n = 3, a range of assumed, reasonable h pre values, and the effective viscosity required by our best linearly viscoelastic lower crust model to fit the first 80 days of postseismic deformation ( Pa s). That estimate was 0.08 to 1.6 MPa, which is reasonable given that actual coseismic stresses in the lower crust are lower than the 5 MPa value we assumed in the 2002 calculation [Hearn et al., 2002a]. [22] We note that assuming n = 3 allows us to bracket admissible values of h pre (section 4.3). To narrow this range further, and more precisely model the postseismic velocities, the postseismic deformation models should be embedded in a preexisting stress field [e.g., Freed et al., 2006b], ideally compatible with the rheology and the preearthquake deformation. [23] The north component of the WRSS is reduced by up to 45%, and the east component by up to 80%, relative to a model with fixed GPS sites. As with the afterslip models, none of the models reduces the north component of the WRSS significantly during the first 6 months. Forty-five days after the Izmit earthquake, the best nonlinear lower crust model (NL1) yields a 58% reduction in WRSS t relative to a model with fixed GPS sites (Table 1). At 1 year, this reduction is 79% and at 2.5 years, the NL1 model yields just a 52% reduction in the WRSS t. Figure 9 shows modeled and GPS site velocities, and residuals 2.5 years after the Izmit earthquake. Similar plots for 1 year after the earthquake are shown in auxiliary material Figure S4. At 1 and 2.5 years, modeled velocities exceed GPS site velocities at 22 and 33 sites, respectively, and the mean velocities are comparable to the mean GPS velocities (v/v o = 0.86 and 1.16). Discrepancies between modeled and GPS vector orientations are similar to those produced by 9of21

10 Table 1. Model Performance a Model Time Epoch WRSS t WRSS o WRSS Reduction (%) v/v o b DM b 45 days FAS1 45 days VSZ1 45 days VSZ2 45 days NL1 45 days DP1 45 days DM b 1 year FAS1 1 year VSZ1 1 year VSZ2 1 year NL1 1 year DP1 1 year DM b 2.5 years FAS1 2.5 years VSZ1 2.5 years VSZ2 2.5 years NL1 2.5 years DP1 2.5 years a The ratio v/v o is the ratio of mean modeled velocity amplitude to mean GPS velocity amplitude. WRSS o is the WRSS for a model in which all GPS sites are stationary. b For the elastic dislocation models [Ergintav et al., 2009], data from 68 GPS sites (rather than 55) were used. Hence, the WRSS o values are greater. Only strike-slip motion was allowed in these dislocation models (for more meaningful comparison with FE results). Including dip slip reduced the DM s WRSS by an additional 3% at 1 year and 0.5% at 2.5 years. the afterslip models, but the larger velocities lead to larger residuals (and WRSS t values) Dual-Process Models: Afterslip Plus Viscoelastic Relaxation [24] Figure 4 and Table 1 show that the FAS1 model does not yield rapid enough postseismic velocities in the vicinity of the Izmit rupture. The model s performance at later time epochs may be improved by supplementing FAS with relaxation of viscoelastic lower crust (or with viscous creep in a vertical, lower crustal shear zone with h/w lower than Pa s m 1 ) and viscoelastic relaxation of the upper mantle. Here, we present results for a model incorporating Newtonian viscoelastic relaxation of lower crust and upper mantle layers, with viscous shear zone h/w fixed at Pa s m 1.Inthis model, frictional afterslip and viscous shear zone creep are both modeled along the NAFZ, and below a depth of 25 km, the process yielding the greater slip rate over each time step is assumed to occur. Figure 10 shows the sensitivity of WRSS to mantle and lower crust viscosities for our dual-process (DP) models. The DP models outperform model FAS1 for a restricted range of mantle and crust viscosities (shaded region on Figure 10). The minimum permissible mantle or lower crust viscosity is about 2 to Pa s: for lower viscosity values, total WRSS increases dramatically. In models with the lowest mantle viscosities, a stiffer lower crust is required to fit the GPS velocity data, and for models with the lowest lower crust viscosities, a stiffer upper mantle is required. A model in which both the mantle and lower crust viscosities are both about Pa s also performs well. Hence, while our models provide firm lower bounds for the effective viscosities of the uppermost mantle and the lower crust, we cannot distinguish between the jelly sandwich and the creme brulee models of lithosphere viscoelastic structure [e.g., Burov and Watts, 2006]. [25] Wenoteherethatifweassumealowerh/w ( Pa s m 1 ) for the viscous shear zone, and we allow viscous creep to take over when frictional afterslip slows down below a depth of 15 km (instead of at 25 km), higher mantle and lower crust viscosities are required. For a model with identical mantle and lower crust effective viscosities, an effective viscosity value of Pa s is required. [26] Figure 11 illustrates that, as is the case with all of our models, the east velocity component is fit much better than the north velocity component. (This is true at all time epochs.) For the entire 900-day period, the east WRSS is reduced by up to 83%, while the north WRSS is reduced by only up to 45%. For the best model, DP1, WRSS is , which corresponds with an RMS error of 60. This error is large because of the very small formal uncertainties for the GPS postseismic velocities [Ergintav et al., 2009]. For example, after 1 and 2.5 years, virtually all of the GPS one-sigma velocity errors (at both CGPS and SGPS sites) are less than 1.5 and 0.5 mm a 1, respectively. (If we included the vertical velocity component, with its larger uncertainties, the RMS error would be about 30% lower.) [27] Adding viscoelastic mantle and lower crust relaxation increases the magnitude of modeled, late postseismic velocities to values approaching the GPS values. This viscoelastic relaxation only modestly reduces WRSS t and WRSS values, because as the modeled velocities approach parity with the GPS velocities, angular misfits cause larger residuals (Table 1). For the best DP model (DP1), WRSS t is 63% smaller than for a model with no site motion, 45 days after the earthquake. After 1 year, the WRSS t is 84% lower, which is comparable to the performance of the best dislocation model. However, v/v o is somewhat smaller than it is for the dislocation model (0.77 versus 0.86). After 2.5 years, the DP1 model s WRSS t is comparable to that of the FAS1 model for that time epoch. Unlike the FAS1 model, the DP model Figure 6. Weighted-residual sum of squares for the VSZ models. Models with depth-varying h/w perform better than models with uniform h/w. The quantity shown, WRSS, is the sum of the squared weighted residuals at all sites for both horizontal velocity components, for day intervals spanning 2.5 years after the Izmit earthquake (see equation (1) and section 3). This value has been scaled by Dots indicate sets of parameter values for which models were run. 10 of 21

11 Figure 7a. (top) Modeled and GPS velocity vectors 2.5 years after the Izmit earthquake and (bottom) residuals for viscous shear zone model VSZ1. In Figure 7b (top), dark arrows are modeled velocities, and light arrows are GPS velocities (with error ellipses showing one-sigma errors (68% uncertainties)). Vectors for 1 year after the earthquake are shown in auxiliary material Figure S2. produces a mean, modeled GPS site velocity amplitude at 2.5 years which is close to the GPS value (v/v o = 1.05), and modeled velocity amplitudes exceed GPS velocities at 32 of the 55 sites. 6. Discussion 6.1. Postseismic Deformation: Afterslip Plus Viscoelastic Relaxation [28] The earliest Izmit postseismic deformation requires afterslip at seismogenic depths [Bürgmann et al., 2002b; Hearn et al., 2002a]; an extremely rapidly decaying postseismic transient component (with a characteristic decay time of 1 day), and a very fast initial afterslip rate [Ergintav et al., 2009; Cakir et al., 2003]. These constraints are consistent with velocity-strengthening frictional afterslip along the NAFZ. Shallow, stress-driven afterslip which is indicated by model FAS1 is also required by kinematic slip inversions [Ergintav et al., 2009]. Furthermore, postseismic slip at the Earth s surface was identified in Golcuk after the Izmit earthquake and along the Düzce rupture after that earthquake (O. Emre, personal communication, 2006), and the amplitude of this slip was roughly consistent with the kinematic inversions [Ergintav et al., 2009]. A combination of frictional afterslip and viscous fault zone creep at depth [e.g., Mehl and Hirth, 2008] is likely at work. [29] Model VSZ1 shows that coseismic stress on the NAFZ was sufficient to drive afterslip at a fast enough rate to explain postseismic deformation for perhaps a year after the Izmit earthquake. However, none of the afterslip models can explain the deformation rates beyond about the first year. This suggests a second process, likely viscoelastic deformation of the upper mantle. Our dual-process models with mantle and lower crust viscosities of the order of to Pa s provide the best fit to the first 2.5 years of postseismic deformation. If a low-h/w viscous shear zone accommodates postseismic strain in the lower crust, rather than a relaxing viscoelastic layer, the mantle effective viscosity will be at the lower end of this range. Gravity surveys [Ergintav et al., 2007] also support a deformation contribution from viscoelastic relaxation in or below the lower crust. These data, spanning 2003 through 2005, suggest increasing gravity south of the NAFZ and decreasing gravity north of the NAFZ, along a profile crossing the eastern Marmara 11 of 21

12 Figure 7b. (top) Modeled and GPS velocity vectors 2.5 years after the Izmit earthquake and (bottom) residuals for viscous shear zone model VSZ2. In Figure 7 (top), dark arrows are modeled velocities, and light arrows are GPS velocities (with error ellipses showing one-sigma errors (68% uncertainties)). Vectors for 1 year after the earthquake are shown in auxiliary material Figure S3. Sea. This is opposite the expected coseismic polarity and is consistent with relaxation of a lower crust or upper mantle layer [Hearn, 2003] but not afterslip or relaxation of a midcrustal detachment. (Afterslip on a south dipping, roughly NAFZ-parallel normal fault such as the Princes Island Fault might also cause this signature.) [30] The only alternative to a multiprocess model is one in which most of the rapid postseismic deformation is due to temporarily low effective viscosity in the lower crust or upper mantle around the Izmit rupture, for example, due to a power law rheology. This has been proposed for Mojave Desert earthquakes [e.g., Freed and Bürgmann, 2004]. We were able to devise a nonlinear lower crust model with n = 3 that fit the earliest Izmit-Düzce postseismic deformation about as well as the afterslip models (after the first week). However, the low preearthquake stresses and strain rates required by this model are problematic. [31] Figures 12a and 12b show the temperatures at which flow laws representative of crustal rocks would yield the required h pre (2 to Pa s) at the required s pre of 0.1 MPa. To generate Figure 12, we calculated effective viscosity for a randomly sampled set of synthetic flow laws with various values of A and Q, assuming that 2.5 < n < 3.5, s pre = 0.1 MPa, and T = 300 to 1200 C. Real flow laws for quartzite and crustal rocks of various types are plotted on Figures 12a and 12b for reference. Heat flux in the vicinity of the Izmit rupture is about 60 mw m 2 [Schindler, 1997], suggesting lower crust temperatures of 450 to 600 C. In this temperature range, some wet quartz flow laws yield the expected early postseismic effective viscosities. However, the required preearthquake stress is so low that linear diffusion creep, rather than nonlinear dislocation creep, should dominate. Another problem with the NL1 model is that much higher effective viscosities (at least Pa s) are required by surface deformation rates later in the earthquake cycle (as described in section 6.3). Finally, for the required interseismic differential stress of about 0.05 to 0.1 MPa, the strain rate in the lower crust must be about s 1. This is too low to be consistent with deformation around a major plate boundary fault (or with the highly localized surface strain around the NAFZ throughout the earthquake cycle) unless a rheologically distinct, very low viscosity shear zone below the NAFZ controls levels of differential stress [e.g., Mehl and Hirth, 2008]. In this case, however, the shear zone would deform via 12 of 21

13 Figure 8. Weighted-residual sum of squares for the nonlinear lower crust models. The quantity shown, WRSS, is the sum of the squared weighted residuals at all sites for both horizontal velocity components for day intervals spanning 2.5 years after the Izmit earthquake (see equation (1) and section 3). This value has been scaled by Dots indicate sets of parameter values for which models were run. diffusion creep, due to grain size reduction [e.g., Mehl and Hirth, 2008; Warren and Hirth, 2006]. Unless the grain size could adjust to the coseismic stress increment by further decreasing over postseismic timescales, the situation would reduce to a VSZ model Comparison With Other Postseismic Deformation Studies [32] We require frictional afterslip, viscous shear zone creep or viscoelastic relaxation of the lower crust, and viscoelastic relaxation of the upper mantle to explain the first 2.5 years of Izmit earthquake postseismic deformation. We have not examined all of the possibilities, for example, more combinations of FAS and VSZ creep (with lower h/w), and lateral heterogeneities, such as shear zones, in the upper mantle [e.g., Warren and Hirth, 2006]. Such lateral heterogeneities, or a transient rheology, or both, are required to make the observed postseismic deformation consistent with highly localized strain around NAFZ segments later in the earthquake cycle. [33] Afterslip has been identified following earthquakes in several tectonic settings, including subduction zones [e.g., Heki and Tamura, 1997; Melbourne et al., 2002; Miyazaki and Larson, 2008]; and along thrust- and strikeslip faults within continental crust [e.g., Hsu et al., 2002; Podgorski et al., 2007]. In several cases, afterslip has been modeled as velocity-strengthening frictional slip [Perfettini and Avouac, 2004, 2007; Johnson et al., 2006, 2008, and references therein]. These studies give (A-B) values of 0.2 to 0.7 MPa, which are consistent with our results, and which likely indicate elevated pore pressures in the fault zone. Other studies, while not explicitly modeling frictional afterslip, show that afterslip occurs near the surface, or within the seismogenic zone between patches of high coseismic slip (where the fault was coseismically loaded, but is not hot enough for significant viscous creep [e.g., Podgorski et al., 2007; Miyazaki and Larson, 2008]). Given the near ubiquity of rapid afterslip following major earthquakes, it would be odd if this were not the cause of the earliest post-izmit deformation. [34] All of our models of stress-driven afterslip prior to the Düzce earthquake indicate a dominant patch of early afterslip near the Izmit hypocenter, and the maximum magnitude of afterslip prior to the Düzce event is about 2 m (depending on the model). The location and magnitude of this afterslip are consistent with kinematic afterslip results from Cakir et al. [2003], obtained from interferometric synthetic aperture radar (InSAR) surface displacements from the first month after the Izmit earthquake. Our FE models suggest another patch of afterslip approximately centered between the Izmit and Düzce hypocenters, which correlates with a patch of afterslip centered about 50 km east of the Izmit hypocenter, in Figure 10 of Cakir et al. [2003]. Our stress-driven afterslip models and the Cakir et al. [2003] inversions place the two afterslip patches to the west of their locations inferred from inversions of GPS data [Ergintav et al., 2009; Bürgmann et al., 2002a]. Our early afterslip results differ somewhat from what we presented in our previous study [Hearn et al., 2002a] because with the new GPS analysis and the higher-resolution mesh, we can now distinguish two patches of slip, rather than a continuous zone of afterslip extending the length of the coseismic rupture. Our VSZ models, and the kinematic slip inversions from Ergintav et al. [2009] also suggest some pre- Düzce earthquake afterslip along or below the future Düzce rupture. This is not noted by Cakir et al. because the Düzce rupture is beyond the east limit of the region covered by their interferograms. [35] Modeling studies of postseismic deformation following major earthquakes in other parts of the world also show that multiple processes contribute to postseismic deformation, and that afterslip (sometimes with other shallow processes) is likely the cause of the earliest deformation. For example, Johnson et al. [2008] find that frictional afterslip and Newtonian viscoelastic relaxation of upper mantle below a depth of 45 to 85 km (with h eff = Pa s) explain the first 4 years postseismic deformation following the 2002 Denali, Alaska earthquake. Freed et al. [2006a] explain Denali postseismic surface displacements in terms of afterslip, viscoelastic relaxation or fault zone creep in the lower crust, poroelastic rebound, and viscoelastic relaxation of the upper mantle. Freed et al. [2006a] find a mantle viscosity at the Moho (>10 19 Pa s at 50 km, decreasing with depth) which is comparable to what we see, based on 2.5 years of postseismic deformation, in Anatolia. They also find, as we do, that the postseismic slip distribution is anticorrelated with coseismic slip, and that coseismic stresses along the fault cannot drive enough afterslip to explain all of the observed postseismic deformation. In a follow-up to their 2006 Denali paper, Freed et al. [2006b] conclude that the mantle beneath the Denali rupture cannot be Newtonian, and we summarize below why we arrive at the same conclusion for central Anatolia Interseismic Deformation Rules Out Maxwell Viscoelastic Mantle [36] Moderate viscosities in the Anatolian upper mantle (i.e., less than Pa s) are consistent with marked attenuation of regional seismic phases (particularly Sn, which 13 of 21

14 Figure 9. (top) Modeled and GPS velocity vectors 2.5 years after the Izmit earthquake and (bottom) residuals for nonlinear viscoelastic lower crust model NL1. In Figure 9 (top), dark arrows are modeled velocities, and light arrows are GPS velocities (with error ellipses showing one-sigma errors (68% uncertainties)). Vectors for 1 year after the earthquake are shown in auxiliary material Figure S4. broadcasts upper mantle properties), and slow V p in the mantle (7.7 km s 1 ) in central Turkey [Sandvol et al., 2001; Gok et al., 2000; Hearn and Ni, 1994]. (Viscosity of the lower crust is not as well constrained by geophysical methods.) Although some models of regional-scale secular deformation in the eastern Mediterranean region require a high, vertically averaged lithosphere viscosity consistent with block-like behavior over hundreds to thousands of years [Jiménez-Munt and Sabadini, 2002; Hubert-Ferrari et al., 2003; Fischer, 2006], others suggest that block-like deformation could be compatible with a moderate to low (10 20 to Pa s) viscosity below the brittle-ductile transition (see Fischer [2006], Anatolian block only, and Provost et al. [2003]) if shear stresses are extremely low along the blockbounding faults. Upper mantle effective viscosity values required by models of postseismic deformation in other parts of the world cluster around to Pa s [Bürgmann and Dresen, 2008]. [37] If the mantle viscosity is moderate, as our postseismic models suggest, and linearly viscoelastic, strain around the NAFZ segments should depend strongly on the time since the last earthquake [Savage, 1990]. Earthquake cycle models incorporating Maxwell viscoelastic material below the brittle-ductile transition produce geodetically observable variations in the pattern of strain around the NAFZ at different times in the earthquake cycle, unless the viscosity is high (at least Pa s [Hearn et al., 2002b]). However, GPS observations show that strain concentration around NAFZ segments is insensitive to the elapsed time since the last great earthquake [e.g., Ayhan et al., 2002; Reilinger et al., 2006]. InSAR studies also show that strain is tightly localized around NAFZ rupture segments that have not experienced major earthquakes for over 60 years [e.g., Wright et al., 2001]. [38] The simplest way around this problem is to model an increase in effective viscosity with time between major earthquakes, due to either a power law or a transient rheology. Nonlinear flow laws for crust and mantle rocks at the temperatures we expect in the lower crust and upper mantle can theoretically produce an effective viscosity of the order of 2 to Pa s within a range of stress levels exceeding the coseismic stress change (Figures 13a and 13b). If the power 14 of 21

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

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

The Effect of Elastic Layering on Inversions of GPS Data for Coseismic Slip and Resulting Stress Changes: Strike-Slip Earthquakes

The Effect of Elastic Layering on Inversions of GPS Data for Coseismic Slip and Resulting Stress Changes: Strike-Slip Earthquakes Bulletin of the Seismological Society of America, Vol. 95, No. 5, pp. 1637 1653, October 2005, doi: 10.1785/0120040158 The Effect of Elastic Layering on Inversions of GPS Data for Coseismic Slip and Resulting

More information

Seven years of postseismic deformation following the 1999, M = 7.4 and M = 7.2, Izmit-Düzce, Turkey earthquake sequence

Seven years of postseismic deformation following the 1999, M = 7.4 and M = 7.2, Izmit-Düzce, Turkey earthquake sequence JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 114,, doi:10.1029/2008jb006021, 2009 Seven years of postseismic deformation following the 1999, M = 7.4 and M = 7.2, Izmit-Düzce, Turkey earthquake sequence S. Ergintav,

More information

Afterslip and viscoelastic relaxation following the 1999 M 7.4 İzmit earthquake from GPS measurements

Afterslip and viscoelastic relaxation following the 1999 M 7.4 İzmit earthquake from GPS measurements Geophys. J. Int. (2009) doi: 10.1111/j.1365-246X.2009.04228.x Afterslip and viscoelastic relaxation following the 1999 M 7.4 İzmit earthquake from GPS measurements L. Wang, 1,2 R. Wang, 1 F. Roth, 1 B.

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

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

Using short-term postseismic displacements to infer the ambient deformation conditions of the upper mantle

Using short-term postseismic displacements to infer the ambient deformation conditions of the upper mantle JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 117,, doi:10.1029/2011jb008562, 2012 Using short-term postseismic displacements to infer the ambient deformation conditions of the upper mantle Andrew M. Freed, 1

More information

Frictional Properties on the San Andreas Fault near Parkfield, California, Inferred from Models of Afterslip following the 2004 Earthquake

Frictional Properties on the San Andreas Fault near Parkfield, California, Inferred from Models of Afterslip following the 2004 Earthquake Bulletin of the Seismological Society of America, Vol. 96, No. 4B, pp. S321 S338, September 2006, doi: 10.1785/0120050808 Frictional Properties on the San Andreas Fault near Parkfield, California, Inferred

More information

Time Dependence of Postseismic Creep Following Two Strike-Slip Earthquakes. Gerasimos Michalitsianos

Time Dependence of Postseismic Creep Following Two Strike-Slip Earthquakes. Gerasimos Michalitsianos Time Dependence of Postseismic Creep Following Two Strike-Slip Earthquakes Gerasimos Michalitsianos 9 April 011, GEOL394 Thesis Advisor: Laurent Montési 1 Table of Contents I. Introduction 4 II. Overview

More information

Afterslip, slow earthquakes and aftershocks: Modeling using the rate & state friction law

Afterslip, slow earthquakes and aftershocks: Modeling using the rate & state friction law Afterslip, slow earthquakes and aftershocks: Modeling using the rate & state friction law Agnès Helmstetter (LGIT Grenoble) and Bruce Shaw (LDE0 Columbia Univ) Days after Nias earthquake Cumulative number

More information

Seven years of postseismic deformation following the 1999, M = 7.4 and M = 7.2, Izmit-Düzce, Turkey earthquake sequence

Seven years of postseismic deformation following the 1999, M = 7.4 and M = 7.2, Izmit-Düzce, Turkey earthquake sequence Seven years of postseismic deformation following the 1999, M = 7.4 and M = 7.2, Izmit-Düzce, Turkey earthquake sequence The MIT Faculty has made this article openly available. Please share how this access

More information

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

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

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

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

Coupled afterslip and viscoelastic flow following the 2002

Coupled afterslip and viscoelastic flow following the 2002 submitted to Geophys. J. Int. Coupled afterslip and viscoelastic flow following the Denali Fault, Alaska Earthquake Kaj M. Johnson, Roland Bürgmann, Jeffrey T. Freymueller Department of Geological Sciences,

More information

On the effects of thermally weakened ductile shear zones on postseismic deformation

On the effects of thermally weakened ductile shear zones on postseismic deformation JOURNAL OF GEOPHYSICAL RESEARCH: SOLID EARTH, VOL. 118, 1 16, doi:10.1002/2013jb010215, 2013 On the effects of thermally weakened ductile shear zones on postseismic deformation Christopher S. Takeuchi

More information

Evidence for postseismic deformation of the lower crust following the 2004 Mw6.0 Parkfield earthquake

Evidence for postseismic deformation of the lower crust following the 2004 Mw6.0 Parkfield earthquake JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 116,, doi:10.1029/2010jb008073, 2011 Evidence for postseismic deformation of the lower crust following the 2004 Mw6.0 Parkfield earthquake Lucile Bruhat, 1,2 Sylvain

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

Controls of shear zone rheology and tectonic loading on postseismic creep

Controls of shear zone rheology and tectonic loading on postseismic creep JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 109,, doi:10.1029/2003jb002925, 2004 Controls of shear zone rheology and tectonic loading on postseismic creep Laurent G. J. Montési Woods Hole Oceanographic Institution,

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

Variability of earthquake nucleation in continuum models of rate-and-state faults and implications for aftershock rates

Variability of earthquake nucleation in continuum models of rate-and-state faults and implications for aftershock rates Click Here for Full Article JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 113,, doi:10.1029/2007jb005154, 2008 Variability of earthquake nucleation in continuum models of rate-and-state faults and implications

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

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

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 114, B07405, doi: /2008jb005748, 2009

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 114, B07405, doi: /2008jb005748, 2009 Click Here for Full Article JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 114,, doi:10.1029/2008jb005748, 2009 Postseismic deformation due to the M w 6.0 2004 Parkfield earthquake: Stress-driven creep on a fault

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

Post-seismic motion following the 1997 Manyi (Tibet) earthquake: InSAR observations and modelling

Post-seismic motion following the 1997 Manyi (Tibet) earthquake: InSAR observations and modelling Geophys. J. Int. (7) 69, 9 7 doi:./j.365-46x.6.33.x Post-seismic motion following the 997 Manyi (Tibet) earthquake: InSAR observations and modelling Isabelle Ryder, Barry Parsons, Tim J. Wright and Gareth

More information

Three-dimensional viscoelastic finite element model for postseismic deformation of the great 1960 Chile earthquake

Three-dimensional viscoelastic finite element model for postseismic deformation of the great 1960 Chile earthquake JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 109,, doi:10.1029/2004jb003163, 2004 Three-dimensional viscoelastic finite element model for postseismic deformation of the great 1960 Chile earthquake Y. Hu, 1 K.

More information

Geophysical Journal International

Geophysical Journal International Geophysical Journal International Geophys. J. Int. (2010) 182, 1124 1140 doi: 10.1111/j.1365-246X.2010.04678.x A unified continuum representation of post-seismic relaxation mechanisms: semi-analytic models

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

Earthquake cycle simulations with rate-and-state friction and power-law viscoelasticity

Earthquake cycle simulations with rate-and-state friction and power-law viscoelasticity 1 2 Earthquake cycle simulations with rate-and-state friction and power-law viscoelasticity 3 4 5 6 7 Kali L. Allison a,1 and Eric M. Dunham a,b a Department of Geophysics, Stanford University, Stanford,

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

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

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

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

SUPPLEMENTARY INFORMATION

SUPPLEMENTARY INFORMATION doi: 1.138/nature962 Data processing A network of 5 continuously recording GPS stations (LAGU, CHIN, ENAP, GUAD and JUAN) was installed after the earthquake (Figure 1, main text). The data considered in

More information

Secondary Project Proposal

Secondary Project Proposal Secondary Project Proposal Post-seismic deformation of Chi-chi earthquake Yunyue (Elita) Li 11:, Wednesday, June 2, 21 Li 2 Secondary project proposal Personal prospective MOTIVATION My interests for earthquake

More information

Originally published as:

Originally published as: Originally published as: Lorenzo Martín, F., Wang, R., Roth, F. (2002): The effect of input parameters on visco-elastic models of crustal deformation. - Física de la Tierra, 14, 33-54 The effect of input

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

INGV. Giuseppe Pezzo. Istituto Nazionale di Geofisica e Vulcanologia, CNT, Roma. Sessione 1.1: Terremoti e le loro faglie

INGV. Giuseppe Pezzo. Istituto Nazionale di Geofisica e Vulcanologia, CNT, Roma. Sessione 1.1: Terremoti e le loro faglie Giuseppe Pezzo Istituto Nazionale di Geofisica e Vulcanologia, CNT, Roma giuseppe.pezzo@ingv.it The study of surface deformation is one of the most important topics to improve the knowledge of the deep

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

Depth variation of coseismic stress drop explains bimodal earthquake magnitude-frequency distribution

Depth variation of coseismic stress drop explains bimodal earthquake magnitude-frequency distribution Click Here for Full Article GEOPHYSICAL RESEARCH LETTERS, VOL. 35, L24301, doi:10.1029/2008gl036249, 2008 Depth variation of coseismic stress drop explains bimodal earthquake magnitude-frequency distribution

More information

The problem (1/2) GPS velocity fields in plate boundary zones are very smooth. What does this smoothness hide?

The problem (1/2) GPS velocity fields in plate boundary zones are very smooth. What does this smoothness hide? Block models The problem (1/2) GPS velocity fields in plate boundary zones are very smooth Figure from Tom Herring, MIT What does this smoothness hide? Continuous deformation? Rigid block motions, with

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

Journal of Geophysical Research (Solid Earth) Supporting Information for

Journal of Geophysical Research (Solid Earth) Supporting Information for Journal of Geophysical Research (Solid Earth) Supporting Information for Postseismic Relocking of the Subduction Megathrust Following the 2007 Pisco, Peru earthquake D.Remy (a), H.Perfettini (b), N.Cotte

More information

ON NEAR-FIELD GROUND MOTIONS OF NORMAL AND REVERSE FAULTS FROM VIEWPOINT OF DYNAMIC RUPTURE MODEL

ON NEAR-FIELD GROUND MOTIONS OF NORMAL AND REVERSE FAULTS FROM VIEWPOINT OF DYNAMIC RUPTURE MODEL 1 Best Practices in Physics-based Fault Rupture Models for Seismic Hazard Assessment of Nuclear ON NEAR-FIELD GROUND MOTIONS OF NORMAL AND REVERSE FAULTS FROM VIEWPOINT OF DYNAMIC RUPTURE MODEL Hideo AOCHI

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

Interseismic strain accumulation across the Manyi fault (Tibet) prior to the 1997 M w 7.6 earthquake

Interseismic strain accumulation across the Manyi fault (Tibet) prior to the 1997 M w 7.6 earthquake GEOPHYSICAL RESEARCH LETTERS, VOL. 38,, doi:10.1029/2011gl049762, 2011 Interseismic strain accumulation across the Manyi fault (Tibet) prior to the 1997 M w 7.6 earthquake M. A. Bell, 1 J. R. Elliott,

More information

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

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

More information

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 transfer and strain rate variations during the seismic cycle

Stress transfer and strain rate variations during the seismic cycle JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 19, B642, doi:1.129/23jb2917, 24 Stress transfer and strain rate variations during the seismic cycle H. Perfettini Institut de Recherche pour le Développement/Laboratoire

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

Today (Mon Feb 27): Key concepts are. (1) how to make an earthquake: what conditions must be met? (above and beyond the EOSC 110 version)

Today (Mon Feb 27): Key concepts are. (1) how to make an earthquake: what conditions must be met? (above and beyond the EOSC 110 version) Today (Mon Feb 27): Key concepts are (1) how to make an earthquake: what conditions must be met? (above and beyond the EOSC 110 version) (2) strain (matrix: cannot be represented as a scalar or a vector

More information

material would flow extremely slowly similarly to a brittle material. The shear zone

material would flow extremely slowly similarly to a brittle material. The shear zone GSA DATA REPOSITORY 21468 Hayman and Lavier Supplementary model description: Lavier et al. (213) showed that formation or reactivation of mixed mode fractures in ductile shear zones might generate variations

More information

Does Aftershock Duration Scale With Mainshock Size?

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

More information

Lecture 5. Rheology. Earth Structure (2 nd Edition), 2004 W.W. Norton & Co, New York Slide show by Ben van der Pluijm

Lecture 5. Rheology. Earth Structure (2 nd Edition), 2004 W.W. Norton & Co, New York Slide show by Ben van der Pluijm Lecture 5 Rheology Earth Structure (2 nd Edition), 2004 W.W. Norton & Co, New York Slide show by Ben van der Pluijm WW Norton; unless noted otherwise Rheology is... the study of deformation and flow of

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

Dynamic analysis. 1. Force and stress

Dynamic analysis. 1. Force and stress Dynamic analysis 1. Force and stress Dynamics is the part of structural geology that involves energy, force, stress, and strength. It's very important to distinguish dynamic concepts from kinematic ones.

More information

Numerical simulation of seismic cycles at a subduction zone with a laboratory-derived friction law

Numerical simulation of seismic cycles at a subduction zone with a laboratory-derived friction law Numerical simulation of seismic cycles at a subduction zone with a laboratory-derived friction law Naoyuki Kato (1), Kazuro Hirahara (2) and Mikio Iizuka (3) (1) Earthquake Research Institute, University

More information

Seismic and flexure constraints on lithospheric rheology and their dynamic implications

Seismic and flexure constraints on lithospheric rheology and their dynamic implications Seismic and flexure constraints on lithospheric rheology and their dynamic implications Shijie Zhong Dept. of Physics, University of Colorado Boulder, Colorado, USA Acknowledgement: A. B. Watts Dept. of

More information

Supplementary information on the West African margin

Supplementary information on the West African margin Huismans and Beaumont 1 Data repository Supplementary information on the West African margin Interpreted seismic cross-sections of the north Angolan to south Gabon west African passive margins 1-3, including

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

Materials and Methods The deformation within the process zone of a propagating fault can be modeled using an elastic approximation.

Materials and Methods The deformation within the process zone of a propagating fault can be modeled using an elastic approximation. Materials and Methods The deformation within the process zone of a propagating fault can be modeled using an elastic approximation. In the process zone, stress amplitudes are poorly determined and much

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

Afterslip and aftershocks in the rate-and-state friction law

Afterslip and aftershocks in the rate-and-state friction law Click Here for Full Article JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 114,, doi:10.1029/2007jb005077, 2009 Afterslip and aftershocks in the rate-and-state friction law Agnès Helmstetter 1 and Bruce E. Shaw

More information

Friction can increase with hold time. This happens through growth and increasing shear strength of contacts ( asperities ).

Friction can increase with hold time. This happens through growth and increasing shear strength of contacts ( asperities ). Friction can increase with hold time. This happens through growth and increasing shear strength of contacts ( asperities ). If sliding speeds up, the average lifespan of asperities decreases This means

More information

Heterogeneous Coulomb stress perturbation during earthquake cycles in a 3D rate-and-state fault model

Heterogeneous Coulomb stress perturbation during earthquake cycles in a 3D rate-and-state fault model Click Here for Full Article GEOPHYSICAL RESEARCH LETTERS, VOL. 35, L21306, doi:10.1029/2008gl035614, 2008 Heterogeneous Coulomb stress perturbation during earthquake cycles in a 3D rate-and-state fault

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

Y. Kaneko, 1 Y. Fialko, 1 D. T. Sandwell, 1 X. Tong, 1 and M. Furuya Introduction

Y. Kaneko, 1 Y. Fialko, 1 D. T. Sandwell, 1 X. Tong, 1 and M. Furuya Introduction JOURNAL OF GEOPHYSICAL RESEARCH: SOLID EARTH, VOL. 118, 316 331, doi:1.19/1jb9661, 13 Interseismic deformation and creep along the central section of the North Anatolian Fault (Turkey): InSAR observations

More information

Inverting geodetic time series with a principal component analysis-based inversion method

Inverting geodetic time series with a principal component analysis-based inversion method Click Here for Full Article JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 115,, doi:10.1029/2009jb006535, 2010 Inverting geodetic time series with a principal component analysis-based inversion method A. P. Kositsky

More information

Influence of anelastic surface layers on postseismic

Influence of anelastic surface layers on postseismic JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 105, NO. B2, PAGES 3151-3157, FEBRUARY 10, 2000 Influence of anelastic surface layers on postseismic thrust fault deformation Gregory A. Lyzenga Department of Physics,

More information

The Harvard community has made this article openly available. Please share how this access benefits you. Your story matters

The Harvard community has made this article openly available. Please share how this access benefits you. Your story matters Estimates of Seismic Potential in the Marmara Sea Region from Block Models of Secular Deformation Constrained by Global Positioning System Measurements The Harvard community has made this article openly

More information

COULOMB STRESS CHANGES DUE TO RECENT ACEH EARTHQUAKES

COULOMB STRESS CHANGES DUE TO RECENT ACEH EARTHQUAKES COULOMB STRESS CHANGES DUE TO RECENT ACEH EARTHQUAKES Madlazim Physics Department, Faculty Mathematics and Sciences of Surabaya State University (UNESA) Jl. Ketintang, Surabaya 60231, Indonesia. e-mail:

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

Postseismic deformation of the Andaman Islands following the 26 December, 2004 Great Sumatra-Andaman Earthquake

Postseismic deformation of the Andaman Islands following the 26 December, 2004 Great Sumatra-Andaman Earthquake GEOPHYSICAL RESEARCH LETTERS, VOL.???, XXXX, DOI:1.129/, Postseismic deformation of the Andaman Islands following the 26 December, 24 Great Sumatra-Andaman Earthquake J. Paul, 1 A. R. Lowry, 2 R. Bilham,

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

Constraints on fault and lithosphere rheology from the coseismic slip and postseismic afterslip of the 2006 M w 7.0 Mozambique earthquake

Constraints on fault and lithosphere rheology from the coseismic slip and postseismic afterslip of the 2006 M w 7.0 Mozambique earthquake JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 117,, doi:10.1029/2011jb008580, 2012 Constraints on fault and lithosphere rheology from the coseismic slip and postseismic afterslip of the 2006 M w 7.0 Mozambique

More information

3D MODELING OF EARTHQUAKE CYCLES OF THE XIANSHUIHE FAULT, SOUTHWESTERN CHINA

3D MODELING OF EARTHQUAKE CYCLES OF THE XIANSHUIHE FAULT, SOUTHWESTERN CHINA 3D MODELING OF EARTHQUAKE CYCLES OF THE XIANSHUIHE FAULT, SOUTHWESTERN CHINA Li Xiaofan MEE09177 Supervisor: Bunichiro Shibazaki ABSTRACT We perform 3D modeling of earthquake generation of the Xianshuihe

More information

Nonuniform prestress from prior earthquakes and the effect on dynamics of branched fault systems

Nonuniform prestress from prior earthquakes and the effect on dynamics of branched fault systems Click Here for Full Article JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 112,, doi:10.1029/2006jb004443, 2007 Nonuniform prestress from prior earthquakes and the effect on dynamics of branched fault systems Benchun

More information

Sendai Earthquake NE Japan March 11, Some explanatory slides Bob Stern, Dave Scholl, others updated March

Sendai Earthquake NE Japan March 11, Some explanatory slides Bob Stern, Dave Scholl, others updated March Sendai Earthquake NE Japan March 11, 2011 Some explanatory slides Bob Stern, Dave Scholl, others updated March 14 2011 Earth has 11 large plates and many more smaller ones. Plates are 100-200 km thick

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

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

Modeling the Thermal-Mechanical Behavior of Mid-Ocean Ridge Transform Faults

Modeling the Thermal-Mechanical Behavior of Mid-Ocean Ridge Transform Faults Excerpt from the Proceedings of the COMSOL Conference 2008 Boston Modeling the Thermal-Mechanical Behavior of Mid-Ocean Ridge Transform Faults Emily C Roland *1, Mark Behn,2 and Greg Hirth 3 1 MIT/WHOI

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

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

Internal Layers of the Earth

Internal Layers of the Earth Lecture #4 notes Geology 3950, Spring 2006; CR Stern Seismic waves, earthquake magnitudes and location, and internal earth structure (pages 28-95 in the 4 th edition and 28-32 and 50-106 in the 5 th edition)

More information

Fault friction parameters inferred from the early stages of afterslip following the 2003 Tokachi-oki earthquake

Fault friction parameters inferred from the early stages of afterslip following the 2003 Tokachi-oki earthquake Click Here for Full Article JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 114,, doi:10.1029/2008jb006166, 2009 Fault friction parameters inferred from the early stages of afterslip following the 2003 Tokachi-oki

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

Separating Tectonic, Magmatic, Hydrological, and Landslide Signals in GPS Measurements near Lake Tahoe, Nevada-California

Separating Tectonic, Magmatic, Hydrological, and Landslide Signals in GPS Measurements near Lake Tahoe, Nevada-California Separating Tectonic, Magmatic, Hydrological, and Landslide Signals in GPS Measurements near Lake Tahoe, Nevada-California Geoffrey Blewitt, Corné Kreemer, William C. Hammond, & Hans-Peter Plag NV Geodetic

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

Homogeneous vs. realistic heterogeneous material-properties in subduction zone models: Coseismic and postseismic deformation

Homogeneous vs. realistic heterogeneous material-properties in subduction zone models: Coseismic and postseismic deformation Homogeneous vs. realistic heterogeneous material-properties in subduction zone models: Coseismic and postseismic deformation T. Masterlark 1, C. DeMets 2, H.F. Wang 2, O. S nchez 3, and J. Stock 4 1 US

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

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

Earthquake distribution is not random: very narrow deforming zones (= plate boundaries) versus large areas with no earthquakes (= rigid plate

Earthquake distribution is not random: very narrow deforming zones (= plate boundaries) versus large areas with no earthquakes (= rigid plate Earthquake distribution is not random: very narrow deforming zones (= plate boundaries) versus large areas with no earthquakes (= rigid plate interiors) Tectonic plates and their boundaries today -- continents

More information

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

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

More information

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

Introduction to Displacement Modeling

Introduction to Displacement Modeling Introduction to Displacement Modeling Introduction Deformation on the Earth surface informs us about processes and material properties below surface Observation tools: GPS (static-dynamic) InSAR (static)

More information

Title: Three-dimensional mechanical models for the June 2000 earthquake sequence in the South Icelandic Seismic Zone

Title: Three-dimensional mechanical models for the June 2000 earthquake sequence in the South Icelandic Seismic Zone Elsevier Editorial System(tm) for Earth and Planetary Science Letters Manuscript Draft Manuscript Number: Title: Three-dimensional mechanical models for the June earthquake sequence in the South Icelandic

More information