Permeabilities Inferred from Poroelastic Modeling. of InSAR-Determined Land Surface Deformation in

Size: px
Start display at page:

Download "Permeabilities Inferred from Poroelastic Modeling. of InSAR-Determined Land Surface Deformation in"

Transcription

1 submitted to Geophys. J. Int. Aquifer and Shallow San Andreas Fault Permeabilities Inferred from Poroelastic Modeling of InSAR-Determined Land Surface Deformation in Coachella Valley, California Ravi Appana 1, Martin O. Saar 1, Beth A. Wisely 2, David A. Schmidt 2 1 Department of Geology and Geophysics, University of Minnesota, Twin Cities 55455, USA 2 Department of Geological Sciences, University of Oregon, Eugene 97403, USA SUMMARY The Coachella Valley groundwater aquifer in southern California, USA, is a major agricultural and municipal water resource. Groundwater overdraft has led to significant reductions of water levels and subsequent land subsidence in some areas of this alluvial basin. Recharge efforts in the Upper Coachella Valley have partially restored groundwater levels and presumably slowed long-term land subsidence. Interferometric Synthetic Aperture Radar (InSAR) between late-1993 through 2000 reveals differential uplift of land surface across two strands of the southern San Andreas Fault (the Garnet Hill and the Banning Faults) in the northwest of the basin, associated with episodes of artificial groundwater recharge at the White Water Recharge Facility. Using InSAR data of land-surface displacement as well as groundwater levels as constraints, we model pore-fluid pressure diffusion and groundwater flow across these two strands of the San Andreas Fault system. Numerical models yield estimates of subsurface hydraulic properties, representing averages over a few kilometers, useful for groundwater management. We constrain the horizontal permeability

2 2 Appana et al. of the White Water sub-basin (WWB) to m 2 k wwb x m 2 and the permeability anisotropy, a = (k z /k x ), of the White Water and the Garnet Hill sub-basins (GHB) to 0.04 a wwb and a ghb 0.016, respectively. In addition, our model suggests permeabilities of m 2 k ghf m 2 and k bf m 2 for the Garnet Hill Fault (GHF) and the Banning Fault (BF), respectively. These results indicate that both faults act as semi-permeable barriers to pore-fluid pressure diffusion and groundwater flow due to a low-permeability fault core. Preliminary calculations result in permeabilities of m 2 k ghf c m 2 for the Garnet Hill Fault core and k bf c m 2 for the Banning Fault core at depths 2 km. Therefore, our study provides estimates for the shallow San Andreas Fault zone permeability field which can serve as an upper limit for fault-zone permeabilities at greater depths and may thus help constrain conceptual and numerical models of active faulting in general and of San Andreas Fault slip dynamics in particular. Key words: Numerical Modeling Permeability Coachella Valley San Andreas Fault Land subsidence Groundwater overdraft. 1 INTRODUCTION Formation permeability is arguably the most important parameter in geological porous media flow as it largely determines fluid flow rates and varies over 15 orders of magnitude in natural settings (e.g., Freeze & Cherry 1979; Saar & Manga 2004; Ingebritsen & Manning 2010). In contrast, other porous media flow parameters, such as hydraulic head gradient, water density, or water viscosity vary much less (e.g., Saar & Manga corresponding author: appan002@umn.edu

3 Poroelastic Modeling in Coachella Valley, CA, USA ; Ingebritsen et al. 2006). In addition to the bulk permeability of the sediment matrix or geologic formations, faults in unconsolidated sediments, bounding or disecting basin aquifers, can act as flow barriers (e.g., Antonellini & Aydin 1994; Rawling et al. 2001; Eichhubl et al. 2005; Mayer et al. 2007), conduits, or combined conduit-barrier systems (e.g., Bredehoeft et al. 1992; Caine et al. 1996; Bense & Person 2006; Bense et al. 2008), depending on their permeability field. Therefore, the permeability of the faults can dictate local to regional groundwater flow paths and rates and potentially affect recharge to groundwater basins (e.g., Marler & Ge 2002; Folch & Mas-Pla 2008). Faults can thus strongly influence the subsurface hydraulic head distribution in faulted aquifer systems (e.g., Haneberg 1995; Mayer et al. 2007). As a result, determination of fault permeabilities is an important component of hydrogeologic investigations. Moreover, improved knowledge of subsurface hydraulic properties can aid with groundwater management, for example to prevent groundwater overdraft and its associated adverse effects such as groundwater storage reduction, land-surface subsidence, and water quality degradation (e.g., Galloway et al. 1999). While faults associated with aquifer systems can affect groundwater flow and storage processes, hydrogeologic conditions can conversely influence fault-slip behavior. For example, many plate-boundary faults, like the San Andreas Fault (SAF), have been interpreted to slip at lower shear stresses than expected for rocks and for faults with similar friction coefficients (e.g., Lachenbruch & Sass 1980; Zoback et al. 1987; Hickman 1991). One possible explanation for this weakness is existence of (large) pore-fluid over-pressures within the fault zones (e.g., Byerlee 1990, 1992; Rice 1992), resulting in the reduction of effective stresses causing fault weakening (e.g., Raleigh et al. 1976; Simpson et al. 1988; Gupta 1992; Saar & Manga 2003; Fulton & Saffer 2009). Furthermore, frictional heating during fault slip can lead to thermal pressurization of the fluids trapped in the low-permeability fault core and resultant reductions in effective stresses, causing dynamic fault weakening (Vredevoogd et al. 2007). Therefore, determining the

4 4 Appana et al. permeability field given by a fault-zone s architecture (dimensions of damage zone vs. fault core as discussed in Sec. 7) can provide insights into the role played by pore-fluid pressures during fault slip. In this paper, we present a case study illustrating the influence of low-permeability faults on the groundwater flow within an unconfined aquifer system in Coachella Valley, California, USA. The three branches of the southern SAF, the Garnet Hill Fault (GHF), the Banning Fault (BF), and the Mission Creek Fault (MCF), are sub-vertical, exhibit dextral strike-slip motion (Yule & Sieh 2003), and comprise the main tectonic features of the upper valley region (Fig. 1). We develop numerical models of pore-fluid pressure diffusion, groundwater recharge and flow, and associated poroelastic land-surface uplift in the Upper Coachella Valley. The models are constrained by both Interferometric Synthetic Aperture Radar (InSAR) data of land-surface uplift and groundwater level measurements in wells (Sec. 3). Our models provide estimates for both basin and fault permeabilities and thus have implications for groundwater and tectonic studies. Most estimates of fault-zone permeability come from laboratory experiments on fault samples from surface outcrops or drill cores (e.g., Chu et al. 1981; Morrow & Lockner 1994; Zhang et al. 1999; Faulkner 2004; Tsutsumi et al. 2004; Mizoguchi et al. 2008; Wibberley et al. 2008; Lockner et al. 2009). Here, we infer permeabilties at regional aquifer scales by modeling pore-fluid pressure changes across fault zones. Our approach yields a range of relatively shallow fault permeabilities, applicable only to basin depths or depths of 2 km at most. However, these values likely serve as upper bounds for deeper fault permeabilities and have thus implications for slip dynamics of faults in general and of the San Andreas Fault (SAF) in particular.

5 Poroelastic Modeling in Coachella Valley, CA, USA 5 2 BACKGROUND In the following sections we introduce the equations governing fluid flow through, and deformation of, porous media. We also provide background information on the hydrogeologic and poroelastic setting of the study region. 2.1 Fluid flow and poroelasticity Laminar, low-inertia fluid flow through porous media under unconfined aquifer conditions is governed by the here linearized Boussinesq equation, S h h 1 t (K h ) = Q s + α ( s), (1) t where S = S y + DS s is storativity with S y, D, and S s denoting specific yield, thickness, and specific storage of the aquifer, respectively. Furthermore, K is the hydraulic conductivity tensor, Q s is a fluid source or sink, and are the gradient and the divergence operators, respectively, α is the Biot-Willis coefficient defined as the ratio of increment in fluid content to volumetric strain while maintaining constant fluid pressure conditions (e.g., Wang 2000), s is the displacement vector (see Equation (10), defined for solid displacement), t ( s) is the temporal change in strain, and h 1 and h are the constant and spatially varying terms of the hydraulic head, h, respectively. Neglecting kinematic energy due to assumption of low groundwater flow velocities, hydraulic head is defined as h = p ρ f g + h z = h + h 1, (2) where h z is elevation head and pρ f 1 g 1 is pressure head with p representing porefluid pressure, ρ f fluid density, and g (vertical) acceleration due to Earth s gravity. Linearization implies that Equation (1) is only valid for small h /h 1 ratios, as given in many settings, including the basins in the study region (see Fig. 5 introduced in Sec. 4). Under steady-state conditions, Equation (1) simplifies to (K h ) = Q s, (3)

6 6 Appana et al. which is solved for h in our fluid-flow models. Darcy s law (Darcy 1886) is then applied to yield specific discharge or Darcy velocity, q, given as q = K h, (4) where the hydraulic conductivity tensor is defined as K = ρ f gk/η, (5) where k is the second-rank intrinsic permeability tensor and η is the fluid s dynamic viscosity. In our two-dimensional models, k is reduced to a horizontal, k x, and a vertical, k z, component (ignoring minor off-diagonal terms) which are related by an anisotropy parameter, a = k z /k x. (6) Under hydrostatic conditions, the pore-fluid pressure, p, is related to the hydraulic head, h, by p = ρ f gh. (7) Therefore, the steady-state groundwater flow or hydraulic head diffusion equation (Equation 3) can also be written in terms of the pore-fluid pressure gradient, p = ρ f g h, then denoting pore-fluid pressure diffusion. Hydraulic diffusivity is given by D u = hk S y = hρ fgk ηs y (8) in unconfined aquifers. Consequently, the subsurface pore-fluid pressure distribution is largely determined by the hydraulic conductivity or the permeability of the porous medium. The effective stress describes the relation between pore-fluid pressure, p, and the stress state as σ t = σ e + αpδ ij, (9)

7 Poroelastic Modeling in Coachella Valley, CA, USA 7 where σ t are the total stresses acting on the porous material, σ e are the effective stresses, and δ ij is the Kronecker delta. Furthermore, plane-strain theory relates changes in pore-fluid pressure to displacement of the porous medium matrix by Y Y 2(1 + ν) 2 s + s = α p, (10) 2(1 + ν)(1 2ν) where Y is the Young s modulus of the aquifer material (porous medium) and ν is its Poisson s ratio (Wang 2000). The numerical model introduced in Section 5, solves Equations (3), (9), and (10) and applies Equation (4) to determine fluid flow lines when desired. 2.2 Hydrogeologic setting The study region is located in the Upper Coachella Valley in southern California, USA (Fig. 1), an area enclosed between Little San Bernardino, San Bernardino, and San Jacinto Mountains. The southern San Andreas Fault (SAF) system branches into several strands in the upper valley region, dividing the basin into sub-basins and respective aquifers. Regional groundwater flow paths extend through the valley along its long-axis from the NW to the SE and drain into the Salton Sea. A much smaller but significant cross-valley component to groundwater flow also exists in the upper valley region due to artificial groundwater recharge at the White Water Recharge Facility as shown later. Runoff from the mountains in the NE, NW, and SW of the study region and baseflow from the San Gorgonio Pass form the main sources of natural groundwater recharge. Evapotranspiration, groundwater pumpage through wells, and outflow to the lower valley contribute to groundwater discharge (CVWD 2002). The sediments of the Upper Coachella Valley basin include unconsolidated, semiconsolidated, and consolidated deposits (Tyley 1974). The main water-bearing sediment units are Pleistocene and Holocene unconsolidated deposits. These are poorly sorted heterogeneous alluvial fan deposits with few fine sediments, forming unconfined aquifers (CVWD 2002). Furthermore, the foothills of the mountains are formed by partly consolidated deposits of Pliocene and Pleistocene age that yield little water.

8 8 Appana et al. Similarly, the pre-tertiary and Tertiary rocks of the basement and the San Jacinto Mountains do not yield any significant amount of water and thus act as groundwater flow barriers (CDWR 1964). The White Water sub-basin (WWB), southwest of the Garnet Hill Fault, hosts the groundwater recharge facility (Fig. 1). This basin has the most permeable deposits in the Upper Coachella Valley, allowing the most direct recharge of the unconfined aquifer below. 2.3 Poroelastic deformation Sneed & Brandt (2007) report aquifer deformation and related land-surface subsidence associated with groundwater overdraft in the Lower Coachella Valley ranging from about 35 to 150 mm in These studies further report an increase in subsidence of as much as 75 to 180 mm in Such deformation can be elastic and/or inelastic. Elastic deformation occurs when decreases in pore-fluid pressures due to groundwater discharge or withdrawal increase effective stresses, σ e, causing aquifer skeleton contraction (e.g., Poland 1981; Wang 2000; Hoffmann et al. 2001). However, if this decrease in pore-fluid pressure causes the effective stresses to increase beyond their pre-consolidation values, deformation is not recoverable, i.e., it is inelastic and results in a permanent loss of aquifer storage capacity (e.g., Poland 1981, 1984; Galloway et al. 1998). In addition, compaction of the clay-rich sediments present in the Lower Coachella Valley may contribute to inelastic deformation in that region (Galloway et al. 1999). Therefore, the recorded decrease in groundwater levels in the Lower Coachella Valley between 1996 and 2005 suggests that part of the observed land-surface subsidence might be permanent, resulting in loss of aquifer storage. To counteract land-surface subsidence, the Coachella Valley Water District (CVWD) has introduced artificial groundwater recharge in the upper valley since 1973 (CVWD 2002). Due to an increase in the amount of artificial groundwater recharge, satellite interferometric methods have registered ground-surface uplift in the upper valley during recent years (late-1993 to 2000). The InSAR data further shows differential land-surface uplift between the sub-basins separated by the BF and the GHF (Fig. 2). We suggest

9 Poroelastic Modeling in Coachella Valley, CA, USA 9 that such land-surface uplift may potentially be interpreted as the combined response of the aquifer-fault system to 1) artificial groundwater recharge and 2) tectonic stresses acting on the basin, with a strong preference toward the former as discussed later. 3 DATA We use both geophysical and hydrologic data to develop a coupled hydrologic and poroelastic model of the Upper Coachella Valley. The rectangular region stretching SW- NE in Figure 1 shows the area of interest which is primarily chosen for its proximity to wells where groundwater levels are available during late-1993 through 2000 (Sec. 3.3). In addition, this region is down-dip from the groundwater recharge area where groundwater levels are elevated due to the artificial groundwater recharge. Finally, the cross-sectional profile is drawn perpendicular to the main hydraulic head gradient, oriented down the axis of the valley. Table 1 lists the parameters used in this study. 3.1 Digital Elevation Model (DEM) data The topography across the actual basin is relatively flat with only about 60 m variation in elevation over a distance of 12 km (Fig. 1). We represent this topography in a two-dimensional (2D) vertical model by averaging land-surface elevations over a series of parallel SW-NE-trending cross-sections within the rectangular region ( km 2 ) shown in Figure 1 (i.e., averaging perpendicular to the cross-sections). Advantages of using such an average surface elevation are that 1) it includes topography information from the region immediately surrounding the model cross-section and 2) any local features deviating from the average topography are subdued, similar to the underlying groundwater-table topography. This averaged land-surface topography does not have a precise physical location on the map and is thus represented by the dashed line passing

10 10 Appana et al. through the mid-point of the rectangular region (Fig. 1), hereafter referred to as the model cross-section. 3.2 InSAR data Interferometric Synthetic Aperture Radar (InSAR) is a repeat-pass satellite data acquisition tool that can be effectively used for monitoring surface deformation due to groundwater extraction and recharge (e.g., Galloway et al. 1998; Amelung et al. 1999; Bawden et al. 2001; Lu & Danskin 2001; Schmidt & Bürgmann 2003; Bell et al. 2008; Wisely & Schmidt 2010). SAR data from the ERS1/2 satellite of the European Space Agency is used to measure surface deformation in the study region. This type of data is most useful in areas with little vegetative cover and in urban areas, and therefore is well suited for the semi-arid Coachella Valley. Differential interferograms can be constructed to span months, even years, depending on the suitability of SAR pairs regarding interferometry and coherence of each scene, and are processed to a spatial resolution of 30 m using the ROI PAC software package (Rosen et al. 2004). Flattening the interferometric phase removes 1) any gradient caused by orbital errors and 2) the horizontal displacement signal associated with plate boundary deformation. Because the perpendicular baseline between the orbital passes is greater than zero, there is a topographic contribution to the phase difference (Bürgmann et al. 2000). This is removed using a 30 m Digital Elevation Model (Fig. 1) from the Shuttle Radar Topography Mission (SRTM). The phase difference is then unwrapped to estimate range change in the satellite s line-of-sight (LOS) or look direction. Stacking differential interferograms estimates an average rate of deformation for each pixel and dampens the effect of atmospheric noise. We stack 23 interferograms (processed from 24 independent SAR scenes) by summing the LOS range change for a given pixel and dividing by the cumulative time spanned by all of the interferograms (Table 2). All 23 differential interferograms used in the analysis span intervals within late-1993 through 2000 and have a perpendicular baseline of < 277 m, averaging 77 m.

11 Poroelastic Modeling in Coachella Valley, CA, USA 11 These short distances in satellite positioning between acquisitions maximize coherence and minimize topographic errors. A phase step in an interferogram across a fault in this region (Fig. 2a) could be caused by subsurface fault slip for which we would expect predominantly horizontal surface displacement. Alternatively, natural or artificial groundwater recharge can cause differential uplift when subsurface hydraulic properties, for example, permeability or storage coefficients, are heterogeneously distributed causing non-uniform pore-fluid pressure and effective stress fields (e.g., Haneberg 1995; Bense & Balen 2004; Mayer et al. 2007). Therefore, a sharp contrast between fault and/or basin aquifer permeabilities can give rise to relatively sharp pore-fluid pressure changes and related uplift step functions across faults resulting in such a phase step in an interferogram. InSAR does not determine the full 3D displacement vector unless multiple satellite look directions can be employed in the analysis. However, we find that the ascending track data, although limited, is consistent in magnitude and direction to the more abundant descending data, thereby suggesting that the vertical direction is the dominant mode of deformation for this hydrogeologic basin (Bawden et al. 2001; Hoffman & Zebker 2003; Wisely & Schmidt 2010). This also suggests that most of the land-surface uplift during late-1993 to 2000 is caused by groundwater recharge. Hereafter, we refer to this poroelastic response of the aquifer simply as uplift. While tectonic causes of, or at least contributions to, the observed uplift pattern cannot be completely excluded, they are likely minor and thus beyond the scope of the present study. Furthermore, since the look vector is nearly vertical (Fig. 2a), we assume that the LOS range change observation is representative of the vertical deformation. The actual vertical deformation is 6% more than the LOS observations once the LOS is back-projected onto a vertical unit vector. However, the spatial averaging of the LOS observations perpendicular to the profile across the basin (discussed below) and the likelihood of only a small horizontal deformation component, means that the uncertainty in vertical deformation is greater than the 6% LOS correction. The patterns of time-dependent surface displacement may also be assessed by time-

12 12 Appana et al. series analysis (Fig. 3). In regions like Coachella Valley, where surface deformation may be due to more than one process, high temporal resolution of a time-series analysis is a method used to separate transient seasonal deformation from long-term deformation. Therefore, we produce a pixel-by-pixel time-series of vertical displacement using an inversion method developed by Schmidt & Bürgmann (2003). The method solves for the incremental range change between SAR scene acquisitions employing a linear inversion which translates the set of interferograms into a range-change time-series. The same interferograms used in stacking are used in the time-series construction (Table 2). Scenes used in multiple interferograms are down-weighted in the time-series inversion process, so that artifacts in repeated scenes do not dominate the time-series of deformation. The resultant time-series is a pixel-by-pixel deformation map sequence from late-1993 to 2000, showing vertical deformation accumulating from scene acquisition date to acquisition date. Analogously to averaging multiple topographic cross-sections (Sec. 3.1), we calculate averages along the model cross-section from the stacked interferogram showing uplift rates (Fig. 2b) and from the time-series data showing cumulative uplift (Fig. 3a). Such averaging provides a more representative uplift profile for the model cross-section and allows for better predictions of average permeabilities for the aquifer-fault system. In part due to this approximation, our estimates of permeability and other parameters along the cross-section are not exact but rather first-order approximations of plausible parameter ranges. Reporting such parameter ranges is, however, consistent with the notion that natural geologic parameters over large spatial scales are likely heterogeneous and/or anisotropic. From the time-series of average uplift (Fig. 3a), we calculate a sub-basin spatial average of the land-surface uplift (Fig. 3b) for all three sub-basins in the upper valley. The uncertainties in these averages are estimated using standard error propogation. Excluding negligible tectonic effects discussed previously (Sec. 3.2), the total uplift recorded in the upper valley sub-basins from late-1993 to 2000 is caused by a convolution of seasonal recharge/discharge and artificial recharge of groundwater at the White Water

13 Poroelastic Modeling in Coachella Valley, CA, USA 13 Recharge Facility. We assume that artificially recharged groundwater does not reach the Mission Creek sub-basin (MCB) due to its distal location to the White Water Recharge Facility as may be inferred from Figure 1. Thus, we attribute the temporal variations in spatially averaged MCB uplift (MCB-curve in Fig. 3b) solely to natural seasonal changes in groundwater content and related pore-fluid pressures. Furthermore, assuming that seasonal groundwater recharge/discharge effects in the White Water subbasin (WWB) and the Garnet Hill sub-basin (GHB) are similar to those in the MCB, we approximately remove the natural seasonal land-surface uplift/subsidence from the total uplift (Fig. 3a) by subtracting the spatially averaged MCB uplift from that of the WWB and the GHB, resulting in a relative land-surface uplift (Fig. 3c). This relative uplift of the WWB and the GHB should thus largely reflect cumulative uplift that is solely caused by the increase in artificial groundwater recharge in the WWB during late to 2000 (Fig. 4), compared to that during 1987 to 1993, as discussed in Section 3.3. Moreover, we observe that the relative uplift in the WWB and the GHB from late-1993 to 2000 (Fig. 3c) is significantly larger than 1) the uncertainties introduced by averaging uplift as discussed above and 2) the seasonal variations in land-surface uplift, indicating that the majority of the uplift is due to groundwater recharge. Therefore, the average uplift rate (Fig. 2b) calculated from the stacked interferogram along the model crosssection is adjusted such that the uplift rate from late-1993 to 2000 is zero throughout the MCB. 3.3 Hydrologic data Figure 4a shows temporal variations in water-table depth below ground-surface elevation (GSE) at nine monitoring wells (Fig. 1) projected onto the model cross-section. The peaks in the water levels of Wells W6, W7, W8, and W9 in the WWB appear to indicate responses to episodes of artificial groundwater recharge (Fig. 4b). In contrast, Wells W1, W2, W3, as well as W4 and W5 in the MCB and the GHB, respectively, are farther away from the artificial groundwater recharge area and thus show approximately constant groundwater levels during the same time period. Although these wells are at

14 14 Appana et al. different topographic elevations (Appendix C), the depth to the water in wells of a given sub-basin are approximately identical. Therefore, we assume that the groundwater-table elevation is a subdued replica of the topography within a sub-basin and is offset across different sub-basins with higher groundwater levels in the NE than in the SW of the cross-section (Fig. 5). The amount of artificial groundwater recharge, Q [m 3 /yr], introduced at the White Water Recharge Facility in the WWB from late-1993 to 1999 is shown in Figure 4b. The horizontal dashed and solid lines represent the average annual artificial groundwater recharge added from 1987 through 1993 ( m 3 /yr) and from late-1993 through 1999 ( m 3 /yr), respectively (CVWD 2002). The net change in artificial groundwater recharge, Q [m 3 /yr], is defined as the difference between the amount of average annual artificial groundwater recharge from late-1993 through 1999 and that from 1987 through However, only a portion of Q, hereafter denoted Q c [m 3 /yr/m], reaches the model cross-section where it causes poroelastic land-surface uplift (Fig. 2). Therefore, the actual value of Q c [m 2 /yr] is estimated as a plausible range. Hence, we consider two scenarios with minimum and maximum extent of alongvalley pore-fluid pressure diffusion of 3 and 9 km, respectively, represented by the two triangles (with bold and thin lines) in Figure 2a. The portion of the artificial groundwater recharge entering the rectangular region in the WWB is estimated to be proportional to the ratio of its area (trapezoid) to the areal extent of groundwater diffusion (the two triangles). This results in plausible Q c values of m 2 /yr Q c m 2 /yr. All simulations are thus conducted for Q c values of (min), (mean), and (max) m 2 /yr.

15 Poroelastic Modeling in Coachella Valley, CA, USA 15 4 CONCEPTUAL MODEL A cross-section across the Upper Coachella Valley, showing our conceptual two-dimensional model, is given in Figure 5. The cross-section extends from the foothills of the San Jacinto mountains in the SW to a distance of 11.5 km toward the NE into the Mission Creek sub-basin (MCB) where the effects of artificial groundwater recharge in the White Water sub-basin (WWB), and the runoff from the Little San Bernardino Mountains, are negligible. The permeability and its anisotropy, hereafter referred to as permeability structure, can be different for different sub-basins. Adjacent domains are linked by requiring continuity of hydraulic head and groundwater flux across the internal boundaries. The faults are modeled as vertical features (Yule & Sieh 2003), as a small deviation in actual fault orientation from vertical does not substantially modify the regional groundwater flow system (Folch & Mas-Pla 2008). Furthermore, we do not model the permeability structure of the MCB as there is no observed uplift across that sub-basin in the region of interest (Fig. 2). Instead, we use permeability estimates of m 2 k mcb x m 2 from previous studies (Tyley 1974; Swain 1978; Mayer et al. 2007) for the MCB. No aquitards are modeled as the clay content of the aquifer sediments is minor in the upper valley region (CVWD 2002). The San Jacinto Mountains toward the SW end of the model cross-section and the low-permeability basement rocks act as barriers to groundwater flow restricting the active groundwater flow zone to the upper 450 m of the basin, which corresponds to a water-saturated sediment thickness of about 300 m (Tyley 1974; Swain 1978). There is no cross-valley flow across the NE boundary of the cross-section, as it is far away from both the groundwater recharge area and the runoff from the mountains to the NE. In our model, increase in pore-fluid pressures can induce poroelastic land-surface uplift in two ways: 1) by increasing the water pressure in saturated pores and 2) by raising the water table into previously unsaturated zones which increases pore-fluid pressures from atmospheric, or sub-atmospheric, to (approximately) hydrostatic conditions. In order to constrain hydrogeologic parameters, we conduct simulations to explore parameter space. To reduce the number of unknowns to a tractable value, we make

16 16 Appana et al. the following assumptions: 1) All uplift is caused by poroelastic deformation of the sediment matrix, i.e., there is no tectonic uplift (Sec. 3.2). 2) The main along-valley groundwater flow, perpendicular to the cross-section, does not affect water-table elevations or pore-fluid pressures, as this flow into and out of the cross-section is equal. 3) Groundwater recharge rates to valley aquifers through the mountains are negligible in comparison to both along-valley and across-valley components of groundwater flow (CDWR 1964). 4) Groundwater flow in the aquifer is assumed to be isothermal due to the small topographic relief and groundwater flow depths encountered within the sedimentary basin and negligible groundwater flow within the adjacent mountains (see Point 3). 5) Artificial groundwater recharge at the surface enters the aquifer system instantaneously due to the large proportion of coarse, high-permeability sediments present in the WWB (CVWD 2002). Therefore, flow in the unsaturated zone is not explicitly simulated. 6) The water table is a subdued replica of the topography within fault-bounded sub-basins (Sec. 3.3). 7) Equilibrium conditions for poroelastic deformation are reached instantaneously whenever subsurface pore-fluid pressures change. This implies that a) diffusion of pore-fluid pressure in the sub-surface is a relatively fast process so that water levels at any instant determine groundwater flow velocity vectors and b) the coupling between pore-fluid pressures and stresses in the sediment matrix is instantaneous. 8) Fault-zones are simulated as isotropic porous media, with a uniform width of 50 m throughout the entire depth of the basin (Mayer et al. 2007), although permeability anisotropy and a different width (of 10 m) are considered later by employing a heterogeneous fault-zone architecture. 9) The permeability structures of all formations (three sub-basins and two faults in Fig. 5) are homogeneous. However, different formations may have different permeabilities, so that the aquifer-fault system as a whole can be heterogeneous. 10) The effect of net artificial groundwater recharge on uplift is negligible in the Mission Creek sub-basin (Sec. 3.2).

17 Poroelastic Modeling in Coachella Valley, CA, USA 17 5 NUMERICAL MODEL We develop a two-dimensional, vertical numerical model, employing the finite-element simulator COMSOL Multiphysics (R), to constrain the permeability structure of the subsurface. Our model simulates instantaneous artificial groundwater recharge causing pore-fluid pressure diffusion and groundwater flow and associated poroelastic stress redistribution leading to deformation of porous media in both horizontal and vertical directions along the model cross-section. The numerical model grid has triangular mesh elements with quadratic shape functions that interpolate parameter values in-between element corners. The triangular elements are of varied sizes with coarser element resolutions in the center of the model sub-domains and finer element resolutions toward subdomain edges. The total number of elements used in the model cross-section ( km 2 ) is The simulation procedure is discussed in the following sections. 5.1 Initial state (late-1993) The initial pore-fluid pressure distribution (Fig. 6a) and the state of stresses (Fig. 6e) in the aquifer are defined by conducting steady-state simulations of fluid flow and planestrain deformation for a baseline set of parameter values. For the 2D fluid-flow model, the water table is approximated from observed groundwater levels during late-1993 (Sec. 3.3). The left, right, and bottom edges of the model cross-section are no-flow boundaries (Table 3). For the 2D plane-strain model, the boundary conditions defining the horizontal and vertical movement of the porous sediment matrix are described in Table 3. The flow lines and the z-displacements of the porous sediment matrix modeled for late-1993 are shown in Figures 6b and 6f, respectively. 5.2 Final State (2000) The final state of the aquifer is modeled by introducing artificial groundwater recharge, i.e., by adding a fluid source term. An integrated constant-flux term, Q c [m 2 /yr],

18 18 Appana et al. representing the net change in artificial groundwater recharge entering the cross-section (Sec. 3.3), is introduced in the White Water sub-basin along a line source. The line source is centered at R(x, z) = (2600, 120) m, approximately at the location of the peak uplift rate along the model cross-section (Fig. 2), and has a horizontal extent of L = 2500 m. Both R and L are constant parameters in the steady-state simulation. The fluid-flow model estimates the resultant pore-fluid pressure redistribution in the subsurface (Fig. 6c). These pore-fluid pressure changes lead to modifications in the Darcy velocity field and associated groundwater flow lines (Fig. 6d). The plane-strain model in turn calculates the resultant stresses (Fig. 6g) and displacements (Fig. 6h) for the new hydrogeologic conditions. Although we show only the vertical displacements in our illustrations, we consider the total two-dimensional displacement vector when calculating the misfit, M [mm/yr], discussed below. The rate of uplift is shown as a function of position, x, along the model crosssection (e.g., Fig. 7b). We explore parameter space of basin and fault permeabilities (as discussed in Sec. 6), minimizing the misfit between modeled and measured uplift rates. The misfit, M [mm/yr], is defined as a root-mean-squared error along the model cross-section, where M = 1 n n ( u(x i )) 2, (11) i=1 u(x i ) u obs (x i ) if u(x i ) u obs (x i ) ū obs (x i ) u(x i ) = 0 if u(x i ) u obs (x i ) < ū obs (x i ). In Equations (11) and (12), x i [km] are distances along the cross-section, u(x i ) [mm/yr] and u obs (x i ) [mm/yr] are the simulated and observed uplift rates at x i, respectively, ū obs (x i ) [mm/yr] are the standard deviations of u obs (x i ) at x i, and n is the number of measurements along the model cross-section. (12)

19 Poroelastic Modeling in Coachella Valley, CA, USA 19 6 RESULTS We employ numerical models (Secs 4 and 5), simulating pore-fluid pressure diffusion, groundwater flow, and land-surface uplift to explore parameter space (Table 4) of k wwb x, a wwb, k ghf (k ghf x = k ghf z ), k ghb, a ghb, and k bf (k bf x x = k bf z ). Other parameters, D, L, R(x, z), b, Y, α, η, ν, and ρ s, described in Table 1, are held constant and assigned typical values for the Upper Coachella Valley, as variations in these parameters are small compared to the variable parameters (numbered in square brackets throughout this paper) listed in Table 4. Appendix A provides detail regarding the interdependence of some model parameters. In the following, sub-basins and faults are discussed in order of appearance along the modeled cross-section from the SW (near the White Water Recharge Facility) to the NE (Fig. 5). As mentioned before, uplift in the MCB beyond that caused by natural grounwater recharge is not observed (Fig. 3c), so that the BF is the last feature discussed along the cross-section. 6.1 White Water sub-basin (WWB) We use a contour plot of misfit (Fig. 7a) to analyze the simulations conducted by varying parameters k wwb x [1] versus a wwb [2] at constant Q c [3] = m 2 /yr, k ghf [4], kx ghb [5], a ghb [6], and k bf [7] as shown in Table 4. The dots in the figure represent parameter combinations of kx wwb [1] and a wwb [2]=(k z /k x ) wwb at which simulations are run. Contours with lower misfit numbers, M, represent regions with better matches between observed and simulated uplift rates. For these simulations, the bold solid line in Figure 7a represents the contour with minimized misfit values of M 0.14 mm/yr. The dashed and dot-dashed curves in Figure 7a are contours with minimized misfit values of M 0.14 mm/yr for simulations conducted at Q c [3] values of and m 2 /yr, respectively, and constant [4]-[7]. However, the best match is obtained only after optimizing all parameters considered, as discussed further in the following sections. As expected, an increase in kx wwb [1], at constant values for parameters [2]-[7], results

20 20 Appana et al. in a decrease in simulated uplift rates in the WWB (Fig. 7b). Similarly, an increase in a wwb = (k z /k x ) wwb [2], at constant [1] and [3]-[7], results in a decrease in simulated uplift rates in the WWB (Fig. 7c). It is important to note that variations in k wwb x [1] and a wwb [2] also affect simulated uplift rates in the adjacent Garnet Hill sub-basin (GHB) but not the Mission Creek sub-basin (MCB). Furthermore, these simulations show that estimates of kx wwb [1] at constant [2] and [4]-[7], depend on parameter values for Q c [3]. Therefore, by considering a plausible range of values for Q c [3] (Sec. 3.3), we constrain the permeability structure of the WWB near the model cross-section (Fig. 1) to m 2 k wwb x m 2 and 0.04 a wwb In summary, the WWB aquifer appears to be anisotropic, with higher horizontal permeability values, kx wwb [1], associated with larger permeability anisotropies, where the latter are indicated by decreasing values of a wwb [2] further away from Garnet Hill Fault (GHF) To estimate the permeability of the Garnet Hill Fault (GHF), we explore the parameter space of k ghf [4] versus a ghb [6], at constant k wwb x [1], a wwb [2], Q c [3], kx ghb [5], and k bf [7] (Table 4). Simulation results suggest that increasing values of a ghb [6]=(k z /k x ) ghb at constant [1]-[3], [5], and [7] requires a corresponding increase in k ghf [4] to maintain minimized misfit values (Fig. 8a). However, this trend is finite so that a bounded elliptical region with minimized misfit, M, exists. The effect of variations in kx ghb [5], the parameter representing the horizontal permeability of the GHB on the permeability, k ghf [4], of the GHF is illustrated by comparing Figures 8a and 8b. Figures 8a and 8b show contour plots of misfit, M, corresponding to simulations conducted with values of kx ghb [5] equal to (minimum) and m 2 (maximum), respectively (Reichard & Meadows 1992), at constant [1]-[3] and [7]. Furthermore, the effect of variations in parameter values of Q c [3] is also explored (Appendix A). In summary, our simulations suggest a permeability range for the GHF of m 2 k ghf m 2. As expected, an increase in Garnet Hill Fault permeability, k ghf [4], at constant values for all other parameters considered, re-

21 Poroelastic Modeling in Coachella Valley, CA, USA 21 sults in an increase in simulated uplift rates in the adjacent GHB (Fig. 8c), as pore-fluid pressure diffusion from the WWB to the GHB is facilitated. However, it is important to note that these variations in k ghf [4] do not appear to influence the simulated uplift in the WWB, where the artificial groundwater recharge occurs. 6.3 Garnet Hill sub-basin (GHB) We use estimates of the horizontal permeability of the GHB from previous studies (Tyley 1974; Swain 1978) as input values for kx ghb [5] in our simulations (for justification, see Appendix A) but do constrain the permeability anisotropy of the GHB to a ghb [6] (Figs 8a-b). As expected, an increase in k ghb x [5], at constant values for parameters [1]-[4], [6], and [7] (Table 4), results in a decrease in simulated uplift rates in the GHB (Fig. 9a). Similarly, an increase in a ghb [6] = (k z /k x ) ghb, at constant [1]-[5] and [7] (Table 4), results in a decrease in uplift rates in the GHB as well (Fig. 9b). However, we note that the variations of kx ghb [5] and a ghb [6] do not affect uplift rates in the adjacent sub-basins, i.e., the WWB or the MCB, as shown in Figures 9a-b. 6.4 Banning Fault (BF) In order to conclusively constrain the permeability of the BF it would be necessary to conduct a 5-dimensional parameter analysis, i.e., a wwb [2] versus Q c [3] versus k ghf [4] versus a ghb [6] versus k bf [7], at constant k wwb x [1] and kx ghb [5], which is beyond the scope of this study (Secs 6.1, 6.2, and 6.3). However, we conduct simulations of land-surface uplift considering six plausible combinations of parameter values for [1]-[6] (Table 4) within their likely variability as determined elsewhere in this study. Figure 10 shows the reduced misfits, δm = M M i min, as a function of the permeability of the BF, where M i min, with i = 1 to 6, are the minimum misfit values corresponding to the simulations run with the six different combinations of parameters [1]-[6], respectively. We observe that irrespective of the parameter values used for [1] to [6], the reduced misfit increases with increasing values for k bf [7] above m 2 (Fig. 10). There-

22 22 Appana et al. fore, we consider that this value represents a maximum permeability for the BF. That only an upper bound can be established for k bf [7] is not surprising, as the fault permeability needs to be just low enough to prevent land-surface uplift in the adjacent MCB (as observed - see Fig. 3c) by sufficiently limiting pore-fluid pressure diffusion and groundwater flow across the BF. 7 DISCUSSION Estimates of permeability are often scale-dependent (e.g., Zlotnick et al. 2000; Ingebritsen et al. 2006). Our study provides additional insights into the subsurface hydrogeology of the Upper Coachella Valley when combined with previous permeability estimates determined on local scales using well logs (Tyley 1974) and regional scales (few km) using steady-state groundwater flow models (Tyley 1974; Swain 1978). Figure 11a shows a summary of the permeability estimates in the Upper Coachella Valley from this study in comparison to previous work. Figure 11b shows the uplift rate along the model cross-section for one plausible combination of parameter values (listed under Fig. 11b in Table 4) giving a close-to optimal fit between simulated and observed uplift rates. However, the model seems to slightly underpredict the observed uplift rates on the left side of the profile (in the WWB) and a better fit may be obtained by changing the position, R(x, z), and/or length, L, of the recharge zone (Sec. 5.2). Since the simulated uplift rates shown are within the zone of acceptable values indicated by the standard deviation of observed uplift rates and given how the misfit, M, is determined (Equations 11 and 12), we do not expect significant changes in the estimated permeabilities resulting from variations in R and L. Therefore we keep these secondary parameters fixed.

23 7.1 Permeability structure of the sub-basins Poroelastic Modeling in Coachella Valley, CA, USA 23 The estimates for the horizontal permeability of the WWB (Fig. 11a) in the Upper Coachella Valley are well within the range of values representative of clean unconsolidated sand deposits (Freeze & Cherry 1979) present in the study region (Proctor 1968). Furthermore, our large-scale (several km) permeability estimates for the WWB of m 2 k wwb x m 2 match closely values of m 2 k wwb x m 2 from previous studies (Tyley 1974; Swain 1978). Our permeability anisotropy estimates of 0.04 a wwb = (k z /k x ) wwb for the WWB (Sec. 6.1) predict higher horizontal permeabilities and may indicate stratification due to alternation of coarse- and fine-grained sediment deposits, resulting from variations in stream power from wet to dry years (CDWR 1964). A possible explanation is that horizontal layers in parallel, formed by such stream processes, provide overall horizontal permeabilities given by the arithmetic mean, which is dominated by the highest-permeability layer (e.g., Saar & Manga 2004; Ingebritsen et al. 2006). Our analysis also suggests that the permeability anisotropy of the GHB of a ghb is approximately an order of magnitude smaller than that of the WWB (Fig. 11a). Furthermore, the horizontal permeabilities of m 2 k ghb x m 2 considered for the GHB in model simulations from a previous study (Reichard & Meadows 1992) are an order of magnitude less than those predicted here for the adjacent WWB (Sec. 6.1). This is expected, as basic sedimentological principles suggest that sediments deposited farther away from the mountains, for example, in the GHB, tend to be relatively finer-grained, and thus less permeable, than those deposited closer to the sediment source, for example, in the WWB (CDWR 1964). We thus suggest that this contrast in the average permeability structure of the WWB and the GHB can be caused by differing sediment types.

24 24 Appana et al. 7.2 Fault permeabilities Our study suggests that the inferred permeability of the GHF of m 2 k ghf m 2 is an important hydrogeologic parameter to consider when managing groundwater resources and related processes in the region, for example, to reverse landsurface subsidence in the GHB (e.g., Fig. 8c). Furthermore, our estimates for k ghf result in permeability contrasts of 10 2 kx wwb /k ghf 10 4 between the GHF and the WWB and 10 1 kx ghb /k ghf 10 3 between the GHF and the GHB. Such permeability contrasts suggest that the GHF acts as a semi-permeable barrier to groundwater flow. In addition, we establish a maximum permeability value for the BF of k bf m 2 (Fig. 11a). This maximum likely permeability is of particular interest when determining the permeability contrast between the sub-basins and the faults as the basins have much higher permeabilities. We calculate a permeability contrast of kx ghb /k bf 10 3 between the GHB and the BF and kx mcb /k bf 10 2 between the MCB and the BF. However, these inferred permeability contrasts are only a conservative minimum estimate. Analogously to the GHF discussed above, these permeability contrasts suggest that the BF also acts as a semi-permeable barrier to groundwater flow. Figure 11a summarizes all results. The permeability structure of a fault zone is often complex, affecting groundwater movement or trapping within the fault zone. For example, field observations suggest that fault zones are generally composed of two main components (e.g., Chester & Logan 1986; Caine et al. 1996): 1) a fault core where comminution produces fine-grained fault gouge or cataclasite, and 2) an adjacent damage zone with enhanced fractures and veins. The permeability structure of faults is strongly determined by the distribution of cross-cutting fractures and the grain size within these fault components. Furthermore, the proportion of fault core to damage zone depends on the type of basin sediments and the geologic setting of the fault (Caine et al. 1996). Faults can act as barriers when 1) they contain finer-grained, and thus low-permeability, fault gouge (Caine et al. 1996), 2) their permeabilities are reduced as a result of mineral precipitation from circulating fluids during fault diagenesis (Ingebritsen et al. 2006), 3) they juxtapose hydrogeologic

25 Poroelastic Modeling in Coachella Valley, CA, USA 25 units of contrasting permeability structures (Maclay & Small 1983; Galloway & Hoffmann 2007), and 4) realigned elongated clasts reduce fault-normal permeability and increase fault-parallel permeability (Davis & DeWeist 1966). Although some permeability contrast exists between the WWB and the GHB as well as between the GHB and the MCB (Fig. 11a), we do not expect this to result from juxtaposition of hydrogeologic units across the GHF and the BF, respectively, because both faults are relatively young (Langenheim et al. 2005) and such a scenario would require significant displacement history. Additionally, the here-inferred bulk fault-zone permeability is two to three orders of magnitude smaller than those of the nearby sub-basins, which argues against Point (3) above, because in such a case, fault-zone permeability would be expected to be on the same order of magnitude as the two sub-basin permeabilities. In contrast, the displacement along both of these faults, although small, is likely sufficient to cause wear and gouge formation, arguing for Points (1) and (4) above. We use a simplified fault-zone model for both the GHF and the BF as shown in Figure 12, to constrain the horizontal permeabilities of the corresponding fault cores (Table 5). We vary the damage-zone permeabilities in our model to consider the possibility of faults acting either as barriers to fluid flow (e.g., Sand Hill Fault, Albuquerque basin, USA; Rawling et al. (2001)) or combined conduit-barrier systems (e.g., Lower Rhine Embayment, Germany; Bense et al. (2008); Baton Rouge Fault, Louisiana, USA; Bense & Person (2006)) in sedimentary basins. In the latter case, the fault zone as a whole is modeled with a large permeability anisotropy (vertical permeability horizontal permeability) such that the majority of fluid flow occurs within the damage zone and along the fault plane while a considerable hydraulic head drop across the fault core is maintained. Therefore, (vertical) layers of the fault-adjacent damage zones and the fault core are located in-series with respect to horizontal groundwater flow. Hence, the bulk horizontal permeability of the fault zone is given by the harmonic mean of the permeabilities of the individual layers as b k b = d k dl + c k c + d k dr, (13)

26 26 Appana et al. where, b [m] is the width of the fault zone considered, k b [m 2 ] is the bulk permeability of the fault zone, c [m] is the width of the fault core, k c [m 2 ] is the (unknown) permeability of the fault core, d = (b c)/2 [m] is the width of the damage zones and k d = k dl = k dr [m 2 ] is the permeability of the damage zone on either side of the fault core. We estimate the fault-core permeabilities using a range of fault-core widths, c, a range of values for the bulk permeabilities of fault zones, k b, established from our numerical simulations, and two different fault-zone widths, b (Table 5). Since our simulations are conducted for a constant fault-zone width, the k b values used in Equation (13) are only valid for b = 50 m. For any other value of b, the bulk permeability of the fault zone, k b, is recalculated as shown in Appendix B. These considerations constrain the fault-core permeabilities of the GHF and the BF to m 2 k ghf c k bf c m 2, respectively m 2 and Direct observations of fault permeabilities at major plate-boundary faults are not readily available in the literature. Chu et al. (1981) found that fault-gouge samples from the SAF exhibit permeabilities between and m 2. Similarly, Mizoguchi et al. (2008) report fault-core permeabilities between and m 2 at the Nojima fault in Japan. Both of these permeability ranges are very low with respect to fluid flow as they are comparable to unfractured metamorphic and igneous rocks and shale (e.g., Freeze & Cherry 1979) and to geologic materials more typically found at depths of approximately 10 km or more (e.g., Manning & Ingebritsen 1999; Saar & Manga 2004). In contrast, our estimates for fault-core permeability are one to two orders of magnitude higher. It is important to note, however, that our fault permeability estimates apply only to the shallow depths of the investigated basins (Secs 6.2 and 6.4). Thus, our fault-core and fault-zone permeabilities may serve as upper limits for permeabilities likely present at greater depths, as permeability is expected to decrease with depth (e.g., Manning & Ingebritsen 1999; Faulkner 2004; Saar & Manga 2004).Furthermore, Mizoguchi et al. (2008) found that permeability values estimated from surface samples compare well

27 Poroelastic Modeling in Coachella Valley, CA, USA 27 with borehole samples down to 2 km depth. Thus, our constraints for fault-core and fault-zone permeabilities may also be applicable for the uppermost 2 km of the SAF in the Coachella Valley region. Quantitative estimates of fault-zone permeabilities provide important constraints for understanding the static and dynamic weakening of faults. Lower fault-core permeabilities can cause elevated pore-fluid pressures within the fault core and lead to reduction of normal stresses across the fault (e.g., Byerlee 1990; Rice 1992). Consequently, lower shear-driving stresses are sufficient to cause fault slip at seismogenic depths (Raleigh et al. 1972, 1976; Talwani & Acree 1984; Simpson et al. 1988; Gupta 1992; Saar & Manga 2003). By considering the SAF as a hydrologic barrier, Fulton & Saffer (2009) show that such elevated pore-fluid pressures can be localized within the faults at depths greater than 5 km when horizontal permeabilities of k x m 2, at a depth of 500 m, are considered for the fault zone. Dynamic rupture is also affected by low permeabilities within the fault core relative to the damage zone or surrounding country rock. For example, several studies have explored thermal pressurization as a dynamic weakening mechanism (e.g., Sibson 1973; Vredevoogd et al. 2007). Finally, through numerical simulations, Bizzarri & Cocco (2006) show that source parameters such as stress drop, slip-weakening distance, and fracture energy all decrease systematically as permeability increases within the fault. Thus, estimates of fault-core and fault-zone permeabilities, as presented here, provide important constraints for conceptual and numerical models of active faulting. 8 CONCLUSIONS Our study suggests that the subsurface permeability structure (i.e., permeability and permeability anisotropy) in the Upper Coachella Valley region is heterogeneous over

28 28 Appana et al. basin scales, with the White Water (WWB), the Garnet Hill (GHB), and the Mission Creek (MCB) sub-basins having permeabilities typical for groundwater aquifers. Furthermore, we infer that the permeability anisotropy of the GHB and the WWB are significantly different, with at least an order of magnitude variability (Fig. 11a), likely caused by different sediment types existing in these sub-basins. In addition, simulations suggest that the permeability of the Garnet Hill Fault (GHF) is an important hydrogeologic parameter to consider when managing groundwater in the region, for example, to reverse land-surface subsidence in the GHB. InSAR observations in conjunction with numerical modeling suggest that the GHF and the Banning Fault (BF) act as semi-permeable barriers to groundwater flow in this region. We calculate a permeability contrast of between the GHF and the surrounding sub-basins and at least 10 2 between the BF and the surrounding sub-basins, with the faults having lower permeabilities in either case. Moreover, our calculations suggest fault-core permeabilities of k ghf c m 2 and k bf c m 2 for the GHF and the BF, respectively, which are one to two orders of magnitude higher than previous estimates for the San Andreas Fault (SAF). However, our estimates of fault-core permeabilities are only valid down to a maximum depth of 2 km, but may thus serve as an upper limit for fault-core permeabilities at greater depths. Therefore, our study provides estimates for the SAF-zone permeability structure and architecture which can potentially be used to constrain conceptual and numerical models of active faulting in general and SAF slip dynamics in particular. 9 ACKNOWLEDGEMENTS We thank Stuart Walsh for help with the numerical analysis, John Boggs for computational support and the reviewers for their helpful comments and suggestions. We also thank the George and Orpha Gibson endowment for its generous support of the Hydro-

29 Poroelastic Modeling in Coachella Valley, CA, USA 29 geology and Geofluids group. This research was supported by NSF grants EAR and EAR , as well as a grant by the Institute for Renewable Energy and the Environment (IREE) at the University of Minnesota. Any opinions, findings, and conclusions or recommendations expressed in this material are those of the authors and do not necessarily reflect the views of the NSF or the IREE. REFERENCES Amelung, F., Galloway, D. L., Bell, J. W., Zebker, H. A., & Laczniak, R., Sensing the ups and downs of Las Vegas: InSAR reveals structural control of land subsidence and aquifer-system deformation, Geology, 27(6), Antonellini, M. & Aydin, A., Effect of faulting on fluid flow in porous sandstone: Petrophysical properties, Am. Assoc. Petrol. Geol. Bull., 78, Bawden, G. W., Thatcher, W., Stein, R. S., Hudnut, K. W., & Peltzer, G., Tectonic contraction across Los Angeles after removal of groundwater pumping effects, Nature, 412, Bell, J. W., Amelung, F., Ferretti, A., Bianchi, M., & Novali, F., Permanent scatterer In- SAR reveals seasonal and long-term aquifer-system response to groundwater pumping and artificial recharge, Water Resour. Res., 44. Bense, V. F. & Balen, R. V., The effect of fault relay and clay smearing on groundwater flow patterns in the Lower Rhine Embayment, Basin Res., 16, Bense, V. F. & Person, M. A., Faults as conduit-barrier systems in siliciclastic sedimentary aquifers, Water Resour. Res., 42(W05421), 18 pp. Bense, V. F., Person, M. A., Chaudhary, K., You, Y., Cremer, N., & Simon, S., Thermal anomalies indicate preferential flow along faults in unconsolidated sedimentary aquifers, Geophys. Res. Lett., 35(L24406), 6 pp. Bizzarri, A. & Cocco, M., A thermal pressurization model for the spontaneous dynamic rupture propagation on a three-dimensional fault: 2. Traction evolution and dynamic parameters, J. Geophys. Res., 111(B05304). Bredehoeft, J. D., Belitz, K., & Sharp-Hansen, S., The hydrodynamics of the Big Horn basin: the study of the role of faults, Am. Assoc. Petrol. Geol. Bull., 76, Bürgmann, R., Rosen, P. A., & Fielding, E. J., Synthetic aperture radar interferometry to measure Earth s surface topography and its deformation, Annu. Rev. Earth Planet. Sci., 28, Byerlee, J., Friction, overpressure, and fault normal compression, Geophys. Res. Lett., 17(12), Byerlee, J., The change in orientation of subsidiary shears near faults containing pore fluid under high pressure, Tectonophysics, 211,

30 30 Appana et al. Caine, J. S., Evans, J. P., & Forster, C. B., Fault zone architecture and permeability structure, Geology, 24(11), CDWR, Coachella Valley Investigation, California Dept. of Water Res. (CDWR), Bulletin No. 108, 145 pp. Chester, F. M. & Logan, J. M., Composite planar fabric of gouge from the Punchbowl fault, California, J. Struc. Geo., 9, Chu, C. L., Wang, C. Y., & Lin, W., Permeability and frictional properties of San Andreas Fault Gouges, Geophys. Res. Lett., 8(6), CVWD, Coachella Valley Final Water Management Plan, Coachella Valley Water District (CVWD), 277 pp. Darcy, H., Les Fontaines Publiques de la Ville de Dijon ( The Public Fountains of the Town of Dijon ), Dalmont, Paris, 647 pp. Davis, S. N. & DeWeist, R. J. M., Hydrogeology, John Wiley, New York. Eichhubl, P., D Onfro, P. S., Aydin, A., Waters, J., & McCarty, D. K., Structure, petrophysics, and diagenesis of shale entrained along a normal fault at Black Diamond Mines, California - Implications for fault seal, Am. Assoc. Petrol. Geol. Bull., 89, Faulkner, D. R., A model for the variation in permeability of clay-bearing fault gouge with depth in the brittle crust, Geophys. Res. Lett., 31(L19611). Folch, A. & Mas-Pla, J., Hydrogeological interactions between fault zones and alluvial aquifer in regional flow systems, Hydrol. Process., 22, Freeze, R. & Cherry, J. A., Groundwater, Princeton-Hall, Inc., Englewood Cliffs, N.J., 604 pp. Fulton, P. M. & Saffer, D. M., Potential role of mantle-derived fluids in weakening the San Andreas Fault, J. Geophy. Res., 114(B07408), 15 pp. Galloway, D. L. & Hoffmann, J., The application of satellite differential SAR interferometryderived ground displacements in hydrogeology, Hydrogeol. J., 15(1), Galloway, D. L., Hudnut, K. W., Ingebritsen, S. E., Phillips, S. P., Peltzer, G., Rogez, F., & Rosen, P. A., Detection of aquifer system compaction and land subsidence using Interferometric Synthetic Aperture Radar, Antelope Valley, Mojave Desert, California, Water Resour. Res., 34(10), Galloway, D. L., Jones, D. R., & Ingebritsen, S. E., Land subsidence in the United States, United States Geological Survey, Circular 1182, 177 pp. Gupta, H. K., Reservoir-Induced Earthquakes, Elsevier Scientific Publishing Co., Amsterdam, p. 365 pp. Haneberg, W. C., Steady-state groundwater flow across idealized faults, Water Resour. Res., 31, Hickman, S., Stress in the lithosphere and the strength of active faults, Rev. Geophys., 29,

31 Poroelastic Modeling in Coachella Valley, CA, USA 31 Hoffman, J. & Zebker, H., Prospecting for horizontal surface displacements in Antelope Valley, California, using satellite radar interferometry, J. Geophys. Res., 108(F1), Hoffmann, J., Galloway, D. L., Zebker, H. A., & Amelung, F., Seasonal subsidence and rebound in Las Vegas Valley, Nevada, observed by Synthetic Aperture Radar interferometry, Water Resour. Res., 37, Ingebritsen, S. E. & Manning, C. E., Permeability of the continental crust: dynamic variations inferred from seismicity and metamorphism, Geofluids, 10, Ingebritsen, S. E., Sanford, W. E., & Neuzil, C., Groundwater in Geologic Processes, Second Edition, Cambridge University Press, Cambridge; New York, 341 pp. Lachenbruch, A. H. & Sass, J. H., Heat flow and energetics of the San Andreas Fault Zone, J. Geophy. Res., 85, Langenheim, V. E., Jachens, R. C., Matti, J. C., Hauksson, E., Morton, D. M., & Christensen, A., Geophysical evidence for wedging in the San Gorgonio Pass structural knot, southern San Andreas fault zone, southern California, GSA Bulletin, 117(11/12), Lockner, D. A., Tanaka, H., & Ito, H., Geometry of the Nojima Fault at Nojima-Hirabayashi, Japan - i. A Simple Damage Structure Inferred from Borehole Core Permeability, Pure Appl. Geophys., 166, Lu, Z. & Danskin, W. R., InSAR analysis of natural recharge to define structure of a groundwater basin, San Bernardino, California, Geophys. Res. Lett., 28(13), Maclay, R. W. & Small, T. A., Hydrostratigraphic subdivisions and fault barriers of the Edwards aquifer, south-central Texas, U.S.A., J. Hydrol., 61, Manning, C. E. & Ingebritsen, S. E., Permeability of the continental crust: Implications of geothermal data and metamorphic systems, Rev. Geophys., 37(1), Marler, J. & Ge, S., The permeability of the Elkhorn Fault Zone, South Park, Colorado, Ground Water, 41(3), Mayer, A., May, W., Lukkarila, C., & Diehl, J., Estimation of fault-zone conductance by calibration of a regional groundwater flow model: Desert Hot Springs, California, Hydrogeol. J., 15(6), Mizoguchi, K., Hirose, T., Shimamoto, T., & Fukuyama, E., Internal structure and permeability of the Nojima fault, southwest Japan, J. Struc. Geo., 30(4), Morrow, C. A. & Lockner, D. A., Permeability differences between surface-derived and deep drillhole core samples, Geophys. Res. Lett., 21(19), Poland, J. F., Subsidence in the United States due to groundwater withdrawal, Journal of Irrigation and Drainage Division, 107(IR2), Poland, J. F., Mechanics of land subsidence due to fluid withdrawal, Studies and Reports in Hydrology, 40, Proctor, J. P., Geology of the Desert Hot Springs upper Coachella Valley area, California,

32 32 Appana et al. California Division of Mines and Geology, Special Report, 94, 43 pp. Raleigh, C. B., Healy, J. H., & Bredehoeft, J. D., Faulting and crustal stress at Rangley, Colorado, in: Flow and Fracture of Rocks, Geophy. Monogr. Ser., pp Raleigh, C. B., Healy, J. H., & Bredehoeft, J. D., An experiment in earthquake control at Rangley, Colorado, Science, 191, Rawling, G. C., Goodwin, L. B., & Wilson, J. L., Internal architecture, permeability structure, and hydrologic significance of contrasting fault-zone types, Geology, 29, Reichard, E. G. & Meadows, J. K., Evaluation of a ground-water flow and transport model in the upper Coachella Valley, California, United States Geological Survey, Water Resour. Invest. Rep., , 101 pp. Rice, J. R., Fault Mechanics and Transport Properties of Rocks, edited by B. Evans and T.-F. Wong, chap. Fault stress states, pore pressure distributions, and the weakness of the San Andreas fault, pp , Elsevier. Rosen, P. A., Hensley, S., Peltzer, G., & Simmons, M., Updated repeat orbit interferometry package released, EOS Trans. AGU, 85(47), 108. Saar, M. O. & Manga, M., Seismicity induced by seasonal groundwater recharge at Mt. Hood, Oregon, Earth Planet. Sci. Lett., 214, Saar, M. O. & Manga, M., Depth dependence of permeability in the Oregon Cascades inferred from hydrogeologic, thermal, seismic, and magmatic modeling constraints, J. Geophys. Res., 109(B04204), 19 pp. Schmidt, D. A. & Bürgmann, R., Time-dependent land uplift and subsidence in the Santa Clara Valley, California, from a large Interferometric Synthetic Aperture Radar data set, J. Geophys. Res., 108(B9), Sibson, R. H., Interaction between temperature and pore-fluid pressure during earthquake faulting - A mechanism for partial or total stress relief, Nature, 243, Simpson, D. W., Leith, W. S., & Scholz, C. H., Two types of reservoir-induced seismicity, Bull. Seismol. Soc. Am., 78, Sneed, M. & Brandt, J. T., Detection and measurement of land subsidence using Global Positioning System surveying and Interferometric Synthetic Aperture Radar, Coachella Valley, California, , United States Geological Survey, Water Resour. Invest. Rep., , 31 pp. Swain, L. A., Predicted water-level and water-quality effects of artificial recharge in the upper Coachella Valley, California, using a finite-element digital model, United States Geological Survey, Water Resour. Invest. Rep., 77-29, 54 pp. Talwani, P. & Acree, S., Pore pressure diffusion and the mechanism of reservoir-induced seismicity, PAGEOPH, 122, Tsutsumi, A., Nishino, S., Mizoguchi, K., Uehara, S., Sato, K., Tanikawa, W., & Shimamoto, T., Principal fault zone width and permeability of the active Neodani fault, Nobi fault system,

33 Poroelastic Modeling in Coachella Valley, CA, USA 33 Southwest Japan, Tectonophysics, 379, Tyley, S., Analog model study of the ground-water basin of the upper Coachella Valley, California, United States Geological Survey, Water-Supply Paper, 2027, 77 pp. Vredevoogd, M. C., Oglesby, D. D., & Park, S. K., Fluid pressurization due to frictional heating on a fault at a permeability contrast, Geophys. Res. Lett., 34, 4 pp. Wang, H. F., Theory of Linear Poroelasticity with Applications to Geomechanics and Hydrogeology, Princeton University Press, Princeton, N.J. Wibberley, C. A. J., Yielding, G., & Toro, G. D., Recent advances in the understanding of fault zone internal structure: a review, Geol. Soc. Spec. Pub., London, 299, Wisely, B. A. & Schmidt, D. A., Deciphering vertical deformation and poroelastic parameters in a tectonically active fault-bound aquifer using InSAR and well level data, San Bernardino basin, California, Geophys. J. Int.. Yule, D. & Sieh, K., Complexities of the San Andreas fault near San Gorgonio Pass: Implications for large earthquakes, J. Geophys. Res., 108(B11), Zhang, S., Tullis, T. E., & Scruggs, V. J., Permeability anisotropy and pressure dependency of permeability in experimentally sheared gouge materials, J. Struc. Geo., 21, Zlotnick, V. A., Zurbuchen, B. R., Ptak, T., & Teutsch, G., Support volume and scale effect in hydraulic conductivity: experimental aspects. in Theory, Modeling, and Field Investigation in Hydrogeology: A Special Volume in Honor of Shlomo P. Neuman s 60th Birthday, Zhang, D., and Winter, C. L., Geol. Soc. Am. Special Paper, 348, Zoback, M. D., Zoback, M. L., Eaton, J. P., Mount, V. S., & Suppe, J., New evidence on the state of stress of the San Andreas Fault, Science, 238,

34 34 Appana et al. FIGURE LEGENDS Fig. 1: The shaded relief map of Upper Coachella Valley in southern California, USA, produced using a 30 m Digital Elevation Model (DEM) from the Shuttle Radar Topography Mission (SRTM). The Banning and Garnet Hill faults represent active strands of the San Andreas Fault. Squares represent monitoring wells close to the rectangular region of interest. The model cross-section is chosen along the dashed line. Artificial groundwater recharge occurs at the White Water Recharge Facility (diamond). Fig. 2: a) Stacked differential interferogram using Synthetic Aperture Radar (SAR) data from late-1993 to 2000 indicating line-of-sight (LOS) range change in the Upper Coachella Valley. Negative values (blue) indicates motion towards the satellite and positive values (red) indicates motion away from the satellite. White regions are decorrelated, i.e., uplift data is not available. Black lines are regional faults. Also shown are the locations of the White Water Recharge Facility (diamond), the study region (rectangle), and a triangular region representing the assumed maximum extent of pore-fluid pressure diffusion down-stream from the White Water Recharge Facility. b) Average land-surface uplift rate calculated across the model cross-section, which is chosen along the dashed line in (a). Fig. 3: Variously averaged (see text for explanation) land-surface uplift along the model cross-section in the Upper Coachella Valley spanning late-1993 to 2000: a) Average cumulative uplift profiles from InSAR time-series data plotted as a function of time. b) Temporal variations of sub-basin-wide spatial averages for cumulative land-surface uplift with standard error bars for each of the three sub-basins. c) Temporal variations of sub-basin-wide spatial averages for relative uplift with standard error bars for each of the three sub-basins estimated by subtracting the corresponding uplift of the Mission Creek sub-basin (MCB) from that of each sub-basin at each time step to remove/reduce seasonal variations.

35 Poroelastic Modeling in Coachella Valley, CA, USA 35 Fig. 4: a) Depth to groundwater, recorded by the Coachella Vally Water District, at wells near the model cross-section (Fig. 1) during late-1993 through 2000, i.e., the period over which InSAR data is measured (Sec. 3.2). W1 through W4 are wells located in the Mission Creek sub-basin (MCB), W5 is the well in the Garnet Hill sub-basin (GHB), and W5 through W9 are wells located in the White Water sub-basin (WWB). b) Artificial groundwater recharge, Q [m 3 /yr], introduced into Upper Coachella Valley from late-1993 to 1999 (CVWD 2002). Data for the year 2000 is not available. The horizontal dashed and solid lines represent average annual artificial groundwater recharge values, Q mean, during 1987 through 1993 and late-1993 through 1999, respectively. Q mean represents the net change in mean annual artificial groundwater recharge between the two time periods considered. Fig. 5: Illustration showing the model cross-section ( 10 vertically exaggerated) from SW to NE across the Upper Coachella Valley. The bottom, left, and right sides of the model cross-section are no-flow boundaries. The Garnet Hill Fault (GHF) and the Banning Fault (BF) mark the boundaries between the White Water sub-basin (WWB) and the Garnet Hill sub-baisn (GHB) and the GHB and the Mission Creek sub-basin (MCB), respectively. The diamond marks the approximate projection of the White Water spreading facility onto the model cross-section, where artificial groundwater recharge, Q, occurs. Fig. 6: Example simulations conducted using steady-state conditions. The water table (black line) is continuous within sub-basins and offset across faults: a) Simulated pore-fluid pressure distribution in the aquifer in late-1993 (initial state). b) Simulated groundwater flow lines in the aquifer in late c) Simulated pore-fluid pressure redistribution in the aquifer in 2000 (final state) resulting from the increase in the artificial groundwater recharge, Q c. d) Simulated groundwater flow lines in the aquifer in e) Modeled vertical stress field in the aquifer in late-1993 (initial state). f) Modeled z- displacement in the aquifer during late g) Modeled vertical stress field in the

36 36 Appana et al. aquifer in 2000 (final state). h) Modeled z-displacement in the aquifer in Fig. 7: a) Contour plot of misfit, M [mm/yr], from simulations exploring the parameter space of k wwb x [1] versus a wwb [2], where values of a wwb [2] decreasing further away from 1 indicate larger permeability anisotropies for the White Water sub-basin (WWB). The dots represent parameter combinations of kx wwb [1] and a wwb [2] at which simulations are run, keeping parameters [3]-[7] fixed (Table 4). The solid, dashed, and dot-dashed lines represent contours with minimized misfit (M 0.14 mm/yr) corresponding to simulation runs using Q c [3] [m 2 /yr] values of 22500, 35000, and 10000, respectively. b) Simulated land-surface uplift rate, u(x), as a function of k wwb x [1] at constant [2]-[7]. c) Simulated uplift rate, u(x), as a function of a wwb [2]=(k z /k x ) wwb at constant [1] and [3]-[7]. Fig. 8: a) Contour plot of misfit, M [mm/yr], from simulations exploring the parameter space of k ghf = k ghf x = kz ghf [4] versus a ghb [6], where values of a ghb [6] decreasing further away from 1 represent larger permeability anisotropies for the Garnet Hill subbasin (GHB). The dots represent parameter combinations of k ghf [4] and a ghb [6] for which simulations are run, keeping values of parameters [1]-[3], [5], and [7] fixed (Table 4). b) Contour plot of misfit, M, from simulations exploring the parameter space of k ghf [4] versus a ghb [6] at kx ghb [5] = m 2 and constant [1]-[3] and [7]. c) Simulated land-surface uplift rate, u(x), as a function of k ghf [4] at constant [1]-[3] and [5]-[7]. Fig. 9: Simulated land-surface uplift rate, u(x), as a function of a) kx ghb [5] at constant [1]-[4], [6], and [7] and b) a ghb [5]=(k z /k x ) ghb at constant [1]-[5] and [7] (Table 4). Fig. 10: Reduced misfit, δm = M M min, as a function of the permeability of the Banning Fault (BF), for simulations conducted at constant parameter values for [1]- [6] (Table 4). Six different sets of values are used for the constants. For simulations conducted with each of these sets, the corresponding minimum misfit is designated as

37 Poroelastic Modeling in Coachella Valley, CA, USA 37 M i min, where i=1 to 6. Fig. 11: a) Summary of parameter estimates from the present study, subscripted one (1), and comparison with values from previous studies; subscript two (2): Tyley (1974); Swain (1978); subscript three (3): Mayer et al. (2007). Note, only an upper limit is found for the permeability of the BF. b) Curve showing simulated uplift rate, u(x), using an optimal parameter set (Table 4). Fig. 12: Illustration showing an idealized cross-section of a fault zone with its two components, the fault core and the damage zone. We assume that 1) the fault core is of uniform width, c [m], along the length of the fault zone and 2) the damage zones, with a width of d [m] around the fault core, are symmetric. Variables are defined in Table 5. The width of the fault zone is b [m]. Fig. 13: The dots represent parameter combinations of k ghf [4] and a ghb [6] for which simulations are run, keeping values of parameters [1]-[2], [5], and [7] fixed (Table 4). Contour plots of misfit, M, from simulations exploring the parameter space of k ghf [4] versus a ghb [6] for Q c [3] values of a) m 2 /yr and b) m 2 /yr. Fig. 14: Contour plot of misfit, M [mm/yr], from simulations exploring the parameter space of kx ghb [5] versus a ghb [6], where values of a ghb [5] decreasing further away from 1 indicate larger permeability anisotropies. The dots represent parameter combinations of k ghb x [5] and a ghb [6] at which simulations are run, keeping parameters [1]-[4] and [7] fixed (Table 4).

38 38 Appana et al. Table 1. Definition of parameters used in this study. Symbol Units Value Definition a permeability anisotropy (k z /k x ) a mcb a ghb permeability anisotropy of the MCB b m 50 width of the faults (in simulations) D m 450 depth of the basin g m/s gravitational acceleration h m hydraulic head h o m initial hydraulic head k m 2 permeability tensor kx mcb m horizontal permeability of MCB K m/s hydraulic conductivity tensor L m 2500 length of the line source for groundwater recharge M mm/yr root mean-squared error (misfit) δm mm/yr reduced misfit n 73 number of observation points Q m 3 /yr amount of artificialgroundwater recharge Q s 1/s fluid source or sink per unit volume Q m 3 /yr net artificial groundwater recharge Q c m 2 /yr Q entering the model cross-section R(x, z) m (2600, 120) co-ordinates of the center of the line source (see L) s m displacement vector u mm/yr land-surface uplift rate ū mm/yr standard deviation of u Y GPa 1 Young s modulus α 0.75 Biot-Willis coefficient η Pa s dynamic viscosity of fluid ν 0.2 Poisson s ratio ρ f kg/m density of the fluid ρ s σ e kg/m 3 Pa 2500 density of the sediment effective stress tensor σ t Pa total stress tensor subscripts/ superscripts c fault core d damage zone of fault bf Banning Fault ghb Garnet Hill sub-basin ghf Garnet Hill Fault mcb Mission Creek sub-basin min minimum obs observed wwb White Water sub-basin x m horizontal dimension z m elevation (vertical)

39 Poroelastic Modeling in Coachella Valley, CA, USA 39 Table 2. Temporal and perpendicular baselines for interferometric pairs (columns 1 and 2) used in both stacking and time-series analysis of surface displacement. Master Slave b-perp yyyymmdd yyyymmdd [m] Table 3. Boundary conditions used in the 2D fluid-flow and displacement model across the Upper Coachella Valley. Here, h is the hydraulic head potential, h o is the water table elevation in late-1993, s is the displacement vector, and x and z indicate horizontal and vertical dimensions of the model cross-section, respectively. The surface, i.e., the top boundary, is free to move in both directions. Condition h/ x = 0 (no-flow) δs x = 0 h/ z = 0 (no-flow) δs x = δs z = 0 h o = h o (x) free δs x and δs z Boundary left/right left/right base base surface surface

40 40 Appana et al. Table 4. Values assigned to the seven parameters [numbered in square brackets] in different simulation runs as discussed in the main text. Both faults are considered to be isotropic, i.e., kx ghf = kz ghf = k ghf and kx bf = kz bf = k bf. Minimum misfit values, M min, achieved are also shown. [1] [2] [3] [4] [5] [6] [7] kx wwb a wwb Q c k ghf kx ghb a ghb k bf M min Fig. [m 2 ] [m 2 /yr] [m 2 ] [m 2 ] [m 2 ] [mm/yr] 7a b c a b c a b b a b

41 Poroelastic Modeling in Coachella Valley, CA, USA 41 Table 5. Estimates of fault core permeabilities of the GHF and the BF. Variables are: b is the width of the fault zone, k b is the bulk permeability of the fault zone (estimated from numerical simulations), c is the width of the fault core, k c is the permeability of the fault core, and k dl = k dr (denoted by k d ) are the permeabilities of the damage zones to the left and the right of the fault core. For the above calculations we consider the following values for k left and k right (see Fig. 12). GHF: k left = kx wwb = m 2 and k right = kx ghb = m 2. BF: k left = k ghb x = m 2 and k right = k mcb x = m 2 k b [m 2 ] k d [m 2 ] b [m] c [m] k c [m 2 ] comments GHF upper limit for k c k c does not depend on k d lower limit for k c see Appedix B k c does not depend on b BF upper limit for k c see Appendix B Table 6. Well number as used in this paper, official well identification number (CVWD), latitude, and longitude of the wells used in this study. Water level depths are shown in Figure 4a. Well No. Well ID Latitude Longitude 1 03S04E12B001S S04E12C001S S04E12F001S S04E12H002S S04E13N001S S04E20F001S S04E20J001S S04E29F001S S04E29R001S

42 42 Appana et al. Figure 1.

43 Poroelastic Modeling in Coachella Valley, CA, USA 43 a) Land surface uplift rate, u [mm/year] WWB GHF GHB b) x [km] Figure 2. BF Observed uplift, u obs Standard deviation MCB

44 44 Appana et al. a) Average land surface uplift [mm] WWB GHB MCB With seasonal variations 5 b) c) Relative land surface uplift, u [mm] Year WWB GHB MCB Reduced/no seasonal variations Year Figure 3.

45 Poroelastic Modeling in Coachella Valley, CA, USA 45 Depth to groundwater [m] W5 W1 W2 W3 W4 W6 W7 W8 W9 a) b) Q [m 3 /year] 2.5 x Net change, Q mean Total Recharge, Q Q mean ( ) Q mean ( ) 0.5? Year Figure 4. Figure 5.

46 46 Appana et al. a) b) c) d) e) f) g) h) Figure 6.

A THESIS SUBMITTED TO THE FACULTY OF THE GRADUATE SCHOOL OF THE UNIVERSITY OF MINNESOTA BY. Ravi Appana

A THESIS SUBMITTED TO THE FACULTY OF THE GRADUATE SCHOOL OF THE UNIVERSITY OF MINNESOTA BY. Ravi Appana Aquifer and Shallow San Andreas Fault Permeabilities Inferred from Poroelastic Modeling of InSAR Measurements of Land Surface Deformation in Coachella Valley, California A THESIS SUBMITTED TO THE FACULTY

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

16 Rainfall on a Slope

16 Rainfall on a Slope Rainfall on a Slope 16-1 16 Rainfall on a Slope 16.1 Problem Statement In this example, the stability of a generic slope is analyzed for two successive rainfall events of increasing intensity and decreasing

More information

Land subsidence due to groundwater withdrawal in Hanoi, Vietnam

Land subsidence due to groundwater withdrawal in Hanoi, Vietnam Land Subsidence (Proceedings of the Fifth International Symposium on Land Subsidence, The Hague, October 1995). 1AHS Publ. no. 234, 1995. 55 Land subsidence due to groundwater withdrawal in Hanoi, Vietnam

More information

Soils, Hydrogeology, and Aquifer Properties. Philip B. Bedient 2006 Rice University

Soils, Hydrogeology, and Aquifer Properties. Philip B. Bedient 2006 Rice University Soils, Hydrogeology, and Aquifer Properties Philip B. Bedient 2006 Rice University Charbeneau, 2000. Basin Hydrologic Cycle Global Water Supply Distribution 3% of earth s water is fresh - 97% oceans 1%

More information

1.72, Groundwater Hydrology Prof. Charles Harvey Lecture Packet #5: Groundwater Flow Patterns. Local Flow System. Intermediate Flow System

1.72, Groundwater Hydrology Prof. Charles Harvey Lecture Packet #5: Groundwater Flow Patterns. Local Flow System. Intermediate Flow System 1.72, Groundwater Hydrology Prof. Charles Harvey Lecture Packet #5: Groundwater Flow Patterns c Local Flow System 10,000 feet Intermediate Flow System Regional Flow System 20,000 feet Hydrologic section

More information

Measuring Water Resources Unit 3: Groundwater withdrawal and fluctuations in the height of Earth's surface student exercise

Measuring Water Resources Unit 3: Groundwater withdrawal and fluctuations in the height of Earth's surface student exercise Measuring Water Resources Unit 3: Groundwater withdrawal and fluctuations in the height of Earth's surface student exercise Eric Small (University of Colorado) and Bruce Douglas (Indiana University) Complete

More information

Seismic Reflection Imaging across the Johnson Ranch, Valley County, Idaho

Seismic Reflection Imaging across the Johnson Ranch, Valley County, Idaho Seismic Reflection Imaging across the Johnson Ranch, Valley County, Idaho Report Prepared for the Skyline Corporation Lee M. Liberty Center for Geophysical Investigation of the Shallow Subsurface (CGISS)

More information

11/22/2010. Groundwater in Unconsolidated Deposits. Alluvial (fluvial) deposits. - consist of gravel, sand, silt and clay

11/22/2010. Groundwater in Unconsolidated Deposits. Alluvial (fluvial) deposits. - consist of gravel, sand, silt and clay Groundwater in Unconsolidated Deposits Alluvial (fluvial) deposits - consist of gravel, sand, silt and clay - laid down by physical processes in rivers and flood plains - major sources for water supplies

More information

Groundwater Hydrology

Groundwater Hydrology EXERCISE 12 Groundwater Hydrology INTRODUCTION Groundwater is an important component of the hydrologic cycle. It feeds lakes, rivers, wetlands, and reservoirs; it supplies water for domestic, municipal,

More information

High-resolution temporal imaging of. Howard Zebker

High-resolution temporal imaging of. Howard Zebker High-resolution temporal imaging of crustal deformation using InSAR Howard Zebker Stanford University InSAR Prehistory SEASAT Topographic Fringes SEASAT Deformation ERS Earthquake Image Accurate imaging

More information

Seasonal subsidence and rebound in Las Vegas Valley, Nevada, observed by synthetic aperture radar interferometry

Seasonal subsidence and rebound in Las Vegas Valley, Nevada, observed by synthetic aperture radar interferometry WATER RESOURCES RESEARCH, VOL. 37, NO. 6, PAGES 1551 1566, JUNE 2001 Seasonal subsidence and rebound in Las Vegas Valley, Nevada, observed by synthetic aperture radar interferometry Jörn Hoffmann and Howard

More information

Reservoir Geomechanics and Faults

Reservoir Geomechanics and Faults Reservoir Geomechanics and Faults Dr David McNamara National University of Ireland, Galway david.d.mcnamara@nuigalway.ie @mcnamadd What is a Geological Structure? Geological structures include fractures

More information

Laboratory Exercise #4 Geologic Surface Processes in Dry Lands

Laboratory Exercise #4 Geologic Surface Processes in Dry Lands Page - 1 Laboratory Exercise #4 Geologic Surface Processes in Dry Lands Section A Overview of Lands with Dry Climates The definition of a dry climate is tied to an understanding of the hydrologic cycle

More information

Las Vegas Valley is a fault-bounded basin containing hundreds of metres of Tertiary and Quaternary sediments derived from lacustrine, paludal, and

Las Vegas Valley is a fault-bounded basin containing hundreds of metres of Tertiary and Quaternary sediments derived from lacustrine, paludal, and Land Subsidence (Proceedings of the Fourth International Symposium on Land Subsidence, May 1991). IAHS Publ. no. 200, 1991. Elevation Changes Associated with Subsidence in Las Vegas Valley, Nevada JOHN

More information

Inverse modeling using PS-InSAR for improved calibration of hydraulic parameters and prediction of future subsidence for Las Vegas Valley, USA

Inverse modeling using PS-InSAR for improved calibration of hydraulic parameters and prediction of future subsidence for Las Vegas Valley, USA doi:10.5194/piahs-372-411-2015 Author(s) 2015. CC Attribution 3.0 License. Inverse modeling using PS-InSAR for improved calibration of hydraulic parameters and prediction of future subsidence for Las Vegas

More information

Evolution of the conceptual hydrogeologic and ground-water flow model for Las Vegas Valley, Clark County, Nevada

Evolution of the conceptual hydrogeologic and ground-water flow model for Las Vegas Valley, Clark County, Nevada Evolution of the conceptual hydrogeologic and ground-water flow model for Las Vegas Valley, Clark County, Nevada Geological Society of America Annual Meeting November 14, 2 David J. Donovan Southern Nevada

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

ERS Track 98 SAR Data and InSAR Pairs Used in the Analysis

ERS Track 98 SAR Data and InSAR Pairs Used in the Analysis ERS Track 98 SAR Data and InSAR Pairs Used in the Analysis Date 1 Date 2 Date 1 Date 2 Date 1 Date 2 Date 1 Date 2 7/17/1992 6/19/2000 7/17/1992 7/2/1993 9/10/1993 10/28/1996 9/3/1995 10/18/1999 9/25/1992

More information

Tu P8 08 Modified Anisotropic Walton Model for Consolidated Siliciclastic Rocks: Case Study of Velocity Anisotropy Modelling in a Barents Sea Well

Tu P8 08 Modified Anisotropic Walton Model for Consolidated Siliciclastic Rocks: Case Study of Velocity Anisotropy Modelling in a Barents Sea Well Tu P8 08 Modified Anisotropic Walton Model for Consolidated Siliciclastic Rocks: Case Study of Velocity Anisotropy Modelling in a Barents Sea Well Y. Zhou (Rock Solid Images), F. Ruiz (Repsol), M. Ellis*

More information

Enabling Technologies

Enabling Technologies Enabling Technologies Mechanical Modelling 1 Key Parameter Mineral System Exploration is reflected in scale-dependent translation A. Gradient in hydraulic potential B. Permeability C. Solubility sensitivity

More information

Satellite radar interferometry time series analysis of surface deformation for Los Angeles, California

Satellite radar interferometry time series analysis of surface deformation for Los Angeles, California GEOPHYSICAL RESEARCH LETTERS, VOL. 31, L23613, doi:10.1029/2004gl021294, 2004 Satellite radar interferometry time series analysis of surface deformation for Los Angeles, California Riccardo Lanari, 1,2

More information

Surface Water and Stream Development

Surface Water and Stream Development Surface Water and Stream Development Surface Water The moment a raindrop falls to earth it begins its return to the sea. Once water reaches Earth s surface it may evaporate back into the atmosphere, soak

More information

Advanced interpretation of land subsidence by validating multiinterferometric SAR data: the case study of Anthemountas basin (Northern Greece)

Advanced interpretation of land subsidence by validating multiinterferometric SAR data: the case study of Anthemountas basin (Northern Greece) Revision of paper Advanced interpretation of land subsidence by validating multiinterferometric SAR data: the case study of Anthemountas basin (Northern Greece) By Raspini et al. General comments This

More information

RESISTIVITY IMAGING IN EASTERN NEVADA USING THE AUDIOMAGNETOTELLURIC METHOD FOR HYDROGEOLOGIC FRAMEWORK STUDIES. Abstract.

RESISTIVITY IMAGING IN EASTERN NEVADA USING THE AUDIOMAGNETOTELLURIC METHOD FOR HYDROGEOLOGIC FRAMEWORK STUDIES. Abstract. RESISTIVITY IMAGING IN EASTERN NEVADA USING THE AUDIOMAGNETOTELLURIC METHOD FOR HYDROGEOLOGIC FRAMEWORK STUDIES Darcy K. McPhee, U.S. Geological Survey, Menlo Park, CA Louise Pellerin, Green Engineering,

More information

DIFFERENTIAL INSAR STUDIES IN THE BOREAL FOREST ZONE IN FINLAND

DIFFERENTIAL INSAR STUDIES IN THE BOREAL FOREST ZONE IN FINLAND DIFFERENTIAL INSAR STUDIES IN THE BOREAL FOREST ZONE IN FINLAND Kirsi Karila (1,2), Mika Karjalainen (1), Juha Hyyppä (1) (1) Finnish Geodetic Institute, P.O. Box 15, FIN-02431 Masala, Finland, Email:

More information

RATE OF FLUID FLOW THROUGH POROUS MEDIA

RATE OF FLUID FLOW THROUGH POROUS MEDIA RATE OF FLUID FLOW THROUGH POROUS MEDIA Submitted by Xu Ming Xin Kiong Min Yi Kimberly Yip Juen Chen Nicole A project presented to the Singapore Mathematical Society Essay Competition 2013 1 Abstract Fluid

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

Topographically driven groundwater flow and the San Andreas heat flow paradox revisited

Topographically driven groundwater flow and the San Andreas heat flow paradox revisited University of Nebraska - Lincoln DigitalCommons@University of Nebraska - Lincoln USGS Staff -- Published Research US Geological Survey 2003 Topographically driven groundwater flow and the San Andreas heat

More information

Streams. Water. Hydrologic Cycle. Geol 104: Streams

Streams. Water. Hydrologic Cycle. Geol 104: Streams Streams Why study streams? Running water is the most important geologic agent in erosion, transportation and deposition of sediments. Water The unique physical and chemical properties of water make it

More information

Darcy's Law. Laboratory 2 HWR 531/431

Darcy's Law. Laboratory 2 HWR 531/431 Darcy's Law Laboratory HWR 531/431-1 Introduction In 1856, Henry Darcy, a French hydraulic engineer, published a report in which he described a series of experiments he had performed in an attempt to quantify

More information

Radar Remote Sensing: Monitoring Ground Deformations and Geohazards from Space

Radar Remote Sensing: Monitoring Ground Deformations and Geohazards from Space Radar Remote Sensing: Monitoring Ground Deformations and Geohazards from Space Xiaoli Ding Department of Land Surveying and Geo-Informatics The Hong Kong Polytechnic University A Question 100 km 100 km

More information

6298 Stress induced azimuthally anisotropic reservoir - AVO modeling

6298 Stress induced azimuthally anisotropic reservoir - AVO modeling 6298 Stress induced azimuthally anisotropic reservoir - AVO modeling M. Brajanovski* (Curtin University of Technology), B. Gurevich (Curtin University of Technology), D. Nadri (CSIRO) & M. Urosevic (Curtin

More information

' International Institute for Land Reclamation and Improvement. 2 Groundwater Investigations. N.A. de Ridder'? 2.1 Introduction. 2.

' International Institute for Land Reclamation and Improvement. 2 Groundwater Investigations. N.A. de Ridder'? 2.1 Introduction. 2. 2 Groundwater Investigations N.A. de Ridder'? 2.1 Introduction Successful drainage depends largely on a proper diagnosis of the causes of the excess water. For this diagnosis, one must consider: climate,

More information

Prof. Stephen A. Nelson EENS 111. Groundwater

Prof. Stephen A. Nelson EENS 111. Groundwater Page 1 of 8 Prof. Stephen A. Nelson EENS 111 Tulane University Physical Geology This page last updated on 20-Oct-2003 is water that exists in the pore spaces and fractures in rock and sediment beneath

More information

Which map shows the stream drainage pattern that most likely formed on the surface of this volcano? A) B)

Which map shows the stream drainage pattern that most likely formed on the surface of this volcano? A) B) 1. When snow cover on the land melts, the water will most likely become surface runoff if the land surface is A) frozen B) porous C) grass covered D) unconsolidated gravel Base your answers to questions

More information

IPMO2-1. Groundwater Modelling of Chiang Rai Basin, Northern Thailand. Sattaya Intanum* Dr.Schradh Saenton**

IPMO2-1. Groundwater Modelling of Chiang Rai Basin, Northern Thailand. Sattaya Intanum* Dr.Schradh Saenton** IPMO2-1 Groundwater Modelling of Chiang Rai Basin, Northern Thailand Sattaya Intanum* Dr.Schradh Saenton** ABSTRACT Chiang Rai basin, situated in Chiang Rai and Phayao provinces covering an area of 11,000

More information

Tensor character of pore pressure/stress coupling in reservoir depletion and injection

Tensor character of pore pressure/stress coupling in reservoir depletion and injection Tensor character of pore pressure/stress coupling in reservoir depletion and injection Müller, B., Altmann, J.B., Müller, T.M., Weißhardt, A., Shapiro, S., Schilling, F.R., Heidbach, O. Geophysical Institute

More information

Surface Processes Focus on Mass Wasting (Chapter 10)

Surface Processes Focus on Mass Wasting (Chapter 10) Surface Processes Focus on Mass Wasting (Chapter 10) 1. What is the distinction between weathering, mass wasting, and erosion? 2. What is the controlling force in mass wasting? What force provides resistance?

More information

Modeling pressure response into a fractured zone of Precambrian basement to understand deep induced-earthquake hypocenters from shallow injection

Modeling pressure response into a fractured zone of Precambrian basement to understand deep induced-earthquake hypocenters from shallow injection Modeling pressure response into a fractured zone of Precambrian basement to understand deep induced-earthquake hypocenters from shallow injection S. Raziperchikolaee 1 and J. F. Miller 1 Abstract Analysis

More information

The financial and communal impact of a catastrophe instantiated by. volcanoes endlessly impact on lives and damage expensive infrastructure every

The financial and communal impact of a catastrophe instantiated by. volcanoes endlessly impact on lives and damage expensive infrastructure every Chapter 1 Introduction The financial and communal impact of a catastrophe instantiated by geophysical activity is significant. Landslides, subsidence, earthquakes and volcanoes endlessly impact on lives

More information

mountain rivers fixed channel boundaries (bedrock banks and bed) high transport capacity low storage input output

mountain rivers fixed channel boundaries (bedrock banks and bed) high transport capacity low storage input output mountain rivers fixed channel boundaries (bedrock banks and bed) high transport capacity low storage input output strong interaction between streams & hillslopes Sediment Budgets for Mountain Rivers Little

More information

NOTES AND CORRESPONDENCE Segmented Faulting Process of Chelungpu Thrust: Implication of SAR Interferograms

NOTES AND CORRESPONDENCE Segmented Faulting Process of Chelungpu Thrust: Implication of SAR Interferograms , Vol. 14, No.2, 241-247, June 2003 NOTES AND CORRESPONDENCE Segmented Faulting Process of Chelungpu Thrust: Implication of SAR Interferograms Chien-Chih Chen 1,*, Chung-Pai Chang 2, and Kun-Shan Chen

More information

RADAR Remote Sensing Application Examples

RADAR Remote Sensing Application Examples RADAR Remote Sensing Application Examples! All-weather capability: Microwave penetrates clouds! Construction of short-interval time series through cloud cover - crop-growth cycle! Roughness - Land cover,

More information

Effects of delayed drainage on subsidence modeling and parameter estimation

Effects of delayed drainage on subsidence modeling and parameter estimation Effects of delayed drainage on subsidence modeling and parameter estimation Tingting Yan Thesis submitted to the faculty of the Virginia Polytechnic Institute and State University in partial fulfillment

More information

Blanca Payàs 1, Ryan Morris 2, Alain Arnaud 1, David Albiol 1, John Bradshaw 3, Javier Duro 1. Abstract

Blanca Payàs 1, Ryan Morris 2, Alain Arnaud 1, David Albiol 1, John Bradshaw 3, Javier Duro 1. Abstract Surface monitoring using radar satellite images to measure ground deformation: applied to Coal Seam Gas production areas to comply with regulatory requirements Blanca Payàs 1, Ryan Morris 2, Alain Arnaud

More information

Faults. Strike-slip fault. Normal fault. Thrust fault

Faults. Strike-slip fault. Normal fault. Thrust fault Faults Strike-slip fault Normal fault Thrust fault Fault any surface or narrow zone with visible shear displacement along the zone Normal fault Strike-slip fault Reverse fault Thrust fault

More information

Essentials of Geology, 11e

Essentials of Geology, 11e Essentials of Geology, 11e Groundwater Chapter 10 Instructor Jennifer Barson Spokane Falls Community College Geology 101 Stanley Hatfield Southwestern Illinois Co Jennifer Cole Northeastern University

More information

GG655/CEE623 Groundwater Modeling. Aly I. El-Kadi

GG655/CEE623 Groundwater Modeling. Aly I. El-Kadi GG655/CEE63 Groundwater Modeling Model Theory Water Flow Aly I. El-Kadi Hydrogeology 1 Saline water in oceans = 97.% Ice caps and glaciers =.14% Groundwater = 0.61% Surface water = 0.009% Soil moisture

More information

San Gorgonio Pass Special Fault Study Area

San Gorgonio Pass Special Fault Study Area San Gorgonio Pass Special Fault Study Area Michele Cooke, David Oglesby and Doug Yule Whitney Behr, Kim Blisniuk, Jim Brune, Sarah Carena, Judi Chester, Gary Fuis, Thomas Goebel, Peter Gold, Egill Hauksson,

More information

2. Governing Equations. 1. Introduction

2. Governing Equations. 1. Introduction Multiphysics Between Deep Geothermal Water Cycle, Surface Heat Exchanger Cycle and Geothermal Power Plant Cycle Li Wah Wong *,1, Guido Blöcher 1, Oliver Kastner 1, Günter Zimmermann 1 1 International Centre

More information

Slow Deformation of Mt. Baekdu Stratovolcano Observed by Satellite Radar Interferometry

Slow Deformation of Mt. Baekdu Stratovolcano Observed by Satellite Radar Interferometry Slow Deformation of Mt. Baekdu Stratovolcano Observed by Satellite Radar Interferometry Sang-Wan Kim and Joong-Sun Won Department of Earth System Sciences, Yonsei University 134 Shinchon-dong, Seodaemun-gu,

More information

,Baynes Lake. TO...?&.?...A 2...KO.?'!!&... Sr. *logical Engineer

,Baynes Lake. TO...?&.?...A 2...KO.?'!!&... Sr. *logical Engineer > i evernment OF BRITISH COLUMBIA a TO...?&.?...A 2....KO.?'!!&... Sr. *logical Engineer... Grou,,water. Section Hydrology Division Wat.er... In~.~s.tiga.ti.On.s..Branck.... 5 u BJECT...C;.roun.dw.ater...Snve

More information

In all of the following equations, is the coefficient of permeability in the x direction, and is the hydraulic head.

In all of the following equations, is the coefficient of permeability in the x direction, and is the hydraulic head. Groundwater Seepage 1 Groundwater Seepage Simplified Steady State Fluid Flow The finite element method can be used to model both steady state and transient groundwater flow, and it has been used to incorporate

More information

FUNDAMENTALS OF ENGINEERING GEOLOGY

FUNDAMENTALS OF ENGINEERING GEOLOGY FUNDAMENTALS OF ENGINEERING GEOLOGY Prof. Dr. HUSSEIN HAMEED KARIM Building and Construction Engineering Department 2012 Preface The impulse to write this book stemmed from a course of geology given by

More information

Inverse modeling of interbed storage parameters using land subsidence observations, Antelope Valley, California

Inverse modeling of interbed storage parameters using land subsidence observations, Antelope Valley, California WATER RESOURCES RESEARCH, VOL. 39, NO. 2, 1031, doi:10.1029/2001wr001252, 2003 Inverse modeling of interbed storage parameters using land subsidence observations, Antelope Valley, California Jörn Hoffmann

More information

Hydraulic properties of porous media

Hydraulic properties of porous media PART 5 Hydraulic properties of porous media Porosity Definition: Void space: n V void /V total total porosity e V void /V solid Primary porosity - between grains Secondary porosity - fracture or solution

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

InSAR-based hydrology of the Everglades, South Florida

InSAR-based hydrology of the Everglades, South Florida InSAR-based hydrology of the Everglades, South Florida Shimon Wdowinski (1), Falk Amelung (1), Fernando Miralles-Wilhelm (2), Tim Dixon (1), and Richard Carande (3) (1) Division of Marine Geology and Geophysics,

More information

Napa Valley Groundwater Sustainability: A Basin Analysis Report for the Napa Valley Subbasin

Napa Valley Groundwater Sustainability: A Basin Analysis Report for the Napa Valley Subbasin Napa Valley Groundwater Sustainability: A Basin Analysis Report for the Napa Valley Subbasin A report prepared pursuant to California Water Code Section 10733.6(b)(3) EXECUTIVE SUMMARY (354.4(A)) 1 1.0

More information

Title: Application and use of near-wellbore mechanical rock property information to model stimulation and completion operations

Title: Application and use of near-wellbore mechanical rock property information to model stimulation and completion operations SPE OKC Oil and Gas Symposium March 27-31, 2017 Best of OKC Session Chairperson: Matthew Mower, Chaparral Energy Title: Application and use of near-wellbore mechanical rock property information to model

More information

EVALUATING HEAT FLOW AS A TOOL FOR ASSESSING GEOTHERMAL RESOURCES

EVALUATING HEAT FLOW AS A TOOL FOR ASSESSING GEOTHERMAL RESOURCES PROCEEDINGS, Thirtieth Workshop on Geothermal Reservoir Engineering Stanford University, Stanford, California, January 31-February 2, 2005 SGP-TR-176 EVALUATING HEAT FLOW AS A TOOL FOR ASSESSING GEOTHERMAL

More information

PUBLICATIONS. Water Resources Research. Estimating temporal changes in hydraulic head using InSAR data in the San Luis Valley, Colorado

PUBLICATIONS. Water Resources Research. Estimating temporal changes in hydraulic head using InSAR data in the San Luis Valley, Colorado PUBLICATIONS Water Resources Research RESEARCH ARTICLE 1.12/213WR14938 Key Points: InSAR data can be used to interpolate/extrapolate hydraulic head data in time Kriging allowed us to calibrate directly

More information

Mohr's Circle and Earth Stress (The Elastic Earth)

Mohr's Circle and Earth Stress (The Elastic Earth) Lect. 1 - Mohr s Circle and Earth Stress 6 Mohr's Circle and Earth Stress (The Elastic Earth) In the equations that we derived for Mohr s circle, we measured the angle, θ, as the angle between σ 1 and

More information

IN-SITU STRESS ESTIMATION IN OFFSHORE EASTERN MEDITERRANEAN WITH FINITE ELEMENT ANALYSIS

IN-SITU STRESS ESTIMATION IN OFFSHORE EASTERN MEDITERRANEAN WITH FINITE ELEMENT ANALYSIS 8 th GRACM International Congress on Computational Mechanics Volos, 12 July 15 July 2015 IN-SITU STRESS ESTIMATION IN OFFSHORE EASTERN MEDITERRANEAN WITH FINITE ELEMENT ANALYSIS Anna Kyriacou 1, Panos

More information

Fault Specific, Dynamic Rupture Scenarios for Strong Ground Motion Prediction

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

More information

Impact : Changes to Existing Topography (Less than Significant)

Impact : Changes to Existing Topography (Less than Significant) 4.2 Land Resources 4.2.1 Alternative A Proposed Action Impact 4.2.1-1: Changes to Existing Topography (Less than Significant) Development of the project site would involve grading and other earthwork as

More information

Table of Contents Chapter 1 Introduction to Geotechnical Engineering 1.1 Geotechnical Engineering 1.2 The Unique Nature of Soil and Rock Materials

Table of Contents Chapter 1 Introduction to Geotechnical Engineering 1.1 Geotechnical Engineering 1.2 The Unique Nature of Soil and Rock Materials Table of Contents Chapter 1 Introduction to Geotechnical Engineering 1.1 Geotechnical Engineering 1.2 The Unique Nature of Soil and Rock Materials 1.3 Scope of This Book 1.4 Historical Development of Geotechnical

More information

Case Study: University of Connecticut (UConn) Landfill

Case Study: University of Connecticut (UConn) Landfill Case Study: University of Connecticut (UConn) Landfill Problem Statement:» Locate disposal trenches» Identify geologic features and distinguish them from leachate and locate preferential pathways in fractured

More information

Water in Soil Sections in Craig

Water in Soil Sections in Craig Water in Soil Sections 2.1-2.6 in Craig Outlines Introduction Darcy s Law Volume of water flowing per unit time Measuring K in laboratory Seepage Theory Flow Net Introduction All soils are permeable materials,

More information

Instructor : Dr. Jehad Hamad. Chapter (7)

Instructor : Dr. Jehad Hamad. Chapter (7) Instructor : Dr. Jehad Hamad Chapter (7) 2017-2016 Soil Properties Physical Properties Mechanical Properties Gradation and Structure Compressibility Soil-Water Relationships Shear Strength Bearing Capacity

More information

FORENSIC GEOLOGY A CIVIL ACTION

FORENSIC GEOLOGY A CIVIL ACTION NAME 89.215 - FORENSIC GEOLOGY A CIVIL ACTION I. Introduction In 1982 a lawsuit was filed on behalf of eight Woburn families by Jan Schlictmann. The suit alleged that serious health effects (childhood

More information

The Potential of High Resolution Satellite Interferometry for Monitoring Enhanced Oil Recovery

The Potential of High Resolution Satellite Interferometry for Monitoring Enhanced Oil Recovery The Potential of High Resolution Satellite Interferometry for Monitoring Enhanced Oil Recovery Urs Wegmüller a Lutz Petrat b Karsten Zimmermann c Issa al Quseimi d 1 Introduction Over the last years land

More information

Lab 5 - Aquifer Elasticity and Specific Storage. Due October 12 14, 2010

Lab 5 - Aquifer Elasticity and Specific Storage. Due October 12 14, 2010 Lab 5 - Aquifer Elasticity and Specific Storage Due October 12 14, 2010 The goal of this experiment is to measure the specific storage S s of a balloon, which simulates aquifer elasticity. The experiment

More information

4D stress sensitivity of dry rock frame moduli: constraints from geomechanical integration

4D stress sensitivity of dry rock frame moduli: constraints from geomechanical integration Title 4D stress sensitivity of dry rock frame moduli: constraints from geomechanical integration Authors Bloomer, D., Ikon Science Asia Pacific Reynolds, S., Ikon Science Asia Pacific Pavlova, M., Origin

More information

Geochemistry, Geophysics, Geosystems. Supporting Information for

Geochemistry, Geophysics, Geosystems. Supporting Information for 1 2 3 4 5 6 7 8 Geochemistry, Geophysics, Geosystems Supporting Information for Volcano Deformation Survey over the Northern and Central Andes with ALOS InSAR Time Series Anieri M. Morales Rivera 1, Falk

More information

UGRC 144 Science and Technology in Our Lives/Geohazards

UGRC 144 Science and Technology in Our Lives/Geohazards UGRC 144 Science and Technology in Our Lives/Geohazards Flood and Flood Hazards Dr. Patrick Asamoah Sakyi Department of Earth Science, UG, Legon College of Education School of Continuing and Distance Education

More information

WATER ON AND UNDER GROUND. Objectives. The Hydrologic Cycle

WATER ON AND UNDER GROUND. Objectives. The Hydrologic Cycle WATER ON AND UNDER GROUND Objectives Define and describe the hydrologic cycle. Identify the basic characteristics of streams. Define drainage basin. Describe how floods occur and what factors may make

More information

From 2D Seismic to Hydrodynamic Modelling

From 2D Seismic to Hydrodynamic Modelling From 2D Seismic to Hydrodynamic Modelling A Complex Study of Turbidites in a Petroleum-Bearing Basin Marcell Lux, HC Exploration Geologist Ahmed Amran, HC Exploration Advisor Marianna Vincze, Geomodel

More information

1. Water in Soils: Infiltration and Redistribution

1. Water in Soils: Infiltration and Redistribution Contents 1 Water in Soils: Infiltration and Redistribution 1 1a Material Properties of Soil..................... 2 1b Soil Water Flow........................... 4 i Incorporating K - θ and ψ - θ Relations

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

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

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

Analyzing effect of fluid flow on surface subsidence

Analyzing effect of fluid flow on surface subsidence Analyzing effect of fluid flow on surface subsidence in mining area Y. Abousleiman", M. Bai\ H. Zhang', T. Liu" and J.-C. Roegiers* a. School of Engineering and Architecture, The Lebanese American University,

More information

MEASUREMENT OF SURFACE DISPLACEMENT CAUSED BY UNDERGROUND NUCLEAR EXPLOSIONS BY DIFFERENTIAL SAR INTERFEROMETRY

MEASUREMENT OF SURFACE DISPLACEMENT CAUSED BY UNDERGROUND NUCLEAR EXPLOSIONS BY DIFFERENTIAL SAR INTERFEROMETRY MEASUREMENT OF SURFACE DISPLACEMENT CAUSED BY UNDERGROUND NUCLEAR EXPLOSIONS BY DIFFERENTIAL SAR INTERFEROMETRY X. Cong a, KH. Gutjahr b, J. Schlittenhardt a, U. Soergel c, a Bundesanstalt für Geowissenschaften

More information

1. classic definition = study of deformed rocks in the upper crust

1. classic definition = study of deformed rocks in the upper crust Structural Geology I. Introduction 1. classic definition = study of deformed rocks in the upper crust deformed includes translation, rotation, and strain (change of shape) All rocks are deformed in some

More information

ES 105 Surface Processes I. Hydrologic cycle A. Distribution % in oceans 2. >3% surface water a. +99% surface water in glaciers b.

ES 105 Surface Processes I. Hydrologic cycle A. Distribution % in oceans 2. >3% surface water a. +99% surface water in glaciers b. ES 105 Surface Processes I. Hydrologic cycle A. Distribution 1. +97% in oceans 2. >3% surface water a. +99% surface water in glaciers b. >1/3% liquid, fresh water in streams and lakes~1/10,000 of water

More information

Deep Borehole Disposal Performance Assessment and Criteria for Site Selection

Deep Borehole Disposal Performance Assessment and Criteria for Site Selection Deep Borehole Disposal Performance Assessment and Criteria for Site Selection Sandia is a multiprogram laboratory operated by Sandia Corporation, a Lockheed Martin Company, for the United States Department

More information

ENVIRONMENTAL EFFECTS OF GROUNDWATER WITHDRAWAL IN SOUTH NYÍRSÉG

ENVIRONMENTAL EFFECTS OF GROUNDWATER WITHDRAWAL IN SOUTH NYÍRSÉG PhD thesis ENVIRONMENTAL EFFECTS OF GROUNDWATER WITHDRAWAL IN SOUTH NYÍRSÉG János Szanyi Szeged, 2004 ENVIRONMENTAL EFFECTS OF GROUNDWATER WITHDRAWAL IN SOUTH NYÍRSÉG Preliminaries, the aims of the dissertation

More information

predictive mineral discovery*cooperative Research Centre A legacy for mineral exploration science Mineral Systems Q4 Fluid flow drivers & pathways

predictive mineral discovery*cooperative Research Centre A legacy for mineral exploration science Mineral Systems Q4 Fluid flow drivers & pathways Mineral Systems Q4 Fluid flow drivers & pathways 1 Key Parameter Mineral System Exploration is reflected in scale-dependent translation A. Gradient in hydraulic potential B. Permeability C. Solubility

More information

Finding Large Capacity Groundwater Supplies for Irrigation

Finding Large Capacity Groundwater Supplies for Irrigation Finding Large Capacity Groundwater Supplies for Irrigation December 14, 2012 Presented by: Michael L. Chapman, Jr., PG Irrigation Well Site Evaluation Background Investigation Identify Hydrogeologic Conditions

More information

4600 Rickenbacker Cswy, Miami 33149, FL. U.S.A., (2)

4600 Rickenbacker Cswy, Miami 33149, FL. U.S.A., (2) POSTSEISMIC DEFORMATION FOLLOWING THE 2010 HAITI EARTHQUAKE: TIME-DEPENDENT SURFACE SUBSIDENCE INDUCED BY GROUNDWATER FLOW IN RESPONSE TO A SUDDEN UPLIFT Shimon Wdowinski (1), Sang-Hoon Hong (1)(2) (1)

More information

Ground deformation in Thessaly, Central Greece, between 1992 and 2000 by means of ERS multi-temporal InSAR

Ground deformation in Thessaly, Central Greece, between 1992 and 2000 by means of ERS multi-temporal InSAR INGV Ground deformation in Thessaly, Central Greece, between 1992 and 2000 by means of ERS multi-temporal InSAR S. Atzori (1), C. Tolomei (1), S. Salvi (1), A. Ganas (2), S. Stramondo (1) and L. Colini

More information

IV. ENVIRONMENTAL IMPACT ANALYSIS G. GEOLOGY AND SOILS

IV. ENVIRONMENTAL IMPACT ANALYSIS G. GEOLOGY AND SOILS IV. ENVIRONMENTAL IMPACT ANALYSIS G. GEOLOGY AND SOILS The following section is a summary of the geotechnical report conducted for the proposed project. The Report of Geotechnical Investigation Proposed

More information

3D temporal evolution of displacements recorded on Mt. Etna from the 2007 to 2010 through the SISTEM method

3D temporal evolution of displacements recorded on Mt. Etna from the 2007 to 2010 through the SISTEM method 3D temporal evolution of displacements recorded on Mt. Etna from the 2007 to 2010 through the SISTEM method Bonforte A., Guglielmino F.,, Puglisi G. INGV Istituto Nazionale di Gofisica e vulcanologia Osservatorio

More information

Uncertainties in rock pore compressibility and effects on time lapse seismic modeling An application to Norne field

Uncertainties in rock pore compressibility and effects on time lapse seismic modeling An application to Norne field Uncertainties in rock pore compressibility and effects on time lapse seismic modeling An application to Norne field Amit Suman and Tapan Mukerji Department of Energy Resources Engineering Stanford University

More information

Lecture 15: Subsidence

Lecture 15: Subsidence Lecture 15: Subsidence Key Questions 1. How does removal of groundwater cause subsidence on a regional scale? 2. Under what conditions does a building sink into sediment? 3. Why do clays consolidate more

More information

Geotechnical Properties of Soil

Geotechnical Properties of Soil Geotechnical Properties of Soil 1 Soil Texture Particle size, shape and size distribution Coarse-textured (Gravel, Sand) Fine-textured (Silt, Clay) Visibility by the naked eye (0.05 mm is the approximate

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

Constrained Fault Construction

Constrained Fault Construction Constrained Fault Construction Providing realistic interpretations of faults is critical in hydrocarbon and mineral exploration. Faults can act as conduits or barriers to subsurface fluid migration and

More information