<<

GEOPHYSICS, VOL. 80, NO. 3 (MAY-JUNE 2015); P. WB11–WB19, 5 FIGS., 3 TABLES. 10.1190/GEO2014-0119.1

Soil , elasticity, and the -water characteristic curve inverted from field-based seismic P- and S-wave velocity in shallow nearly saturated layered

Jie Shen1, Juan M. Lorenzo1, Chris D. White2, and Frank Tsai3

ABSTRACT soil property values by minimizing the misfit between field-based velocity profiles and predicted velocity profiles based on the Soil density, , elastic moduli, and the soil-water char- Hertz-Mindlin and Biot-Gassmann theories. The results from acteristic curve (SWCC) are important properties for soil charac- seismic soil property inversion are validated by comparison with terization. However, geotechnical and laboratory tests for soil geotechnical as as laboratory results conducted independ- properties are costly and limited to point sampling sites. Seismic ently in the same area as the seismic survey. For each seismically surveys can provide laterally continuous, seismic soil property recognizable layer, soil types are interpreted from the inverted soil values that may complement geotechnical tests with less density and elasticity, aided by the SWCC to help detect thin units cost. We have developed a workflow to invert for soil properties that are below the original seismic resolution of the field data. and the SWCC from seismic P- and S-wave velocity-versus-depth There is flexibility to apply our suggested workflow in future profiles interpreted from shallow (<25 m depth) unconsolidated studies. For a known geologic setting, empirical relationships sediments under conditions of near-full saturation (>99%). The and other velocity prediction models could also be incorporated inversion is performed by using the covariance matrix adaptation into the suggested workflow to improve inversion results and evolution strategy to search automatically for optimal input extract additional information in soils.

INTRODUCTION In this paper, we use an indirect inversion process to determine in situ soil density, elastic moduli, porosity, and the SWCC by min- Soil properties such as density, elastic moduli, porosity, and the imizing the misfit (cost function) between the predicted and field- soil-water characteristic curve (SWCC) are important for assessing derived velocity profiles (Figure 1). To perform the inversion, an stability (Bell, 1992), and monitoring contaminant optimization technique automatically searches for input soil-prop- movement and soil aeration (Terzaghi, 1996). These soil properties erty-parameter values that can best explain the field velocity pro- depend on soil grain size, mineral composition, overburden pres- sure, and stress history (Fredlund and Xing, 1994). Soil properties files. Compared with other seismic inversion techniques (such as can be measured directly in the laboratory (Van Genuchten, 1980), the widely used full-waveform inversion), the major advantages but these tests are costly, and the necessary equipment may not be of our inversion are that it uses a global optimization technique readily accessible. Laboratory soil property tests are performed on and requires only a velocity-versus-depth profile, which is inter- either core or bulk sediment samples, which may not be represen- preted from the seismic survey. A global optimization technique tative of in situ sediments. The borehole locations are usually distant searches for a global minimal value of the misfit throughout the from each other (>100’s of meters), so that lateral soil character- input parameter range. Unlike local optimization, which is used istics between are difficult to predict. by full-waveform inversion, global optimization is not affected Downloaded 10/17/16 to 130.39.188.137. Redistribution subject SEG license or copyright; see Terms of Use at http://library.seg.org/

Manuscript received by the Editor 9 March 2014; revised manuscript received 6 September 2014; published online 23 April 2015. 1Louisiana State University, Department of and Geophysics, Baton Rouge, Louisiana, USA. E-mail: [email protected]; [email protected]. 2Louisiana State University, Craft and Hawkins Department of Petroleum Engineering, Baton Rouge, Louisiana, USA. E-mail: [email protected]. 3Louisiana State University, Department of Civil and Environmental Engineering, Baton Rouge, Louisiana, USA. E-mail: [email protected]. © 2015 Society of Exploration Geophysicists. All rights reserved.

WB11 WB12 Shen et al.

by starting parameters. Currently, full-waveform inversion often re- (>10 kHz) in unconsolidated soil (Wang, 2001). Seismic velocity lies on reflections (Masoni et al., 2014), which can be hard to get in is ultimately computed from the relationships between velocity, shallow near-saturated soils because of attenuation, especially for P- effective elasticity, and bulk density (e.g., Ikelle and Amund- waves. For our inversion, we incorporate refraction and reflection sen, 2005). first arrivals through 1D P- and S-wave velocity-versus-depth start- The application of the Hertz-Mindlin model in -dominant ing models (Lorenzo et al., 2014). Moreover, the inversion of 1D sediments is debatable because the model is originally derived using velocity models is simpler and faster computationally than 2D or spherical grains. For example, shear modulus is overestimated by 3D waveform matching. Hertz-Mindlin for angular grains (Bachrach et al., 2000) and clay The SWCC shows the relationship between water saturation and grains have a platy shape. One heuristic approach is to lower the the capillary head (Van Genuchten, 1980), both of which are readily coordination number for computing predicted velocities of angular determined in our inversion. At the same capillary head, water sat- grains (Bachrach et al., 1998; Velea et al., 2000). Particularly in uration increases with soil plasticity (Fredlund and Xing, 1994). shallow unconsolidated sediments, loosely packed, highly angular Therefore, water saturation is lower in sandy soils than in clayey grains lead to a lower coordination number and higher porosity. As soils. The SWCC is used to estimate other soil behavior parameters a result, high-porosity (∼65%) clay has a relatively lower co- such as unsaturated , permeability, and pore-size dis- ordination number than medium-porosity (∼40%) (Murphy, tribution (Fredlund and Rahardjo, 1993a; Fredlund, 1995; Fredlund 1982). The Hertz-Mindlin theory predicts elastic moduli success- et al., 1996). fully in clay-dominant rocks, such as shales (Avseth et al., 2005) Our velocity prediction model is based on the Hertz-Mindlin con- and claystones (Takahashi and Tanaka, 2009). tact theory (Hertz, 1882; Mindlin, 1949) and the Biot-Gassmann Additional debate surrounds the application of a high Poisson’s fluid substitution theory (Gassmann, 1951; Biot, 1962) because ratio (>0.25) with the Hertz-Mindlin theory in wet sediments. In they are the currently acceptable constitutive models used to relate saturated unconsolidated (e.g., Dvorkin and Nur, 1996), as effective elasticity and soil properties in shallow unconsolidated well as in almost fully saturated clays (such as in our case), the sediments (Bachrach et al., 1998; Bachrach and Nur, 1998; Bach- effective Poisson’s ratio can be larger than 0.25. Dvorkin and rach et al., 2000; Zimmer et al., 2006; Bachrach and Avseth, 2008; Nur (1996) use the Hertz-Mindlin theory to predict velocities suc- Dutta et al., 2010). The Hertz-Mindlin model uses grain elasticity, cessfully in saturated loose sands with a high effective Poisson’s porosity, and grain contact geometry under (Digby, ratio close to 0.5. 1981; Walton, 1987; Dvorkin and Nur, 1996) to calculate the iso- Another factor that needs to be considered for clay at shallow tropic elastic moduli of a homogeneous, granular matrix comprising depth when using the Hertz-Mindlin model is the effective stress. identical spherical grains. The Hertz-Mindlin model also assumes In sand, the dominant stresses are and matric

dry, no-slip, spherical contacts between the grains, and as a result, suction ( is close to 0), whereas in clay, cohesion (up to Poisson’s ratio for the grains is expected to be <0.25. The Biot- 20 kPa) also plays a significant role. In this study, we modify the Gassmann fluid substitution theory accounts for pore fluid variation effective stress by also incorporating the effect of matric suction in a porous medium and estimates effective elastic moduli from dry and cohesion. This modification helps to predict velocities in agree- matrix elasticity, grain elasticity, and water saturation. The Biot- ment with field velocities in clay-and-sand mixed soils (Crane, 2013). Gassmann theory more accurately predicts velocities for data Among the various optimization algorithms used to effectively at low frequencies (10–200 Hz) than at higher frequencies search for those parameters that explain field velocity profiles, we use the covariance matrix adaptation evolution strategy (CMA-ES) (Hansen, 2011). CMA-ES belongs to the class of evolutionary algorithms, and it is a stochastic, derivative-free algorithm used for nonlinear local and global optimization (Mezura-Montes and Coello Coello, 2011). One of the advantages in CMA-ES over the ge- netic algorithm is its well-designed, internal, parameter-tuning mechanism that selects new candidates for input parameter values while ap- proaching a best fit between prediction and ob- servation. The parameter-tuning mechanism is based on updating the covariance matrix between variables in the distribution (Hansen, 2011)as the candidate values converge toward the global optimum. In CMA-ES, population size is crucial to the success of global optimization — this

Downloaded 10/17/16 to 130.39.188.137. Redistribution subject SEG license or copyright; see Terms of Use at http://library.seg.org/ number, by increasing logarithmically with the number of unknown inputs, is designed to avoid a local optimum (Hansen, 2006). CMA-ES is ap- Figure 1. The procedure of soil property inversion by minimizing the misfit between predicted and field-based seismic velocities. Predicted velocities are calculated by the plied to model fluid flow (Bayer et al., 2009) and Hertz-Mindlin and Biot-Gassmann models. Covariance matrix adaptation evolution to facilitate groundwater remediation (Bayer and strategy (CMA-ES) is the optimization program to minimize misfit. Finkel, 2007). Seismic soil property inversion WB13

We attempt our inversion for seismic soil properties in almost Table 1. Key parameters with known values used in the fully saturated soils of the lower Mississippi River swamps and Hertz-Mindlin and Biot-Gassmann models (Appendix A). marshes (Lorenzo et al., 2014) because the seismic velocity is most sensitive to water saturation at near-saturation conditions (water sat- Parameters Values uration >99%). Near-saturated soils occur permanently or season- ally in wetlands, including coastal, floodplains, the margins of Water table (m) 36 (assumed for calculation) lakes, and other areas below groundwater level or where precipita- g (m∕s2) 9.80665 tion is sufficiently high (Gilman, 1994). The velocity prediction ρ ∕ 3 models (e.g., Bachrach and Nur, 1998) and field-based velocity pro- w (kg m )1 ρ ∕ 3 × −3 files (e.g., Grelle and Guadagno, 2009) show that P-wave velocities a (kg m ) 1.18 10 9 change from ∼200 to ∼1500 m∕s in the transition zone from near- Kw (Pa) 2.2 × 10 5 saturated soils to saturated soils. As the soil approaches full satu- Ka (Pa) 1.01 × 10 ration, the pore water substitutes almost all of the air. The bulk modulus of water is more than four orders of magnitude larger than that of air (Table 1). As a result, higher water saturation yields a stiffer soil with an increased compressional velocity. Because S- waves are insensitive to pore fluids and travel more slowly than P-waves, S-waves are less attenuated by fluids and have more res- olution (Harris, 2009). As a result, S-waves are sensitive to soil type changes (Hayashi et al., 2013) and S-wave reflections can be used to identify soil layers.

METHODS Field-based data The field data we use for inversion include field-based P- and S-wave velocity-versus-depth profiles interpreted from seismic sur- veys at two field sites (Figure 2). The seismic survey is conducted in

the almost fully saturated shallow soil (<25 m) beneath a New Or- leans flood-protection levee (Lorenzo et al., 2014). In these data, the seismic frequency reaches 30 Hz. In velocity-versus-depth profiles, there is 1merror in layer depths (Lorenzo et al., 2014) and 2% error in P- and S-wave velocities. The seismic survey line is ∼100 m wide. As a result, the velocity profiles represent the aver- age of the survey width. Two sites (A and B) help validate the in- version results under a variety of water-saturation conditions and sand and clay percentages. At both sites, the soil is composed of a majority of clay with several small sand units between the depths of 7.5 and 15 m, and these sand units are thicker at site A than at site B(Lorenzo et al., 2014).

Velocity prediction model The Hertz-Mindlin contact theory and Biot-Gassmann fluid sub- stitution theory (Appendix A) predict seismic velocities from soil properties, water saturation, and effective stress (Figure 1). Water saturation and effective stress are major factors that contribute to ∼ ∕ ∼ ∕ the increase in P-wave velocity from 200 m s to 1200 m s. Figure 2. Predicted and field-based P- and S-wave velocity-versus- In unconsolidated sediments, the Hertz-Mindlin model accounts depth profiles and geotechnical cone penetration testing (CPT) soil for mechanical compaction from confining pressure, so that the pre- type profiles from the seismic survey area (Lorenzo et al., 2014)at dicted velocity is much more strongly dependent on effective stress (a) site A and (b) site B. The quality of the inversion results can be than porosity (Avseth et al., 1998). In almost fully saturated soil quantified by the misfit between the field and predicted velocity- versus-depth profiles. Misfits are calculated by the sum of the nor- (>99%), the velocity model predicts that a 1% change in water sat- malized root-mean-square error (NRMSE) between the predicted

Downloaded 10/17/16 to 130.39.188.137. Redistribution subject SEG license or copyright; see Terms of Use at http://library.seg.org/ uration leads to a 20% change in P-wave velocity, whereas a 1% and field velocity profiles. The CPT soil type profile shown here change in soil properties (e.g., porosity, coordination number, den- is 100 m wide and covers the same area as the seismic survey. sity, and elasticity) leads to less than a 1% change in P-wave veloc- At both sites, predicted velocities are calculated at discrete depths every 0.005 m. The depth error in the field velocity is 1m. The ity. To explain field velocity increases with depth (Figure 2), we also errors in field velocities are <2%. The soil types determined from need to relate the changes in water saturation and effective stress the combination of density, elastic moduli, and SWCC are labeled with depth. We use SWCC to predict the relationship between depth for each layer. WB14 Shen et al.

and water saturation (Appendix A). Effective stress relates to the ðua − uwÞ¼ρwghc; (1) confining pressure as well as matric suction, and both of the param- eters vary with depth. Density and elasticity (both matrix and effec- where ua is the atmospheric pressure, uw is the , tive) are calculated using water saturation and effective stress, hc is the capillary head, ρw is the water density, and g is the gravi- which are themselves depth dependent (Figure 1). tational acceleration (Table 1). Accurate prediction of seismic velocity in shallow (<25 m depth) In shallow unconsolidated sediments, effective stress (Peff ) is the unconsolidated sediments requires the incorporation of matric suc- sum of the net overburden pressure (σ − ua), matric suction tion and cohesion into the estimation of effective stress in the Hertz- (ua − uw), and cohesion (σco)(Lu and Likos, 2006): Mindlin and Biot-Gassmann models (Lu and Sabatier, 2009; Crane, ¼ σ − − ð − Þþσ 2013; Revil and Mahardika, 2013). In shallow, unconsolidated, Peff ua Sw ua uw co; (2) clay-rich soil, matric suction and cohesion can be several orders σ ρ ρ of magnitude larger than overburden pressure and dominate effec- where is the overburden pressure and equals eff gh ( eff is the tive stress (Lu and Sabatier, 2009). Rock physics models (Bachrach effective density of soil with pore fluids, and h is the depth of soil), and Nur, 1998) indicate that seismic velocity is proportional to one- ua is the atmospheric pressure and is assumed to be zero, and ð − Þ sixth of the power of the effective stress. The incorporation of ma- Sw ua uw is the matric suction contribution weighted by the per- tric suction and cohesion in clay-rich soil almost doubles the pre- centage of water saturation Sw (Lu and Sabatier, 2009). dicted seismic velocity and brings the results closer to real data (Crane, 2013). Cohesion is one kind of interparticle stress, and it Parameter constraint before inversion arises from physicochemical forces between mineral grains. The CMA-ES optimization is accelerated by constraining the uncer- matric suction is equal to the total stress difference at the air-water tainty of input parameter values within reasonable ranges (Table 2) − − interface (ua uw). In equilibrium, matric suction (ua uw) around for local soil types. Because our seismic survey was conducted on a capillary tube is balanced by the weight of the water column alluvial-deltaic soils (Lorenzo et al., 2014), the main soil types to pulled up by surface tension (Fredlund and Rahardjo, 1993b): consider are organic clay, clay, and sand. Organic clay has lower cohesion (Waltham, 2001) and larger compress- ibility than clay (Robertson, 1990) because of the Table 2. The 11 unknown parameters with published ranges used in the Hertz- Mindlin and Biot-Gassmann models (Appendix A). The ranges of each organic residue in clay. Input parameters in the parameter span those in organic clay, clay, and sand. These parameters are velocity prediction model include pore fluid assumed to be constant within each seismically recognizable layer. properties, soil properties, and water-saturation related parameters. To give more flexibility to the search for optimal values by CMA-ES, we Lower Upper Parameters boundary boundary References choose the largest published range of input parameter values for unconsolidated sediments 3 ρ0 (kg∕m ) 1.4 2.65 Walmsley (1977), Mavko et al. (2009) (Table 2). For the correction of the overestimated × 6 × 10 K0 (Pa) 3.4 10 3.66 10 Mittal et al. (2004), Mavko et al. (2009) shear modulus in the Hertz-Mindlin model, we 5 10 use a coordination number of one, which is be- G0 (Pa) 1.56 × 10 4.5 × 10 Mittal et al. (2004), Mavko et al. (2009) ϕ 0.35 0.8 Nimmo (2004) low a value with physical meaning. A co- ordination number of one is found to estimate C 18 Crane (2013), Allen (1985) velocity accurately in clay and sand mixtures σ × 4 co (Pa) 0 2 10 Fredlund (1991), Bishop et al. (1960) (Crane, 2013). Pore fluids are assumed to be θr 0 0.436 Leong and Rahardjo (1997) water and air, so their density and elastic proper- m 01 Van Genuchten (1980) ties are well known (Table 1). Because the maxi- a 01 Van Genuchten (1980) mum P-wave velocity throughout the profile is ∕ n 0 49.9 Van Genuchten (1980), about 1400 m s (Lorenzo et al., 2014) but below Leong and Rahardjo (1997) fully saturated velocity values, around 1500 m∕s (Grelle and Guadagno, 2009), for the purpose of SWCC calculation, the water table (value of 36 m) is assumed to be slightly below the depth Table 3. Common soil density and dynamic elastic moduli values for three of the profile. different soil types (organic clay, clay, and sand) used to calculate the average soil properties for each layer in a geotechnical soil model. Seismic soil property inversion Soil property inversion is carried out sepa- Density Bulk Shear rately within each seismically recognizable layer. Soil type (kg∕m3) modulus (Pa) modulus (Pa) References

Downloaded 10/17/16 to 130.39.188.137. Redistribution subject SEG license or copyright; see Terms of Use at http://library.seg.org/ Soil properties and the three fitting parameters in Organic clay 1.4 3.4 × 106 1.56 × 105 Walmsley (1977), SWCC (Table 2) are the outputs from inversion Mittal et al. (2004) when the misfit between the predicted and field Clay 2.6 2.1 × 1010 7 × 109 Mavko et al. (2009) velocity profiles reaches its minimum. From the Sand 2.65 3.66 × 1010 4.5 × 1010 Mavko et al. (2009) fitting parameters, we then compute inverted SWCCs with an assumed water table (value of Seismic soil property inversion WB15

36 m) (Figure 1). Three seismic layers can be determined from cases. The inverted density and elastic moduli correlate well in gen- sharp changes in the field-derived S-wave seismic velocity profiles eral with computed geotechnical results (Figure 3), except for the for both sites (Figure 2). We use CMA-ES optimization to minimize middle layer at site B (Figure 3b). the misfit between the field-velocity profile and the predicted veloc- The inverted soil density, elastic moduli profiles, and the SWCC ity profile by varying the input soil-property parameter values detect meaningful variations among three seismically recognizable within reasonable ranges (Table 2). The soil property parameters layers and between sites A and B (Figures 3 and 4) as expected. The used to explain the field velocity profile are the optimal values that inverted soil property values are much larger at site A than at site B represent the local soil characteristics. Based on the assumptions of because the soil is sandier at the depth of the middle seismic layer the velocity prediction model and considering the resolution of the (Figures 2 and 3). seismic data, we assume that the sediments are homogeneous and The inverted water saturation, bulk density, and porosity values at isotropic within each seismic layer. As a result, soil properties site B are also in agreement within an error of 15% with indepen- (Table 2) are constant within each seismic layer. Soil properties re- dent laboratory results from a well near site B (Figure 5). The largest present average values over the width of seismic survey area mismatch arises in the bulk density and porosity profiles in the (∼100 m) and over the depth of each layer. At each depth, water first layer. saturation is an average of patchy saturation over the seismic survey area (∼100 m). Our seismic data have effectively low frequency (≤30 Hz), so that pore fluid heterogeneity is unresolvable with the dominant seismic wavelet (Johnson, 2001). In each seismic layer, the misfit between the field velocity profile and the predicted velocity profile is quantified by the normalized root-mean-square error (NRMSE) at discrete depths every 0.005 m. NRMSE is cal- culated by normalizing the root-mean square of the difference be- tween the predicted and field-based velocities by the average of the field-based velocities. The 2% velocity error and 1‐m depth error from the original field velocity data propagate into errors in the inverted soil proper- ties and water saturation. For example, the depth error in the in- verted results is 1m, which carries over from the depth errors (1m) in the field-based velocity model. We estimate the final in- version errors via a Monte Carlo simulation. First, we randomly

generate 100 field-velocity-profile cases within their velocity error range of 2%. Then, we invert for soil properties with the 100 dif- ferent scenarios. The range of each resultant soil property value pro- vides the estimated error.

Determination of soil types from inverted results Inverted soil properties can be used to determine the soil types for each seismic layer by reference to published soil properties (Ta- ble 3). For different soil types, published density and elastic moduli vary over 70% (Table 3). Within single soil types, the variation in elastic moduli is about 20%–30% for clay, less than 5% for sand (Mavko et al., 2009), and about 50% for different organic clay (Mit- tal et al., 2004) because of different mineral content, overburden pressures, and organic content. We assign soil types based on where the inverted soil properties fall within these ranges. If the inverted soil property value lies outside the range for one soil type alone but between the values for two soil types, we consider that the soil type is a mixture of the two. Inverted soil property values that span the ranges of two or more soil types may indicate additional layering beyond the resolution of the seismic technique. SWCC can either be used to confirm a heterogeneity or locate thin units in a hetero- geneous layer that may not be sensed by inverted density and elastic moduli values. Figure 3. Density, bulk modulus, and shear modulus determined from seismic inversion and from geotechnical soil profile at (a) site

Downloaded 10/17/16 to 130.39.188.137. Redistribution subject SEG license or copyright; see Terms of Use at http://library.seg.org/ A and (b) site B. The quality of the inverted soil properties can be quantified by the misfit between the inverted soil properties and the RESULTS soil properties based on the CPT soil profile. For most layers, the inverted soil property values fall within the error ranges of geotech- The predicted velocity-versus-depth profile matches the field nical results. The largest misfit of 70% arises in the middle layer at velocity-versus-depth profile (Figure 2) in each of the three seismi- site B. Soil types determined from inverted density and elastic cally recognizable layers with NRMSEs of less than 0.15 in all moduli are labeled for each layer. WB16 Shen et al.

At sites A and B, the inverted soil properties show some common DISCUSSION trends that can be used to identify the soil types within the three different seismic layers. In the top and bottom seismic layers, in- Validation of inversion results and interpretations verted soil density and elastic moduli values fall within the range One measure of the usefulness of the inversion results for our for clay properties (Figure 3), but in addition, the relatively smaller three seismic layers is to compare them with other independent es- values of the top seismic layer may indicate the presence of organic timates of density and elastic moduli. If large differences appear material. We note that for both sites, the water saturation (based on between the two results, it may imply seismically unresolvable thin inverted SWCCs) shifts sharply to lower values across the top and units. We assign known values of density and elasticity in soil bottom boundaries of the middle seismic layers (Figure 4). Labo- (Table 3) to geotechnical soil-behavior types determined by cone ratory data confirm a similar change to lower water saturation values penetration testing (CPT) (Lorenzo et al., 2014) (Figure 2) to cal- at similar depths. For site A, the inverted soil density and elastic culate an average geotechnical density and elasticity for three new moduli values are relatively larger in the middle seismic layer than equivalent layers. The vertical resolution of the CPT soil profile is in the bottom layer. We interpret this to indicate that there may be 0.1 m. Computed geotechnical soil properties will have a range of additional sand within this clayey unit (Figure 3a, middle layer). values attributable to the variation of published soil properties (Ta- Soil types indicated by the inverted cohesion values are in agree- ble 3). A comparison of the inverted and computed geotechnical ment with soil types determined from the inverted density and elas- results of soil density and elastic moduli shows similar values ticity. Inverted cohesion values in the bottom seismic layers at both and implications for soil types (Figure 3). The greatest difference sites are close to 20 kPa and indicate that the soil type is possibly between the inverted and geotechnical results occurs at depths clay. Inverted cohesion values in the top seismic layers are about equivalent to the middle layer at site B (Figure 3b), where the in- 50% lower than in the bottom layers and may indicate the presence verted elastic moduli are almost 70% lower than those from the of less cohesive organics. In the middle seismic layer at site A, the CPT. One explanation for this big difference is the presence of thin inverted cohesion value is close to zero and indicates that the layer sand units (<1m) shown in the CPT soil profile (Figure 2b, the consists mainly of sand. In the middle seismic layer at site B, the middle layer). When seismic waves pass through these thin units inverted cohesion value is close to 20 kPa and indicates clayey soils. (<1mat site B), the changes in soil density and elastic moduli may not be seismically resolvable. For example, a P-wave velocity of ~200–1200 m/s and a dominant frequency of 30 Hz suggest a dominant wavelength of ~7–40 m. These thin sand units lead to a larger computed geotechnical soil property values than the in- verted results derived from seismic profiles. At site A, the sand layers are sufficiently thick. Thus, the inverted soil property values are closer to the geotechnical results and inverted soil types are comparable (Figures 2 and 3). These lateral differences in sand layer

Figure 5. Water saturation, bulk density, and porosity from seismic

Downloaded 10/17/16 to 130.39.188.137. Redistribution subject SEG license or copyright; see Terms of Use at http://library.seg.org/ Figure 4. SWCCs from seismic inversion at (a) site A and (b) site inversion at site B and laboratory tests from a well near site B. The B. The error in inverted SWCC is 0.1%. SWCC can help identify quality of the inversion can be quantified by differences between the heterogeneity within a layer. The shift of SWCC to a lower value inverted and laboratory results. Inverted water saturations are within within the middle layer indicates the presence of sand within the 2% of laboratory results. In the second and third layers, the inverted middle clay-dominant layer at both sites. Soil types determined bulk density and the density are within 5% of laboratory results. In from the combination of density, elastic moduli, and SWCC are la- the first layer, the inverted bulk density and density differ by 15% beled for each layer. from the laboratory results. Seismic soil property inversion WB17

thickness may also be responsible for the noticeable differences in An interesting possibility is that seismic velocity calculations over- density and elastic between the two sites (Figure 3). estimate the true saturation, which may be lowered by evaporation Inverted SWCCs can help confirm heterogeneity or recognize across the soil surface. In our velocity prediction model, water sat- thin units that cannot be sensed by the soil density and elastic uration is determined from the SWCC, which does not account for moduli alone. If the soil type changes, the at the same evaporation effects. As a result, the predicted water saturation from capillary head will also change (Fredlund and Xing, 1994). Because SWCC is greater than in the true field conditions, and the calculated the seismic velocity is most sensitive to the change in water satu- P-wave velocity is larger than the actual field velocity. ration in near-saturation conditions (Bachrach and Nur, 1998), so There is flexibility with the application of our suggested work- may the SWCC inverted from seismic velocity resolve changes flow in future studies. Herein, our attempt at soil property inversion in average water saturation at the same capillary head and help de- begins with a general velocity prediction model but without any termine soil type changes. In the middle seismic layers at both sites, empirical relationship to either simplify the inversion process or re- the shift in the SWCC to a lower value indicates the presence of duce possible errors such as those that can arise from using the sand units in the layer (Figure 4). If a clay layer contains sand units, wrong empirical relationship for a certain field setting. Ideally, the average water saturation will be lower than in a homogeneous the incorporation of other empirical relationships, such as porosity clay layer at the same capillary head. A combination of the results of versus coordination number and porosity versus depth, requires a inverted soil density, elastic moduli, and SWCC suggests that the good prior knowledge of the geologic setting, which often is not middle layers at both sites may contain sand (Figure 4). always the case. The influence of grain-size distribution could also Inverted water saturation, bulk density, and porosity appear con- be taken into account because in the inversion results, poor sorting sistent with results from laboratory tests from a well near site B leads to a decrease in porosity and an increase in the coordination (Figure 5). The laboratory water content and bulk density are di- number. In this study, we use geologic data from geotechnical and rectly measured on soil cores from the well, and porosity is calcu- laboratory tests to validate our inversion results. Thus, for the in- lated from these measurements (Lorenzo et al., 2014). Trends in the version, we only use field velocity profiles without the support inverted results at site B match laboratory results throughout the of extensive geologic data. In other known geologic settings, em- well. Most of the inverted soil property values are within the error pirical relationships could be incorporated into the velocity predic- range of laboratory results. Because velocity is more sensitive to the tion model to achieve a better inversion result. For example, the change in water saturation than other soil properties, inverted water unexpected increase of porosity with depth observed using labora- saturation has the smallest error (0.1%) compared with errors in tory results (Figure 5) could be incorporated into the future work to other inverted soil properties (which vary from 1% to 20%) improve inversion results. With the suggested workflow in this pa- (Figures 3–5). In the inverted and laboratory results, the water sat- per, other more complex velocity prediction models (such as those that include the effect of patchy saturation) could also be used to

uration shifts to lower values from the first layer to the second layer, and it shifts to higher values from the second to the third layer. As invert for additional soil-property and water-saturation information. previously mentioned, the shift in water saturation can help to detect Although geotechnical borehole tests (0.1 cm in our case) may seismically unresolvable thin layers. In the first organic clay layer, have higher vertical resolution than seismic-derived inverted results the inverted and laboratory-based bulk density and porosity show (>1m), seismic surveys advantageously provide continuous lateral the largest differences. One explanation of these differences is that seismic soil property values that may complement geotechnical the inverted soil properties represent average values over the seismic borehole tests for lower cost. Between geotechnical test sites, the survey area (which covers ∼100 m) and may not represent proper- inverted soil property results can highlight anomalies in the lateral ties at the specific well location. In organic clay, soil properties may changes of density, elasticity, and water saturation (Figures 3 and 4). have a larger lateral variation because of the difference in organic Based on the magnitude and the location of these anomalies, addi- content. tional geotechnical tests can be proposed and sited efficiently. An- other advantage is that soil property inversion uses seismic data, Comparison of inversion quality between different which sample in situ lateral variations of pressure, saturation, seismic layers and organic content, if any are present. The inversion results can reflect these lateral variations in soil properties between geotechni- Within each seismic layer, the corresponding soil behavior types cal boreholes. Currently, seismic soil property inversion for water (from CPT data) may vary and so they do not meet the homogeneity saturation is most sensitive if it is applied in near fully saturated assumption of the velocity prediction model. The quality of inver- conditions in which the field P-wave velocity can increase by more sion is best (with a small misfit of 2%) within the deepest layer than one order of magnitude with only a 1% change in the satura- (Figure 2). The difference between the inverted and geotechnical tion, and it is also most likely to detect lithologic changes. Inverted soil properties are also smallest within the deepest layers (Figure 3). soil stratigraphy of this type can improve with the improved reso- For the top and middle seismic layers, the soil behavior types are a lution of seismic velocity field profiles. mixture of clay and organic content or sand. Only for the bottom seismic layer is the soil type homogeneous clay. One of the assump- CONCLUSIONS tions of the Hertz-Mindlin and Biot-Gassmann theories is that the

Downloaded 10/17/16 to 130.39.188.137. Redistribution subject SEG license or copyright; see Terms of Use at http://library.seg.org/ porous medium is homogeneous (Wang, 2001). Thus, the inversion In shallow (<25 m) near-saturated soil (saturation >99%), we in- works best in the relatively homogeneous clay-rich layer in our vert for seismic soil properties by minimizing the misfit between three-layer seismic model. field-based velocity profiles and predicted velocity profiles based Inversion results near the top of the first layer, corresponding to on the Hertz-Mindlin and Biot-Gassmann theories. CMA-ES opti- the very near-surface soils (<5m), unexpectedly predict a velocity mizes the inversion results by searching for optimal input soil prop- that is higher than that seen in the field velocity profile (Figure 2). erty values that can best explain field velocity profiles. WB18 Shen et al.

The inverted density and elastic moduli can be used to interpret where Keff is the effective bulk modulus, Geff is the effective shear ρ major soil types and can detect variations in sand thickness between modulus, and eff is the effective density of the soil matrix with pore two field sites. By comparing inversion results with geotechnical fluids and expressed as (Dvorkin and Nur, 1996) results, the inverted soil properties appear valid in general except ρ ¼ ϕð ρ þð − Þρ Þþð − ϕÞρ for one layer, which probably contains seismically unresolvable eff Sw w 1 Sw a 1 0; (A-3) sandy units. ϕ ρ The inverted SWCC can help recognize thin sand units that are where is the porosity, SW is the water saturation, w is the density ρ ρ below the original seismic resolution of the field data. Laboratory of water, a is the density of air, and 0 is the density of the soil results validate the inversion results, as well as indicate that the re- grains. sults can be improved with a good prior empirical relationship be- The effective elastic moduli are calculated by the Gassmann fluid tween porosity and depth. The SWCC shifts to a lower value when substitution theory (Dvorkin and Nur, 1996): the thin unresolvable layers are sandier than the clay-dominated   K soil. In combination, the inverted density, elastic moduli, and Km þ fl K0 SWCC correspond to soil types that are in agreement with soil types K − K ϕðK − K Þ K ¼ 0 m 0 fl ; (A-4) derived from geotechnical data (CPT). eff K K 1 þ m þ fl For our three-layer seismic model at our two field sites, the in- K − K ϕðK − K Þ version works best in the relatively homogeneous clay-rich bottom 0 m 0 fl layer. Soil within this layer meets the assumption of homogeneity in the Hertz-Mindlin and Biot-Gassmann theories. ¼ Geff Gm; (A-5) Seismic surveys can provide continuous lateral seismic soil prop- erty values that may complement geotechnical borehole tests at where K0 is the bulk modulus of the soil grains, Km is the bulk lower cost. Inversion results can highlight anomalies in lateral modulus of the dry soil matrix, Gm is the shear modulus of the changes of density, elasticity, and water saturation to suggest addi- dry soil matrix, and Kfl is the bulk modulus of the pore fluids. tional geotechnical tests. We assume the pore fluid to be water and air, so that the bulk Although we use a general velocity model without any empirical modulus of the pore fluids is (Dvorkin and Nur, 1996) relationship in the current workflow, there is flexibility to apply our   suggested workflow in future studies. With a known geologic set- − −1 ¼ Sw þ 1 Sw ting, empirical relationships could be incorporated to improve the Kfl ; (A-6) Kw Ka inversion results. Other velocity prediction models could also be used for the inversion of additional information on soils. where Kw is the bulk modulus of water and Ka is the bulk modulus of air. ACKNOWLEDGMENTS Matrix elastic moduli are estimated using the Hertz-Mindlin con- tact theory (Dvorkin and Nur, 1996): This work is supported by a Shell Exploration and Production sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi – Company Grant (2011 2015). We thank H. Pham who helped us 2 2 2 3 C ð1 − ϕÞ G with CMA-ES and C. Wilson for valuable discussions. We thank ¼ 0 Km 2 2 Peff ; (A-7) the following for their support with scholarships and graduate 18π ð1 − νÞ assistantship to the first author: Geometrics, Inc., a student travel sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi grant for the Symposium on the Application of Geophysics to En- 2 2 2 5 − 4ν 3 3C ð1 − ϕÞ G gineering and Environmental Problems; the Donald Towse Memo- G ¼ 0 P ; (A-8) m ð − νÞ π2ð − νÞ2 eff rial Fund awarded by an AAPG Grant-in-Aid; and the EnCana 5 2 2 1 Scholarship and Imagine Resources Scholarship awarded by the Louisiana State University (LSU) Department of Geology and Geo- where C is the coordination number, G0 is the shear modulus of soil ν ’ physics. Special thanks go to the LSU Department of Geology and grains, is the Poisson s ratio of the soil grains, and Peff is the ef- Geophysics for their active support of graduate student research. fective stress. The SWCC fitting function determines the relationship between the effective water saturation Se and the capillary head h (Van Gen- APPENDIX A uchten, 1980):   VELOCITY PREDICTION MODEL m ¼ 1 Se n ; (A-9) The seismic P-wave velocity (VP) and S-wave velocity (VS) are 1 þ½ah calculated from the effective elastic moduli and bulk density (Dvor- kin and Nur, 1996): sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi where a, n, and m are fitting parameters. Water saturation is related to effective water saturation as (Van þ 4 Keff Geff Downloaded 10/17/16 to 130.39.188.137. Redistribution subject SEG license or copyright; see Terms of Use at http://library.seg.org/ ¼ 3 Genuchten, 1980) VP ρ ; (A-1) eff ðϕ − θ Þþθ sffiffiffiffiffiffiffiffi ¼ Se r r Sw ϕ ; (A-10) ¼ Geff VS ρ ; (A-2) eff where θr is the residual volumetric water content. Seismic soil property inversion WB19

Hansen, N., 2011, The CMA evolution strategy: A tutorial, http://www.lri.fr/ REFERENCES ~hansen/cmatutorial.pdf, accessed 6 February 2015. Harris, J. B., 2009, Hammer-impact SH-wave seismic reflection methods in Allen, J. R. L., 1985, Principles of physical sedimentology: Allen and neotectonic investigations: General observations and case histories from Unwin. the Mississippi Embayment, USA: Journal of Earth Science, 20, 513–525, Avseth, P., J. Dvorkin, G. Mavko, and J. Rykkje, 1998, Diagnosing doi: 10.1007/s12583-009-0043-y. high-porosity sands for reservoir characterization using sonic and Hayashi, K., T. Inazaki, K. Kitao, and T. Kita, 2013, Statistical soil type seismic: 68th Annual International Meeting, SEG, Expanded Abstracts, estimation using cross-plots of S-wave velocity and resistivity in Japanese 1024–1027. levees: Presented at Symposium on the Application of Geophysics to En- Avseth, P., T. Mukerji, and G. Mavko, 2005, Quantitative seismic interpre- gineering and Environmental Problems. tation: Applying rock physics tools to reduce interpretation risk: Cam- Hertz, H., 1882, On the contact of rigid elastic and on hardness, in D. bridge University Press. E. Jones, and G. A. Schott, eds., Miscellaneous papers by H. Hertz: Mac- Bachrach, R., and P. Avseth, 2008, Rock physics modeling of unconsoli- Millan, 163–183. dated sands: Accounting for nonuniform contacts and heterogeneous Ikelle, L. T., and L. Amundsen, 2005, Introduction to petroleum seismology: stress fields in the effective media approximation with applications to hy- SEG. drocarbon exploration: Geophysics, 73, no. 6, E197–E209, doi: 10.1190/1 Johnson, D. L., 2001, Theory of frequency dependent acoustics in patchy- .2985821. saturated porous media: Journal of the Acoustical Society of America, Bachrach, R., J. Dvorkin, and A. Nur, 1998, High-resolution shallow-seis- 110, 682–694, doi: 10.1121/1.1381021. mic experiments in sand, Part II: Velocities in shallow unconsolidated Leong, E. C., and H. Rahardjo, 1997, Review of soil-water characteristic sand: Geophysics, 63, 1234–1240, doi: 10.1190/1.1444424. curve equations: Journal of Geotechnical and Geoenvironmental Engi- Bachrach, R., J. Dvorkin, and A. M. Nur, 2000, Seismic velocities and Pois- neering, 123, 1106–1117, doi: 10.1061/(ASCE)1090-0241(1997)123: son’s ratio of shallow unconsolidated sands: Geophysics, 65, 559–564, 12(1106). doi: 10.1190/1.1444751. Lorenzo, J. M., J. Hicks, and E. E. Vera, 2014, Integrated seismic and cone Bachrach, R., and A. Nur, 1998, High-resolution shallow-seismic experi- penetration test observations at a distressed earthen levee: Marrero, Loui- ments in sand, Part I: Water table, fluid flow, and saturation: Geophysics, siana, USA: Engineering Geology, 168,59–68, doi: 10.1016/j.enggeo 63, 1225–1233, doi: 10.1190/1.1444423. .2013.10.019. Bayer, P., E. Duran, R. Baumann, and M. Finkel, 2009, Optimized ground- Lu, N., and W. J. Likos, 2006, Suction stress characteristic curve for unsatu- water drawdown in a subsiding urban mining area: Journal of , rated soil: Journal of Geotechnical and Geoenvironmental Engineering, 365,95–104, doi: 10.1016/j.jhydrol.2008.11.028. 132, 131–142, doi: 10.1061/(ASCE)1090-0241(2006)132:2(131). Bayer, P., and M. Finkel, 2007, Optimization of concentration control Lu, Z., and J. M. Sabatier, 2009, Effects of soil water potential and moisture by evolution strategies: Formulation, application, and assessment of content on sound speed: Society of America Journal, 73, remedial solutions: Water Resources Research, 43, W02410, doi: 10 1614–1625, doi: 10.2136/sssaj2008.0073. .1029/2005WR004753. Masoni, I., W. Zhou, R. Brossier, L. Métivier, S. Operto, and J. Virieux, Bell, F. G., 1992, Engineering properties of soils and rocks: Butterworth- 2014, Near-surface full waveform inversion using surface waves and re- Heinemann Ltd. flected waves: 76th Annual International Conference and Exhibition, Biot, M. A., 1962, Mechanics of deformation and acoustic propagation in EAGE, Extended Abstracts, doi: 10.3997/2214-4609.20140531. porous media: Journal of Applied Physics, 33, 1482–1498, doi: 10.1063/1 Mavko, G., T. Mukerji, and J. Dvorkin, 2009, The rock physics handbook: .1728759. Tools for seismic analysis of porous media: Cambridge University Press. Bishop, A. W., I. Alpan, G. E. Blight, and I. B. Donald, 1960, Factors Mezura-Montes, E., and C. A. Coello Coello, 2011, Constraint-handling in controlling the strength of partly saturated cohesive soils: Presented at nature-inspired numerical optimization: Past, present and future: Swarm the Research Conference on Shear Strength of Cohesive Soils. and Evolutionary Computation, 1, 173–194, doi: 10.1016/j.swevo.2011 Crane, J. M., 2013, Effects of stress and water saturation on seismic velocity .10.001. and attenuation in near surface sediments: Ph.D. dissertation, Louisiana Mindlin, R. D., 1949, Compliance of elastic bodies in contact: Journal of State University. Applied Mechanics, 16, 259–268. Digby, P. J., 1981, The effective elastic moduli of porous granular rocks: Mittal, B., H. Yi, V. M. Puri, A. S. McNitt, and C. F. Mancino, 2004, Bulk Journal of Applied Mechanics, 48, 803–808, doi: 10.1115/1.3157738. mechanical behavior of rootzone sand mixtures as influenced by Dutta, T., G. Mavko, and T. Mukerji, 2010, Improved granular medium shape, moisture and : Particle and Particle Systems Characterization, model for unconsolidated sands using coordination number, porosity, 21, 303–309, doi: 10.1002/ppsc.200400934. and pressure relations: Geophysics, 75, no. 2, E91–E99, doi: 10.1190/ Murphy, W. F., 1982, Effects of microstructure and pore fluids on the acous- 1.3333539. tic properties of granular sedimentary materials: Ph.D. dissertation, Stan- Dvorkin, J., and A. Nur, 1996, Elasticity of high-porosity sandstones: ford University. Theory for two North Sea data sets: Geophysics, 61, 1363–1370, doi: Nimmo, J. R., 2004, Porosity and pore size distribution: in D. Hillel, ed., 10.1190/1.1444059. Encyclopedia of Soils in the Environment, vol. 3: Elsevier, 295–303. Fredlund, D. G., 1991, How negative can pore-water pressures get?: Geo- Revil, A., and H. Mahardika, 2013, Coupled hydromechanical and electro- technical News, 9,44–46. magnetic disturbances in unsaturated porous materials: Water Resources Fredlund, D. G., 1995, The scope of unsaturated : An over- Research, 49, 744–766, doi: 10.1002/wrcr.20092. view, in E. E. Alonso , and P. Delage, eds., Proceedings of the First Robertson, P. K., 1990, using the : International Conference on Unsaturated Soils, vol. 3: Balkema, 1155– Canadian Geotechnical Journal, 27, 151–158, doi: 10.1139/t90-014. 1177. Takahashi, T., and S. Tanaka, 2009, Rock physics model for interpreting Fredlund, D. G., and H. Rahardjo, 1993a, An overview of unsaturated soil elastic properties of soft sedimentary rocks: International Journal of behaviour: in S. L. Houston, and W. K. Wray, eds., Proceedings of ses- the Japanese Committee for Rock Mechanics, 4,53–59. sions of ASCE Convention: American Society of Civil Engineers, Spe- Terzaghi, K., 1996, Soil mechanics in engineering practice: John Wiley & cialty Series on Unsaturated Soil Properties, 1–31. Sons. Fredlund, D. G., and H. Rahardjo, 1993b, Soil mechanics for unsaturated Van Genuchten, M. T., 1980, A closed-form equation for predicting the hy- soils: John Wiley & Sons. draulic conductivity of unsaturated soils: Soil Science Society of America Fredlund, D. G., and A. Xing, 1994, Equations for the soil-water character- Journal, 44, 892–898, doi: 10.2136/sssaj1980.03615995004400050002x. istic curve: Canadian Geotechnical Journal, 31, 521–532, doi: 10.1139/ Velea, D., F. D. Shields, and J. M. Sabatier, 2000, Elastic wave velocities in t94-061. partially saturated Ottawa sand: Soil Science Society of America Journal, Fredlund, D. G., A. Xing, M. D. Fredlund, and S. L. Barbour, 1996, The 64, 1226–1234, doi: 10.2136/sssaj2000.6441226x. relationship of the unsaturated soil shear to the soil-water characteristic Walmsley, M. E., 1977, Physical and chemical properties of peat: Muskeg curve: Canadian Geotechnical Journal, 33, 440–448, doi: 10.1139/t96- and the Northern Environment in Canada: University of Toronto Press, 065. 83–129. Gassmann, F., 1951, Elastic waves through a packing of spheres: Geophys- Waltham, T., 2001, Foundations of engineering geology: CRC Press. ics, 16, 673–685, doi: 10.1190/1.1437718. Walton, K., 1987, The effective elastic moduli of a random packing of Gilman, K., 1994, Hydrology and wetland conservation: John Wiley & spheres: Journal of the Mechanics and Physics of Solids, 35, 213–226, Downloaded 10/17/16 to 130.39.188.137. Redistribution subject SEG license or copyright; see Terms of Use at http://library.seg.org/ Sons. doi: 10.1016/0022-5096(87)90036-6. Grelle, G., and F. M. Guadagno, 2009, Seismic refraction methodology for Wang, Z., 2001, Fundamentals of seismic rock physics: Geophysics, 66, groundwater level determination: “Water seismic index”: Journal of 398–412, doi: 10.1190/1.1444931. Applied Geophysics, 68, 301–320, doi: 10.1016/j.jappgeo.2009.02.001. Zimmer, M. A., M. Prasad, G. Mavko, and A. Nur, 2007, Seismic velocities Hansen, N., 2006, The CMA evolution strategy: A comparing review, in P. of unconsolidated sands: Part 1 — Pressure trends from 0.1 to 20 MPa: Larrañaga, J. A. Lozano, I. Inza, and E. Bengoetxea, eds., Towards a new Geophysics, 72, no. 1, E1–E13, doi: 10.1190/1.2399459. evolutionary computation: Springer, 75–102.