soil dynamics and earthquake engineering64 a. fathi et al. / soil dynamics and earthquake...

19
Three-dimensional P- and S-wave velocity proling of geotechnical sites using full-waveform inversion driven by eld data Arash Fathi a , Babak Poursartip b , Kenneth H. Stokoe II b , Loukas F. Kallivokas a,b,n a Institute for Computational Engineering and Sciences, The University of Texas at Austin, Austin, TX, USA b Department of Civil, Architectural and Environmental Engineering, The University of Texas at Austin, Austin, TX, USA article info Article history: Received 22 December 2015 Received in revised form 18 April 2016 Accepted 19 April 2016 Available online 7 May 2016 Keywords: Site characterization Full-waveform inversion (FWI) Elastic waves perfectly-matched-layers (PMLs) Spectral-Analysis-of-Surface-Waves (SASW) Vibroseis Field data processing abstract We discuss the application of a recently developed full-waveform-inversion-based technique to the imaging of geotechnical sites using eld-collected data. Specically, we address the proling of arbitrarily heterogeneous sites in terms of P- and S-wave velocities in three dimensions, using elastic waves as probing agents. We cast the problem of nding the spacial distribution of the elastic soil properties as an inverse medium problem, directly in the time domain, and use perfectly-matched-layers (PMLs) to ac- count for the semi-innite extent of the site under investigation. After briey reviewing the theoretical and computational aspects of the employed technique, we focus on the characterization of the George E. Brown Jr. Network for Earthquake Engineering Simulation site in Garner Valley, California (NEES@UCSB). We compare the proles obtained from our full-wave- form-inversion-based methodology against the proling obtained from the Spectral-Analysis-of-Surface- Waves (SASW) method, and report agreement. In an attempt to validate our methodology, we also compare the recorded eld data at select control sensors that were not used for the full-waveform in- version, against the response at the same sensors, computed based on the full-waveform-inverted proles. We report very good agreement at the control sensors, which is a strong indicator of the cor- rectness of the inverted proles. Overall, the systematic framework discussed herein seems robust, general, practical, and promising for three-dimensional site characterization purposes. & 2016 Elsevier Ltd. All rights reserved. 1. Introduction Reliable three-dimensional P- and S-wave velocity proles of the near-surface soil have great value in geotechnical engineering practice, and can play a vital role in the safe design of critical components of the civil infrastructure, such as nuclear power plants, bridges, and hospitals, among others. Wave velocity pro- ling of the soil is referred to as seismic site characterization. Seismic site characterization techniques can be broadly classied into two categories: invasive and non-invasive. Invasive techni- ques, such as borehole-based methods, are costly and can typically provide only localized information of the site under investigation, while, non-invasive techniques are capable of providing global imaging, and are generally more cost-efcient. Non-invasive proling of the soil shares common elements with medical ima- ging [17]. In both cases, elastic waves are used to interrogate the unseeable medium, and the medium's response to this probing is subsequently used for driving the imaging. In the mathematical sciences, this process is referred to as solving an inverse medium problem. Our goal is to nd the distribution of the P-wave velocity ( ) c x p , and S-wave velocity ( ) c x s of the soil in three-space dimensions. 1 We use a seismic vibrator to apply loads on the ground surface, which results in the propagation of elastic waves in the soil. Due to the possibly heterogeneous character of the site, multiple reec- tions and refractions occur. The time-history response of the soil medium to this loading is recorded by using geophones, spatially arranged over the formation's surface, and is, hereafter, referred to as measured response (Fig. 1). Arriving at a material prole can then be achieved by minimizing the difference between the measured response at sensor locations, and a computed response corre- sponding to a trial distribution of the material parameters, under Contents lists available at ScienceDirect journal homepage: www.elsevier.com/locate/soildyn Soil Dynamics and Earthquake Engineering http://dx.doi.org/10.1016/j.soildyn.2016.04.010 0267-7261/& 2016 Elsevier Ltd. All rights reserved. n Corresponding author at: Department of Civil, Architectural and Environmental Engineering, The University of Texas at Austin, Austin, TX, USA. E-mail addresses: [email protected] (A. Fathi), [email protected] (B. Poursartip), [email protected] (K.H. Stokoe II), [email protected] (L.F. Kallivokas). 1 =( ) xyz x , , is the position vector. Soil Dynamics and Earthquake Engineering 87 (2016) 6381

Upload: others

Post on 08-Apr-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Soil Dynamics and Earthquake Engineering64 A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81. mathematical framework for tackling the inverse medium pro-blem

Soil Dynamics and Earthquake Engineering 87 (2016) 63–81

Contents lists available at ScienceDirect

Soil Dynamics and Earthquake Engineering

http://d0267-72

n CorrEnginee

E-mbabakp@loukas@

journal homepage: www.elsevier.com/locate/soildyn

Three-dimensional P- and S-wave velocity profiling of geotechnicalsites using full-waveform inversion driven by field data

Arash Fathi a, Babak Poursartip b, Kenneth H. Stokoe II b, Loukas F. Kallivokas a,b,n

a Institute for Computational Engineering and Sciences, The University of Texas at Austin, Austin, TX, USAb Department of Civil, Architectural and Environmental Engineering, The University of Texas at Austin, Austin, TX, USA

a r t i c l e i n f o

Article history:Received 22 December 2015Received in revised form18 April 2016Accepted 19 April 2016Available online 7 May 2016

Keywords:Site characterizationFull-waveform inversion (FWI)Elastic wavesperfectly-matched-layers (PMLs)Spectral-Analysis-of-Surface-Waves (SASW)VibroseisField data processing

x.doi.org/10.1016/j.soildyn.2016.04.01061/& 2016 Elsevier Ltd. All rights reserved.

esponding author at: Department of Civil, Arcring, The University of Texas at Austin, Austinail addresses: [email protected] (A. Fathiutexas.edu (B. Poursartip), [email protected] (L.F. Kallivokas).

a b s t r a c t

We discuss the application of a recently developed full-waveform-inversion-based technique to theimaging of geotechnical sites using field-collected data. Specifically, we address the profiling of arbitrarilyheterogeneous sites in terms of P- and S-wave velocities in three dimensions, using elastic waves asprobing agents. We cast the problem of finding the spacial distribution of the elastic soil properties as aninverse medium problem, directly in the time domain, and use perfectly-matched-layers (PMLs) to ac-count for the semi-infinite extent of the site under investigation.

After briefly reviewing the theoretical and computational aspects of the employed technique, wefocus on the characterization of the George E. Brown Jr. Network for Earthquake Engineering Simulationsite in Garner Valley, California (NEES@UCSB). We compare the profiles obtained from our full-wave-form-inversion-based methodology against the profiling obtained from the Spectral-Analysis-of-Surface-Waves (SASW) method, and report agreement. In an attempt to validate our methodology, we alsocompare the recorded field data at select control sensors that were not used for the full-waveform in-version, against the response at the same sensors, computed based on the full-waveform-invertedprofiles. We report very good agreement at the control sensors, which is a strong indicator of the cor-rectness of the inverted profiles. Overall, the systematic framework discussed herein seems robust,general, practical, and promising for three-dimensional site characterization purposes.

& 2016 Elsevier Ltd. All rights reserved.

1. Introduction

Reliable three-dimensional P- and S-wave velocity profiles ofthe near-surface soil have great value in geotechnical engineeringpractice, and can play a vital role in the safe design of criticalcomponents of the civil infrastructure, such as nuclear powerplants, bridges, and hospitals, among others. Wave velocity pro-filing of the soil is referred to as seismic site characterization.Seismic site characterization techniques can be broadly classifiedinto two categories: invasive and non-invasive. Invasive techni-ques, such as borehole-based methods, are costly and can typicallyprovide only localized information of the site under investigation,while, non-invasive techniques are capable of providing globalimaging, and are generally more cost-efficient. Non-invasive

hitectural and Environmental, TX, USA.),xas.edu (K.H. Stokoe II),

profiling of the soil shares common elements with medical ima-ging [17]. In both cases, elastic waves are used to interrogate theunseeable medium, and the medium's response to this probing issubsequently used for driving the imaging. In the mathematicalsciences, this process is referred to as solving an inverse mediumproblem.

Our goal is to find the distribution of the P-wave velocity ( )c xp ,and S-wave velocity ( )c xs of the soil in three-space dimensions.1

We use a seismic vibrator to apply loads on the ground surface,which results in the propagation of elastic waves in the soil. Due tothe possibly heterogeneous character of the site, multiple reflec-tions and refractions occur. The time-history response of the soilmedium to this loading is recorded by using geophones, spatiallyarranged over the formation's surface, and is, hereafter, referred toasmeasured response (Fig. 1). Arriving at a material profile can thenbe achieved by minimizing the difference between the measuredresponse at sensor locations, and a computed response corre-sponding to a trial distribution of the material parameters, under

1 = ( )x y zx , , is the position vector.

Page 2: Soil Dynamics and Earthquake Engineering64 A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81. mathematical framework for tackling the inverse medium pro-blem

Fig. 1. Problem definition: (a) interrogation of a heterogeneous semi-infinite domain by an active source; and (b) computational model truncated from the semi-infinitemedium via the introduction of perfectly-matched-layers (PMLs).

A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–8164

the same loading, using a computational model. The computa-tional model should mimic the physics of the problem as closely aspossible. A key challenge is proper termination of the semi-infiniteextent of the domain of interest.2 We use perfectly matched-layers(PMLs) for domain truncation (Fig. 1) [5,15,39], since they areamong the best available tools in the presence of heterogeneity[15]. A PML is a (practically reflectionless) buffer zone that sur-rounds the domain of interest, where outgoing waves are forced todecay.

Obtaining the P- and S-wave velocity profiles of the soil inthree-dimensional, arbitrarily heterogeneous domains, remains achallenging problem. The inverse medium problem at hand isnotoriously ill-posed, computationally expensive, posed over asemi-infinite medium, and has challenging physics. Due to thesecomplexities, and in order to render the problem tractable, currenttechniques rely on simplifying assumptions. We classify thesesimplifications as follows: (a) limiting the spatial variability of thesoil properties, whereby it is assumed that the soil is horizontallylayered (one-dimensional) [25,29,33], or has properties varyingonly within a plane (two-dimensional) [3,12,19,21]; (b) assumingthat the measured response of the soil at sensor locations is dueonly to Rayleigh waves, thus neglecting other wave types, such ascompressional and shear waves, as is the case in the SASW [33], orits close variant, the MASW method [30]; (c) idealizing the soilmedium, which is porous, and, generally, partially or fully satu-rated, as an elastic solid; (d) imaging only one elastic property,such as the shear wave velocity or an equivalent counterpart[2,11,27,28]; and (e) grossly simplifying the boundary conditionsassociated with the semi-infinite extent of the medium, due to thecomplexity and computational cost that a rigorous treatmentwould require [11,36]. In recent years, the ubiquity of parallelcomputers, and significant advances in computational geosciences,has created the opportunity of developing a toolkit that is capableof robust, accurate, and three-dimensional characterization ofgeotechnical sites.

In this paper, we extend recent advances in site characterizationusing full-waveform inversion (FWI) to the all-important case ofthree-dimensional site profiling using actual field data. In recentwork [19], we described a methodology for accommodating field-

2 The domain of interest is the region one aims to characterize.

data, which are inherently three-dimensional, into two-dimensionalfull-waveform-inversion-based software, whenever the site condi-tions would support plane strain assumptions. We demonstrated themethod's viability by characterizing a site in Austin, Texas [19]. In thiscommunication, we abandon the plane strain assumption, and con-sider arbitrary heterogeneity in three space dimensions. To be able toextend the framework described in [19] to the fully three-dimen-sional case, various algorithmic modifications are needed, includingmodifications to the forward three-dimensional wave simulator, per[15], to the inversion methodology, per [14], and to the manner thefield experiment is designed and the associated field data are pro-cessed post-recording. We use the mathematical apparatus that wedescribed recently in [13-15], and excerpt herein, but focus on thethree-dimensional characterization of the NEES@UCSB site in GarnerValley, California, as a case study. To this end, we integrate recentadvances in several areas. Specifically, we use: (a) a parallel, three-dimensional, state-of-the-art wave simulation tool for domains ter-minated by PMLs [15]; (b) a partial-differential-equation (PDE)-con-strained optimization framework, through which the misfit mini-mization between the measured response and the computed responseis enforced [14]; (c) regularization schemes to alleviate the ill-po-sedness and solution multiplicity inherent in inverse problems,where the regularization factor is computed adaptively at each in-version iteration [14,20]; (d) continuation schemes that provide al-gorithmic robustness [20]; and (e) a biasing scheme for the robustsimultaneous inversion of both ( )c xp and ( )c xs profiles [14,21].

Most inversion approaches result in profile reconstructions thatare physically acceptable. Validation of the inverted profiles, in theabsence of borehole data, becomes challenging, if not impossible.Herein, to aid in the validation of the inverted profiles, we use a setof control sensors. Specifically, out of the 35 sensors that we de-ployed at the NEES@UCSB site, we use only 30 of them to feed ourfull-waveform-inversion algorithms. The remaining 5 sensorsserve as controls. We compute a time-history response corre-sponding to the full-waveform-inverted soil profiles, and comparethem against the actual recorded field response at the location ofthe 5 control sensors. We report remarkable agreement at thecontrol sensors.

The remainder of this paper is organized as follows: first, wereview the simulation of elastic waves in a three-dimensionalPML-truncated medium, followed by discussing a systematic

Page 3: Soil Dynamics and Earthquake Engineering64 A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81. mathematical framework for tackling the inverse medium pro-blem

Fig. 2. Wave attenuation via perfectly-matched-layers (PMLs): (a) implementation at the computational domain level; (b) conceptual depiction of amplitude decay withinthe PML.

Fig. 3. Equipment used for the field experiments at the NEES@UCSB site.

A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81 65

mathematical framework for tackling the inverse medium pro-blem. Next, we report on the design of the field experiment, datacollection, data processing, and characterization of the NEE-S@UCSB site in Garner Valley, California, using full-waveform in-version. We then compare the resulting imaging against the pro-filing obtained from the SASW method. We also compute thetime-history response of the site corresponding to the invertedprofiles, and compare them against field measurements. Lastly, weconclude with summary remarks.

2. Theoretical aspects – a review

The mathematical development of our full-waveform-inver-sion-based methodology has been discussed in [14,15]. Here, forcompleteness, we excerpt key parts: there are two importantcomponents, namely, the forward wave simulator, and the inversemedium problem optimizer.

2.1. The forward problem

We review first the numerical simulation of elastic wave pro-pagation in a three-dimensional, semi-infinite, arbitrarily hetero-geneous elastic medium, referred to as the forward problem. Dueto the semi-infinite extent of the medium, domain truncation isnecessary to arrive at a finite computational model. Limiting theunboundedness can be realized by placing PMLs at the truncationboundaries (Fig. 2(a)), such that, ideally, when outgoing wavestravel through the reflectionless interface, they get attenuatedwithin the PML buffer zone, as illustrated in Fig. 2(b).

The forward problem needs to be solved repeatedly during theinversion process. Therefore, having a computationally efficientforward solver can influence the total computational costsignificantly. To this end, we use a recently developed, state-of-the-art hybrid approach, which uses a displacement–stress for-mulation for the PML buffer, coupled with a standard displace-ment-only formulation for the interior (regular) domain. We refer

Page 4: Soil Dynamics and Earthquake Engineering64 A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81. mathematical framework for tackling the inverse medium pro-blem

A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–8166

to [15] and references therein for the complete development andparallel implementation of the forward wave solver. Herein, weonly repeat the resulting coupled system of equations.

Accordingly, find ( )tu x, in Ω Ω∪RD PML , and ( )tS x, , in ΩPML (seeFig. 2(a) for domain and boundary designations), where u and S,reside in appropriate function spaces and:

μ λ ρ Ω{ [∇ + (∇ ) ] + ( ) } = ¨ × ( )div u u u udiv in J, 1aT RD

Λ Λ Λ ρ

Ω

( + + ¯ ) = ( ¨ + + + ¯ )

× ( )

a b cdiv S S S u u u ud

in J, 1b

Te

Tp

Tw

PML

μ Λ Λ Λ Λ Λ

Λ λ Λ Λ Λ

Ω

¨ + + + ¯

= [(∇ ) + (∇ ) + (∇ ) + (∇ ) + (∇ ¯ )

+ (∇ ¯ ) ] + [ ( ) + ( ) + ( ¯ )]

× ( )

a b cS S S Su u u u u

u u u u

d

div div div

in J. 1c

e eT

p pT

w

wT

e p w

PML

The system is initially at rest, and subject to the followingboundary and interface conditions:

Fig. 4. Field experimental layout.

Fig. 5. Theoretical time history of the chir

μ λ Γ{ [∇ + (∇ ) ] + ( ) } = × ( )gu u u ndiv on J, 2aT

n NRD

Λ Λ Λ Γ( + + ¯ ) = × ( )S S S n 0 on J, 2bT

eT

pT

w NPML

Fig. 7. Time-history of the source load shown in Fig. 6.

p load and its Fourier spectrum.

Fig. 6. A source located at ( ) = ( )x y, 20, 35 m, and four equidistant sensors.

Page 5: Soil Dynamics and Earthquake Engineering64 A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81. mathematical framework for tackling the inverse medium pro-blem

Fig. 8. Vertical displacement time-histories at the four sensors shown in Fig. 6.

Fig. 9. Sensors used for full-waveform inversion (left), and validation (right), respectively. Control sensors are marked sp1 through sp5.

Table 1Parameters used during each inversion stage.

Stage Temporalduration

Normalized regularization fac-tor parameter ϱ [14]

Cumulativeiterations

1 0.5 0.60 102 1.0 0.55 1803 1.5 0.50 3504 2.0 0.45 4305 2.5 0.40 4606 3.0 0.35 500

A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81 67

Γ= × ( )u 0 on J, 2cDPML

Γ= × ( )u u on J, 2dRD PML I

μ λ

Λ Λ Λ Γ

{ [∇ + (∇ ) ] + ( ) }

= ( + + ¯ ) × ( )

u u u

n S S S n

div

on J, 2e

T

Te

Tp

Tw

I

Page 6: Soil Dynamics and Earthquake Engineering64 A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81. mathematical framework for tackling the inverse medium pro-blem

Fig. 10. P-wave velocity (in m/s) profile of the site obtained via full-waveform in-version. The vertical section cuts through the domain from ( ) = ( )x y, 30, 60 (right),to ( )30, 30 , to ( )60, 30 (left).

Fig. 11. S-wave velocity (in m/s) profile of the site obtained via full-waveform in-version. The vertical section cuts through the domain from ( ) = ( )x y, 30, 60 (right),to ( )30, 30 , to ( )60, 30 (left).

A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–8168

where temporal and spatial dependencies are suppressed forbrevity; u is the displacement vector, ρ is the mass density of themedium, λ and μ are the Lamé parameters, is the second-orderidentity tensor, S represents the stress tensor, a dot ( ) denotesdifferentiation with respect to time, and a bar (¯) indicates historyof the subtended variable3; ΩRD represents the interior (regular)domain, ΩPML denotes the region occupied by the PML buffer zone,ΓI is the interface boundary between the interior and PML do-mains, ΓN

RD and ΓNPML denote the free (top surface) boundary of the

interior domain and PML, respectively, = ( ]TJ 0, is the time in-terval of interest, and gn is the prescribed surface traction.Moreover, Λe, Λp, and Λw are the so-called stretch tensors, whichenforce dissipation of waves in ΩPML , and a, b, c, and d are productsof certain elements of the stretch tensors [15]. Eq. (1a) is thegoverning PDE for the interior elastodynamic problem, whereasEqs. (1b) and (1c) are the equilibrium and combined kinematicand constitutive equations, respectively, for the PML. Eqs. (2a) and(2b) describe the surface traction boundary condition for the

3 For instance, ∫ τ τ¯ ( ) = ( )tu x u x, , dt

0.

interior domain and PML, respectively, and Eq. (2e) enforces thebalance of tractions at the interface between the interior domainand PML. We remark that at the interface ΓI, Λ =e , andΛ Λ= = 0p w .

We use a standard Galerkin finite element approach for thespatial discretization of the interior elastodynamic domain, wherethe unknowns are nodal displacements, whereas within the PMLdomain, displacements and stress components (i.e., u S, ) are bothtreated as unknowns. It can be shown [15] that the following first-order system of ordinary differential equations results:

⎣⎢⎢⎢

⎦⎥⎥⎥

⎣⎢⎢⎢

⎦⎥⎥⎥

⎣⎢⎢⎢

⎦⎥⎥⎥

⎣⎢⎢⎢

⎦⎥⎥⎥

=− − −

+( )

t

xx

Mx

0 I 00 0 IG K C

xxx

00

f

dd

,

3

0

1

2

0

1

2 st

where M C K G, , , , are system matrices, = ¯x d0st, =x d1

st, = x d2st,

= ( )d u S, ,hT

hT Tst is the vector of nodal unknowns comprising nodal

displacements uh, in Ω Ω¯ ∪ ¯RD PML, and nodal stress components Sh,only in ΩPML , and f st is the vector of applied forces. We use spectralelements with a Legendre–Gauss–Lobatto (LGL) quadrature rule torender M diagonal, and use an explicit fourth-order Runge–Kutta(RK-4) method for integrating (3) in time.

2.2. The inverse medium problem

Our goal is to find the spatial distribution of the P- and S-wavevelocity profile of three-dimensional, arbitrarily heterogeneousformations. Due to the expression of the material properties interms of the Lamé parameters in Eqs. (1) and (2), the mathematicalframework relies on inverting for λ( )x and μ( )x , followed bycomputing ( )c xp and ( )c xs through post-processing, according to:

μρ

λ μρ

( ) = ( )( )

( ) = ( ) + ( )( ) ( )

c cxxx

xx x

x,

2.

4s p

The inverse medium problem can be formulated as the mini-mization of the misfit between the measured response um, atsensor locations, and the computed response corresponding to atrial material profile:

∫ ∫∑λ μ δ Γ

λ μ

( )≔ ( − )·( − ) ( − )

+ ( ) ( )

λ μ Γ=tu u u u x xmin ,

12

d d

, , 5

j

N T

m m j, 1 0

r

m

where u is the computed response, obtained from the forward pro-blem, which is governed by the initial- and boundary-value problem(1) and (2); is the objective functional,4 Nr denotes the total numberof sensors, T is the total observation time, Γm is the part of the groundsurface where the sensor response has been recorded, δ( − )x xj is theDirac delta function, which formalizes measurements at sensor loca-tions xj, and λ μ( ), is a regularization term, discussed next.

Inverse medium problems are notoriously ill-posed. They sufferfrom solution multiplicity; that is, material profiles that are verydifferent from each other, and, potentially non-physical, can becomesolutions to the misfit minimization problem. Regularization of λ andμ alleviates the ill-posedness. Herein, we use a Tikhonov regulariza-tion scheme [34], which penalizes large material gradients:

∫ ∫λ μ λ λ Ω μ μ Ω( ) = ∇ ·∇ + ∇ ·∇ ( )λ

Ω

μ

Ω

R R,

2d

2d , 6RD RD

where Rλ and Rμ are the so-called λ- and μ-regularization factor,respectively, and control the amount of penalty imposed via (6) onthe gradients of λ and μ.

4 See [4,7,9,10] for other possibilities.

Page 7: Soil Dynamics and Earthquake Engineering64 A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81. mathematical framework for tackling the inverse medium pro-blem

Fig. 12. Cross sections of cp and cs profiles obtained via full-waveform inversion at =x 0 m and =x 20 m.

5 See Appendix A for the definition and notation.

A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81 69

A solution of (5) requires the satisfaction of the first-orderoptimality conditions [38]. We use the (formal) Lagrangian ap-proach [31,37] to impose the PDE-constraint (1) and (2). Specifi-cally, we introduce Lagrange multiplier vector function w, andLagrange multiplier tensor function T , residing in appropriatefunction spaces, to enforce the initial- and boundary-value pro-blem (1) and (2) in its weak form. It can be shown [14] that theLagrangian functional becomes:

∫ ∫

∫ ∫

∫ ∫

∫ ∫

∫ ∫

∫ ∫

∫ ∫ ∫ ∫

( )

λ μ δ Γ λ μ

μ λ Ω

Λ Λ Λ Ω

ρ Ω

ρ Ω

Γ

Ω

μ Λ Λ Λ Λ Λ

Λ λ Λ Λ Λ Ω

( ) = ( − )·( − ) ( − ) + ( )

− ∇ { [∇ + (∇ ) ] + ( ) }

− ∇ ( + + ¯ )

− · ¨

− · ( ¨ + + + ¯ )

+ ·

− ( ¨ + + + ¯) +

[(∇ ) + (∇ ) + (∇ ) + (∇ ) + (∇ ¯ )

+ (∇ ¯ ) ] + [ ( ) + ( ) + ( ¯ )]

Γ

Ω

Ω

Ω

Ω

Γ

Ω Ω

=

7a

g

t

t

t

t

a b c t

t

a b c t

t

u S w T u u u u x x

w u u u

w S S S

w u

w u u u u

w

T S S S S T

u u u u u

u T u u u

, , , , ,12

d d ,

: div d d

: d d

d d

d d d

d d

: d d d

:

: div div div d d .

j

Nr T

mm m j

T T

T Te

Tp

Tw

T

T

T

n

T T

e eT

p pT

w

wT

e p w

1 0

0 RD

0 PML

0 RD

0 PML

0 NRD

0 PML 0 PML

2.2.1. Optimality conditionsWe compute the first-order optimality conditions for (5) by

using the Lagrangian functional (7a). To this end, the Gâteauxderivative5 (or first variation) of the Lagrangian functional withrespect to all variables must vanish.

We start by taking the derivatives of the Lagrangian functionalwith respect to w and T in directions w and T , residing in

appropriate function spaces, and setting it to zero:

λ μ′( )( ˜ ˜ ) = ( )u S w T w T, , , , , , 0. 8

This results in the forward problem, whose semi-discrete form is (3).Next, we require that the derivative of with respect to u and

S, in directions u and S, residing in appropriate function spaces,vanish. This yields

λ μ′( )( ˜ ˜) = ( )u S w T u S, , , , , , 0, 9

which results in the adjoint problem, with semi-discrete form

⎢⎢⎢

⎥⎥⎥

⎢⎢⎢

⎥⎥⎥

⎢⎢⎢

⎥⎥⎥

⎢⎢⎢

⎥⎥⎥

=−

+

( )t

yy

My

0 I 00 0 I

G K C

yyy

00

f

dd

,

10T T T

0

1

2

0

1

2adj

Page 8: Soil Dynamics and Earthquake Engineering64 A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81. mathematical framework for tackling the inverse medium pro-blem

Fig. 13. Cross sections of cp and cs obtained via full-waveform inversion at =x 40 m and =x 60 m.

Fig. 14. Poisson's ratio obtained via processing of the full-waveform inversionprofiles.

A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–8170

where = ¯y d0adj, =y d1

adj, = y d2adj, = ( )d w T,h

ThT Tadj is the vector of

nodal unknowns comprising discrete values of w in Ω Ω¯ ∪ ¯RD PML

and discrete values of T only in ΩPML , and fadj is a vector com-prising the misfit at sensor locations.6 Moreover, system matricesM C K G, , , , are identical to those of the forward problem.

We remark that the adjoint problem (10) is a final-value pro-blem (i.e., ( ) =Ty 00 , ( ) =Ty 01 , and ( ) =Ty 02 ), and thus, is solvedbackwards in time.7 We use spectral elements to render M diag-onal, and employ an explicit RK-4 scheme to integrate (10) in time[14].

Lastly, we require vanishing derivatives of with respect to λand μ in directions λ and μ, which yield the gradients with respectto λ and μ, respectively:

λ μ λ′( )( ˜) = ( )u S w T, , , , , 0, 11a

λ μ μ′( )( ˜) = ( )u S w T, , , , , 0. 11b

6 Superscript “adj” refers to the adjoint problem.7 See [22,35] for alternative approaches.

Page 9: Soil Dynamics and Earthquake Engineering64 A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81. mathematical framework for tackling the inverse medium pro-blem

A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81 71

Discretization of (11) results in:

˜ = + ( )λ

λλ λRMg g g , 12areg mis

˜ = + ( )μ

μμ μRMg g g , 12breg mis

where M is a mass-like matrix, λg and μg are the vector of discretevalues of the gradient for λ and μ, respectively, and λgreg,

μgreg and

λgmis,μgmis are the associated vectors corresponding to the

Fig. 15. Locations of the SASW-method array lines. (For interpretation of the re-ferences to color in this figure caption, the reader is referred to the web version ofthis paper.)

Fig. 16. Comparison of site shear wave velocity profiles obtained via SASW and full-waveprofiles are shown along 3 array lines at =y 10 m, 30 m, 50 m.

regularization-part and misfit-part of λg and μg . We refer to [14]for matrix and vector definitions, guidelines for selecting theregularization factors λR , μR , and a comprehensive treatment of theinverse medium problem in three-dimensional PML-truncateddomains.

2.2.2. Gradient-based optimizationIn a gradient-based scheme, discrete material properties are

updated iteratively, using the gradient of the objective functional(5) with respect to λ and μ, until the misfit is minimized. We startwith an assumed initial spatial distribution of the material para-meters λ( )x and μ( )x , and solve the state problem (3) to obtain

= ( )d u S, ,hT

hT Tst . With the misfit known, we solve the adjoint pro-

blem (10) and obtain = ( )d w T,hT

hT Tadj . With uh and wh known, the

(reduced) material gradients, i.e., λg and μg , can be computed from(12). Next, the vector of material values, at iteration +k 1, can beobtained by using a search direction via:

form inversion (FWI); SASW profile is horizontally layered; FWI is spatially variable:

Fig. 17. Inverted profile for cs via the SASW method.

Page 10: Soil Dynamics and Earthquake Engineering64 A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81. mathematical framework for tackling the inverse medium pro-blem

Fig. 18. Comparison of measured surface displacement time-histories against those resulting from the SASW and full-waveform-inversion-based profiles at the controlsensors (sensors not used for the full-waveform inversion).

A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–8172

λ λ α= + ( )λ λ+ s , 13ak k k k1

μ μ α= + ( )μ μ+ s , 13bk k k k1

where λ and μ comprise the vector of discrete values for λ and μ,respectively, α λ

k , α μk are step lengths, and λsk ,

μsk are the searchdirections for λk and μk. Herein, we use the L-BFGS method tocompute the search directions [26].8 Moreover, to ensure sufficientdecrease of the objective functional at each inversion iteration, weemploy an Armijo backtracking line search [26]. Furthermore, toalleviate the ill-posedness, and assist the inversion process, webias the search direction of λ based on that of μ at the earlyiterations of the inversion process, as detailed in [14].

8 We store m¼15 L-BFGS vectors.

3. Field experiments

We report next on the design and data processing of two fieldexperiments that we performed at the NEES@UCSB site in GarnerValley, California. The first field experiment is aimed at collectingdata for the three-dimensional cp and cs profiling of the site,using the full-waveform inversion methodology discussed earlier,and the second experiment seeks to collect data for re-constructing a one-dimensional cs profile, using the SASWmethod. We first introduce the equipment that we used in thefield experiments, followed by the experimental layout. Next, wediscuss the key characteristics of the surface load signal, andcomment on the post-processing needs of the collected data. Wethen present the full-waveform-inversion-based cp and cs profilesof the site, and compare them against the cs profiling obtainedfrom the SASW method. Lastly, we show simulated time-historyresponse at control sensor locations due to the soil profiles ob-tained from the full-waveform inversion, and that of the SASWmethod. We compare the simulated time histories with actual

Page 11: Soil Dynamics and Earthquake Engineering64 A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81. mathematical framework for tackling the inverse medium pro-blem

Fig. 19. Comparison of measured surface displacement time-histories against those resulting from the full-waveform-based inversion and SASW ( =x 0 m).

A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81 73

field measurements in an attempt to assess the quality of theinversion.

3.1. Field equipment

We use the T-Rex Vibroseis of the NEES@UTexas equipment sitefor applying loads on the ground surface (Fig. 3(a)). The equipmentcan apply vertical loads with a maximum force amplitude of267 kN within a frequency range of 12–120 Hz. The Vibroseis isalso capable of applying loads outside this frequency range, albeit,at a lower amplitude. We record the resulting motion on theground surface with 1-Hz geophones (Fig. 3(b)).

3.2. Experimental site and layout

We aim at the seismic site characterization of a subregion of theNEES@UCSB site in Southern California. It is located in a narrowvalley, within the Peninsular Ranges batholith, 23 km east of Hemet,20 km southwest of Palm Springs, and is situated just 7 km from theSan Jacinto fault, and 35 km from the San Andreas fault [1].

We consider a portion of the site of length and width 66 m ×68 m, and 40 m depth. The experiment layout is shown in Fig. 4.The main grid lines are 10 m apart. The sensors are shown with(black) solid circles, while sources are indicated by (red) squares.Overall, we “shake” the site at 12 locations, and record the site'sresponse using 35 sensors. The experiment was performed onMarch 13, 2012 [6].

3.3. Surface load signal characteristics

Loads that have a broad range of frequencies can probe the sitemore effectively than those containing only a few frequencies. In thisstudy, we use chirps, since their frequency content can be adjusted toincrease progressively from low to higher frequencies with respect totime. They have been used successfully in site characterization [19],and are also common in radar and other geophysical applications.Specifically, we use a linear chirp of the form:

⎛⎝⎜

⎛⎝⎜

⎞⎠⎟

⎞⎠⎟π( ) = +f t f

kttsin 2

2,0

Page 12: Soil Dynamics and Earthquake Engineering64 A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81. mathematical framework for tackling the inverse medium pro-blem

Fig. 20. Comparison of measured surface displacement time-histories against those resulting from the full-waveform-based inversion ( =+x 10 m).

A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–8174

where f0 is the starting frequency, typically decided based on thegeophone's resonant frequency, and k is the chirp rate, which con-trols how fast the frequency content of the signal increases withrespect to time. We use characteristic parameters =f 3 Hz0 , and

= −k 2.8 s 2, with a total active time duration of 2.5 s, which results ina loading signal with strong frequency components within the rangeof –3 10 Hz. Fig. 5 shows the (theoretical) time history of the load andits corresponding Fourier spectrum.

It is worth mentioning that for the numerical solution of theadjoint problem (10), and evaluation of the gradients (12), solutionhistory of the state variables9 is required at all time steps. Thus, asignal with a longer time duration may necessitate more storage10

in addition to increasing the computational cost. Therefore, thetemporal duration of the signal should be only long enough toprobe the site effectively.

9 Storage of the adjoint variables can be avoided by updating the gradients onthe fly, while solving the adjoint problem backwards in time.

10 Checkpointing strategies may be exploited to relieve storage requirements,albeit at additional computational cost [11].

We remark that T-Rex is unable to generate a chirp signal in itsexact form, nor is this vital. Thus, by installing accelerometers onits baseplate, we compute the force that the equipment applies onthe ground surface. This load is then used in the solution of theforward problem (3).

3.4. Processing of the collected field data

The recorded field data is inevitably contaminated with noise. Weidentify the parts of the signal (in the frequency-domain) with lowsignal-to-noise ratio, and eliminate them, by using an equiripple fi-nite-impulse-response (FIR) filter, which preserves the signal's phaseinformation [32]. Specifically, we use a bandpass filter, with high andlow cuts of 1.5 Hz and 14 Hz, respectively, and high and low slopes of66.6 dB/Hz and 15 db/Hz, respectively. We use a sampling frequencyof 200 Hz for digital data collection, which, according to the Nyquistsampling theorem [32], is adequate, and prevents aliasing, if therecorded data is contaminated with noise up to 100 Hz. Moreover, toreduce the effects of ambient noise, we repeat each loading fivetimes, and use the averaged recording for inversion.

Page 13: Soil Dynamics and Earthquake Engineering64 A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81. mathematical framework for tackling the inverse medium pro-blem

Fig. 21. Comparison of measured surface displacement time-histories against those resulting from the full-waveform-based inversion ( =+x 30 m).

11 We do not invert for the mass density.

A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81 75

Next, we present a subset of the processed recorded motion atselect sensor locations. Specifically, we consider the displacementtime-history at four sensors, placed at ( ) = ( ) ( )x y, 0, 60 , 40, 60 ,( ) ( )0, 10 , 40, 10 m, due to loading at ( ) = ( )x y, 20, 35 m. Thesesensors are equidistant from the source, and are shown with solidcircles in Fig. 6, while the source is indicated by a square. The loadand displacement time-histories at these locations are shown inFigs. 7 and 8, respectively. Fig. 8 depicts both the raw sensor re-cordings, as well as the processed records, per the earlier outlinedprocedure. Notice that, although there is general agreement be-tween the time-histories of the equidistant sensors, there aredifferences, which are indicative of the heterogeneity of the site.

3.5. Full-waveform inversion using field data

Based on the inversion framework discussed in the precedingsections, we use the collected field data to compute the three-di-mensional cp and cs profiles of the probed site. From the 35 sensorswhere the time-history response of the site was recorded, we useonly 30 of them to drive the inversion, and use the remaining

5 control sensors to validate the resulting profiles. The two sets ofsensors are shown in Fig. 9.

We consider a cubic (regular) domain of length, width, and depth× ×66 m 68 m 40 m. A 10 m-thick PML is placed at the truncation

boundaries. For the PML parameters, we choose α = 5o , β = −500 so1,

and a quadratic profile for the attenuation functions, i.e., m¼2 (see[15] for notation and other details regarding the PML parameters).The mass density is considered to be ρ¼1760 kg/m3 for− ≤ ≤z2 m 0, ρ = 1880 kg/m3 for − ≤ ≤ −z4 m 2 m, andρ = 2000 kg/m3 for − ≤ ≤ −z40 m 4 m, according to prior siteinvestigations.11 The material properties at the interfaces ΓI are ex-tended into the PML. The interior and PML domains are discretizedby quadratic hexahedral spectral elements with 2 m sides (i.e., 27-noded bricks, and quadratic–quadratic pairs of approximation fordisplacement and stress components in the PML, and, also, quadraticapproximation for material properties). The time step is Δ = −t 10 s3 .This discretization leads to 2,433,408 state unknowns, and 379,086material parameters. To probe the site, we use the chirp signal

Page 14: Soil Dynamics and Earthquake Engineering64 A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81. mathematical framework for tackling the inverse medium pro-blem

Fig. 22. Comparison of measured surface displacement time-histories against those resulting from the full-waveform-based inversion ( =+x 40 m).

A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–8176

discussed in Section 3.3. Owing to linearity, we apply all of the 12loads simultaneously, and consider their superimposed responses ateach sensor location, after synchronizing the recorded response.

Due to the construction of the chirps, their frequency contentincreases with time, which can be exploited to advantage by fur-ther regularizing the inversion process. That is, we start with aportion of the signal to drive inversion, and as the inversionevolves, we increase the temporal duration of the signal, thus,progressively including higher frequencies, which allows for pro-file refinement [19,23,24]. Table 1 shows the details of this process.

The inversion process begins with an initial guess for the ma-terial profiles, which is then updated iteratively, until the objectivefunctional (5) is minimized. The speed of convergence, and eventhe success of the inversion itself, relies heavily on the initial as-sumption for the material profiles. Although, in principle, it ispossible to start with a homogeneous initial guess,12 oftentimes,prior information about the site may be of assistance to the

12 A homogeneous initial guess may necessitate a multi-scale approach, using asequence of finer grids, to avoid trapping in local minima [8,16].

optimizer. For instance, a smoothed profile obtained from theSASW method may be used as the initial guess: specifically, weperformed SASW at one location, smoothed out the staircase,layered profile the SASW produced, and used it as initial guess forthe three-dimensional full-waveform inversion.

After 500 iterations, the misfit between the measured responseand the computed response becomes reasonably small, with nostrong update for the material parameters. The corresponding three-dimensional cp and cs profiles for the Garner Valley site are shown inFigs. 10 and 11, respectively, whereas Figs. 12 and 13 show the crosssections of the velocity profiles at =x 0, 20, 40, and 60m. Theseprofiles indicate that the site under investigation is heterogeneous.Having obtained the ( )xcp and ( )xcs profiles, we then seek to computePoisson's ratio ν( )x everywhere in the probed region.13 Poisson's ratiodistribution is shown in Fig. 14. Except for a small region, con-centrated around =x 60 m, Poisson's ratio varies between 0.30 and0.35, which is quite reasonable. Since the P- and S-wave velocities

13 The computation is based on post-processing the inverted-for distributionsof the Lamé parameters λ( )x and μ( )x , according to ν λ λ μ( ) = ( ) ( ( ) + ( ))x x x x/2 .

Page 15: Soil Dynamics and Earthquake Engineering64 A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81. mathematical framework for tackling the inverse medium pro-blem

Fig. 23. Comparison of measured surface displacement time-histories against those resulting from the full-waveform-based inversion ( =+x 60 m).

A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81 77

were obtained by simultaneously inverting for both velocities, andsince we had not imposed a constraint on Poisson's ratio, the fact thatPoisson's ratio ended up being within a narrow and physically ac-ceptable range is a strong indicator of the correctness of the invertedprofiles.

3.6. Comparison with SASW

Having obtained the cp and cs profiles of the site via full-wave-form inversion, next, we compare them against the cs profile ob-tained from the SASW method. While full-waveform inversion usesthe complete waveform of the recorded response at sensor locations,and is capable of imaging three-dimensional, arbitrarily hetero-geneous formations, SASW considers only surface (Raleigh) waves,and outputs a one-dimensional layered profile for the shear wavevelocity: the method relies on the dispersive nature of Raleigh wavesin layered media. A material profile, based on the SASW method, isobtained by: (a) constructing an experimental dispersion curve,based on a field experiment; (b) constructing a theoretical dispersion

curve, by assuming a trial distribution for the material properties of alayered profile; and (c) varying the material properties of the layeredprofile, and repeating the previous step, until a match between theexperimental dispersion curve, and the theoretical dispersion curveis attained [18,33].

We performed field tests using the SASW method to obtain the csprofile of the site along two array lines. The profiles are obtainedalong =x 0 m and 40 m, which are marked with red lines in Fig. 15,and are shown in Fig. 16. The corresponding cs profiles obtained fromthe full-waveform inversion at three different locations along theselines are also shown in these plots. There is very good agreementbetween the profiles at near-surface regions (top 25 m), while at thedeeper part of the site, the full-waveform-inversion-based cs-profilepredicts lower values than those obtained via SASW.

3.7. Time-history comparisons at sensor locations

We compute the time-history response corresponding toa forward wave simulation of the site, based on profilesobtained from full-waveform inversion and that of the SASW

Page 16: Soil Dynamics and Earthquake Engineering64 A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81. mathematical framework for tackling the inverse medium pro-blem

Fig. 24. Comparison of measured surface displacement time-histories against those resulting from the full-waveform-inversion-based profiles, with the groundwater levelset at 6 m (FWI-6), at 20-m (FWI-20), and at infinity (FWI). The response is evaluated at the control sensors (sensors not used for the full-waveform inversion).

A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–8178

method14 (Fig. 17), and compare them against the measuredfield response at various sensor locations. We perform thecomparison in two parts: (1) for the 5 control sensors sp1– sp5,which were not used during the inversion, shown in Fig. 9; and(2) for the remaining 30 sensors that were used during the in-version process. Part 1 is an attempt to validate the profilesobtained via the full-waveform-inversion-based methodology,while part 2 gauges the success of the misfit-minimization al-gorithm when using field data.

Time-history comparisons at the 5 control sensors sp1– sp5 areshown in Fig. 18. These sensors are placed along the

14 To compute the cp profile for the SASW method we set ν = 0.33 when> −z 6 m, and set =c 1500 m/sp otherwise. The = −z 6 m corresponds to an es-

timation of the location of the water table. Furthermore, since the SASW profiles forthe two array lines are almost identical, we average the velocity values for time-history simulations.

=x 0, 10, 30, 40, 60 m array lines. Overall, there is very goodagreement between the measured field response and the responsecomputed based on using the full-waveform inversion profiles.Specifically, the agreement at sensors sp sp sp, ,1 2 3 and sp5 is re-markable. At sp4, the amplitudes are not well-matched; moreover,the measured response and full-waveform-inversion-based re-sponse are out of phase after, approximately, 1.5 s. As it can also beseen, the time-history response at the control sensors computedbased on the SASW profiles, differ from the recorded response:overall, it appears that the SASW-based traces match the fre-quency content of the measurements, but not the amplitude.

Time-history comparisons for the remaining 30 sensors (seeFig. 9) are shown in Figs. 19–23. Time history traces corresponding tothe computed response based on the SASW-inverted profiles areincluded for the first 6 sensors only (Fig. 19). As it can be seen, theSASW traces depart from the recorded response; this trend is true for

Page 17: Soil Dynamics and Earthquake Engineering64 A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81. mathematical framework for tackling the inverse medium pro-blem

Fig. 25. Comparison of measured surface displacement time-histories against those resulting from the full-waveform-inversion-based profiles, with the groundwater levelset at 6 m (FWI-6), at 20-m (FWI-20), and at infinity (FWI). The response is evaluated at the array line =x 0 m.

A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81 79

the remaining 24 sensors too. We remark that it may not be fair tocompare the SASW-based time histories at these locations with thoseresulting from the full-waveform inversion: these sensors were usedduring the inversion process to minimize the misfit between themeasured response and the computed response; therefore, goodagreement between the full-waveform-inversion-based time-his-tories and the measured response is expected, as is indeed the case.Overall, the plots indicate that the optimizer performed successfullyin matching the computed response with the measured response.

3.8. Groundwater level effect

According to prior investigations conducted at the site, thegroundwater table is estimated to be between 6 m and 20 m.While one would have expected that in fully saturated zones, theP-wave velocity would take values closer to 1500 m/s, the full-waveform inversion did not return a cp profile suggestive of sa-turation at any depth up to 40 m. Thus, in this section, we in-vestigate the sensitivity of the computed response to variations ofthe groundwater level. To this end, we perform two forward wave

simulations: one that corresponds to a groundwater level at= −z 6 m, henceforth denoted by FWI-6, and another one, FWI-20,

which corresponds to a groundwater level at = −z 20 m. Wecompare the computed responses corresponding to FWI, FWI-6, andFWI-20, against the measured response. In these simulations, weuse the full-waveform-inversion-based cs profile. Moreover, forsoil depths above the water table, we use the cp profile that hadbeen obtained earlier from full-waveform inversion; by contrast,for soil depths below the water table, we set =c 1500 m/sp uni-formly. The resulting time history comparisons at the 5 controlsensors sp1–sp5 are shown in Fig. 24. It can be seen that the time-histories are not very sensitive to the precise location of thegroundwater level for the considered cases. In Fig. 25, we comparethe time histories for sensors located at =x 0 m, where a con-clusion similar to the above can be drawn. The response at theremaining sensors (not shown) follows a similar trend.

In summary, the elastodynamic model at the heart of the full-waveform-inversion does not account for saturation, and cannotreveal the groundwater level, as the sensitivity results suggest.This limitation is also present in most other site characterization

Page 18: Soil Dynamics and Earthquake Engineering64 A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81. mathematical framework for tackling the inverse medium pro-blem

A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–8180

approaches. We remark that overcoming the limitation is possiblewithin the FWI framework by appropriate modifications. Such anenhancement will be addressed in a future communication.

4. Conclusions

We discussed recent progress in three-dimensional site char-acterization using full-waveform inversion. We reviewed key com-ponents required for developing a general and robust framework toimage the three-dimensional P- and S-wave velocity profiles of near-surface deposits. These components are: (a) a computational modelfor the simulation of elastic waves in domains terminated by PMLs;(b) a PDE-constrained optimization framework, endowed with reg-ularization schemes, through which the computed response is “mat-ched” to the measured response, via iteratively updating an initiallyassumed material distribution for the soil; and (c) guidelines fordesigning a field experiment that can be well-coupled with the al-gorithmic development. We then reported on the imaging of theNEES@UCSB site in Garner Valley, using the reviewed computationalframework. We compared the resulting imaging against the profilingobtained from the SASW method, and reported general agreement.

To drive the inversion process, we used only a subset of thesensors that we had deployed during the field experiment. In anattempt to validate the resulting profiles, we then used the re-maining sensors as controls for time-history comparisons betweenthe measured response, and the computed response based on eitherthe full-waveform-inversion- or SASW-based profiles. The re-ported agreement at the control sensors was remarkably good,which, when considered in tandem with the excellent agreementof the time histories between computed and measured responsesat the rest of the sensors, attests to the validity of the full-wave-form-inversion-based profiles.

To the best of our knowledge, this is the first attempt that fielddata have been successfully used for three-dimensional P- andS-wave velocity full-waveform-based profiling of near-surfacedeposits, where PMLs have been used for the rigorous treatment ofthe truncation boundaries. Overall, the full-waveform inversionmethodology discussed herein, appears robust, general, and pro-mising for the imaging of near-surface deposits, in support ofgeotechnical site characterization investigations.

Acknowledgments

We wish to thank Dr. Jamison Steidl, Dr. Farn-Yuh Menq, Mr.Paul Hegarty, Mr. David Umberg, Dr. Y.-C. Lin, Ms. Julia Roberts,and Mr. Cecil Hoffpauir for assisting with conducting the fieldexperiments and processing the SASW tests.

Partial support for the authors' research has been provided bythe National Science Foundation under grant awards CMMI-0619078 and CMMI-1030728, and through an Academic AllianceExcellence grant between the King Abdullah University of Scienceand Technology in Saudi Arabia (KAUST) and the University ofTexas at Austin. Support for operation of the NEES@UCSB andNEES@UTexas equipment sites was provided by the National Sci-ence Foundation grant awards CMMI-0927178 and CMMI-0927178, respectively. This support is gratefully acknowledged.

Appendix A. Gradient of a functional

The gradient of a functional →: , where is a Hilbertspace, is defined as the Riesz-representation of the derivative

′( )( ˜ )q q , such that

( ( ) ˜ ) = ′( )( ˜ ) ∀ ˜ ∈ ( )q q q q q, , A.1

where denotes the gradient, and we use the following notationfor the Gâteaux derivative of with respect to q in a direction q:

′( )( ˜ ) = ( + ˜ ) − ( )( )→

hh

q qq q q

lim . A.2h 0

References

[1] Garner Valley Downhole Array. ⟨http://nees.ucsb.edu/facilities/GVDA⟩.[2] Albocher U, Barbone PE, Richards MS, Oberai AA, Harari I. Approaches to ac-

commodate noisy data in the direct solution of inverse problems in in-compressible plane strain elasticity. Inverse Probl Sci Eng 2014;22(8):1307–28.

[3] Amrouche M, Yamanaka H. Two-dimensional shallow soil profiling usingtime-domain waveform inversion. Geophysics 2015;80(1):EN27–41.

[4] Banerjee B, Walsh TF, Aquino W, Bonnet M. Large scale parameter estimationproblems in frequency-domain elastodynamics using an error in constitutiveequation functional. Comp Meth Appl Mech Eng 2013;253:60–72.

[5] Basu U. Explicit finite element perfectly matched layer for transient three-dimensional elastic waves. Int J Numer Methods Eng 2009;77:151–76.

[6] Bielak J, Civilini F, Crempien J, Fathi A, Gee R, Hegarty P, et al. T-Rex shaking atGarner Valley: toward three-dimensional full-waveform inversion; 2013.http://dx.doi.org/10.4231/D3BK16P79.

[7] Bozda E, Trampert J, Tromp J. Misfit functions for full waveform inversionbased on instantaneous phase and envelope measurements. Geophys J Int2011;185(2):845–70.

[8] Bunks C, Saleck F, Zaleski S, Chavent G. Multiscale seismic waveform inversion.Geophysics 1995;60(5):1457–73.

[9] Diaz MI, Aquino W, Bonnet M. A modified error in constitutive equation ap-proach for frequency-domain viscoelasticity imaging using interior data.Comput Methods Appl Mech Eng 2015;296:129–49.

[10] Engquist B, Froese B. Application of the Wasserstein metric to seismic signals.Commun Math Sci 2014;12(5):979–88.

[11] Epanomeritakis I, Akcelik V, Ghattas O, Bielak J. A Newton-CG method forlarge-scale three-dimensional elastic full-waveform seismic inversion. InverseProbl 2008;24(3):034015.

[12] Eslaminia M, Guddati, MN. A novel wave equation solver to increase the ef-ficiency of full waveform inversion. In: SEG technical program expanded ab-stracts; 2014. p. 1028–32.

[13] Fathi A. Full-waveform inversion in three-dimensional PML-truncated elasticmedia: theory, computations, and field experiments [Ph.D. thesis]. The Uni-versity of Texas at Austin; 2015.

[14] Fathi A, Kallivokas LF, Poursartip B. Full-waveform inversion in three-dimen-sional PML-truncated elastic media. Comput Methods Appl Mech Eng2015;296:39–72.

[15] Fathi A, Poursartip B, Kallivokas LF. Time-domain hybrid formulations forwave simulations in three-dimensional PML-truncated heterogeneous media.Int J Numer Methods Eng 2015;101(3):165–98.

[16] Fichtner A, Trampert J, Cupillard P, Saygin E, Taymaz T, Capdeville Y, et al.Multiscale full waveform inversion. Geophys J Int 2013;194(1):534–56.

[17] Gholami A, Mang A, Biros G. An inverse problem formulation for parameter esti-mation of a reaction-diffusion model of low grade gliomas. J Math Biol 2015:1–25.

[18] Joh SH. Advances in the data interpretation technique for spectral-analysis-of-surface-waves (SASW) measurements [Ph.D. thesis]. The University of Texas atAustin; 1996.

[19] Kallivokas LF, Fathi A, Kucukcoban S, Stokoe II KH, Bielak J, Ghattas O. Sitecharacterization using full waveform inversion. Soil Dyn Earthq Eng2013;47:62–82.

[20] Kang JW, Kallivokas LF. The inverse medium problem in heterogeneous PML-truncated domains using scalar probing waves. Comput Methods Appl MechEngrg 2011;200:265–83.

[21] Kucukcoban S. The inverse medium problem in PML-truncated elastic media[Ph.D. thesis]. The University of Texas at Austin; 2010.

[22] Monteiller V, Chevrot S, Komatitsch D, Fuji N. A hybrid method to computeshort-period synthetic seismograms of teleseismic body waves in a 3-D re-gional model. Geophys J Int 2013;192(1):230–47.

[23] Monteiller V, Chevrot S, Komatitsch D, Wang Y. Three-dimensional full wa-veform inversion of short-period teleseismic wavefields based upon the SEM–

DSM hybrid method. Geophys J Int 2015;202(2):811–27.[24] Mora P. Nonlinear two-dimensional elastic inversion of multioffset seismic

data. Geophysics 1987;52(9):1211–28.[25] Na S-W, Kallivokas LF. Direct time-domain soil profile reconstruction for one-

dimensional semi-infinite domains. Soil Dyn Earthq Eng 2009;29:1016–26.[26] Nocedal J, Wright S. Numerical optimization. In: Springer Series in Operations

Research, second edition. New York, NY: Springer; 2006.[27] Oberai AA, Gokhale NH, Doyley MM, Bamber M. Evaluation of the adjoint

equation based algorithm for elasticity imaging. Phys Med Biol 2004;49:2955–74.[28] Oberai AA, Gokhale NH, Feijo GR. Solution of inverse problems in elasticity

imaging using the adjoint method. Inverse Probl 2003;19(2):297–313.[29] Pakravan A, Kang JW, Newtson CM. A Gauss–Newton full-waveform inversion

for material profile reconstruction in viscoelastic semi-infinite solid media.

Page 19: Soil Dynamics and Earthquake Engineering64 A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81. mathematical framework for tackling the inverse medium pro-blem

A. Fathi et al. / Soil Dynamics and Earthquake Engineering 87 (2016) 63–81 81

Inverse Probl Sci Eng 2016;24(3):393–421. http://dx.doi.org/10.1080/17415977.2015.1046861.

[30] Park CB, Miller RD, Xia J. Multichannel analysis of surface waves. Geophysics1999;64:800–8.

[31] Petra N, Stadler G. Model variational inverse problems governed by partialdifferential equations. ICES Report 11-05, The Institute for ComputationalEngineering and Sciences, The University of Texas at Austin; 2011.

[32] Proakis JG, Manolakis DK. Digital signal processing. fourth edition. EnglewoodCliffs, NJ: Prentice Hall; 2006.

[33] Stokoe II KH, Wright SG, Bay JA, Roësset JM. Characterization of geotechnicalsites by SASW method. In: Woods RD, editor. Geophysical characterization ofsites. New Delhi, India: Oxford & IBH Pub. Co.; 1994. p. 15–25.

[34] Tikhonov A. Solution of incorrectly formulated problems and the regulariza-tion method. Sov Math Dokl 1963; 5: 1035/1038.

[35] Tong P, Chen CW, Komatitsch D, Basini P, Liu Q. High-resolution seismic arrayimaging based on an SEM-FK hybrid method. Geophys J Int 2014;197(1):369–95.

[36] Tran KT, McVay M. Site characterization using Gauss–Newton inversion of 2-Dfull seismic waveform in the time domain. Soil Dyn Earthq Eng 2012;43:16–24.

[37] Tröltzsch F. Optimal control of partial differential equations: theory, methods,and applications. In: Graduate studies in mathematics, vol. 112. AmericanMathematical Society; Providence, Rhode island, 2010.

[38] Vogel CR. Computational methods for inverse problems. Frontiers in appliedmathematics. Philadelphia, PA: Society for Industrial and Applied Mathe-matics (SIAM); 2002.

[39] Xie Z, Komatitsch D, Martin R, Matzen R. Improved forward wave propagationand adjoint-based sensitivity kernel calculations using a numerically stablefinite-element PML. Geophys J Int 2014;198(3):1714–47.