Busch 2014

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

IEEE TRANSACTIONS ON GEOSCIENCE AND REMOTE SENSING, VOL. 52, NO.

7, JULY 2014 3947

Improved Characterization of Fine-Texture Soils


Using On-Ground GPR Full-Waveform Inversion
Sebastian Busch, Jan van der Kruk, Member, IEEE, and Harry Vereecken

Abstract—Ground-penetrating radar (GPR) uses the record- content, ion concentration, and soil texture [3], the investigation
ing of electromagnetic waves and is increasingly applied for a of these properties using hydrogeophysical methods improves
wide range of applications. Traditionally, the main focus was on the characterization and the understanding of the various highly
the analysis of the medium permittivity since estimates of the
conductivity using the far-field approximation contain relatively dynamic processes taking place in the subsurface.
large errors and cannot be interpreted quantitatively. Recently, a At the point scale, time-domain reflectometry (TDR) is the
full-waveform inversion (FWI) scheme has been developed that most accurate method to estimate the permittivity and conduc-
is able to reliably estimate permittivity and conductivity values tivity for the same sensing volume [1], [4]–[7]. A relatively
by analyzing reflected waves present in on-ground GPR data. It new method to estimate the apparent relative permittivity of the
is based on a frequency-domain solution of Maxwell’s equations
including far, intermediate, and near fields assuming a 3-D subsur- soil is the Theta probe method operating with a fixed-frequency
face. Here, we adapt the FWI scheme for on-ground GPR to invert impedance sensor [8]. However, compared with TDR results,
the direct ground wave traveling through the shallow subsurface. Theta probe measurements overestimate the water content of
Due to possible interference with the airwaves and other reflec- the soil by ∼1.5 [9], and the performance depends on the texture
tions, an automated time-domain filter needed to be included in the and compaction of the soil [9], [10].
inversion. In addition to the obtained permittivity and conductiv-
ity values, also the wavelet center frequency and amplitude return At the field scale, ground-penetrating radar (GPR) is one of
valuable information that can be used for soil characterization. the promising tools for the electrical characterization of the
Combined geophysical measurements were carried out over a silty subsoil. Until now, many studies have investigated the potential
loam with significant variability in the soil texture. The obtained of GPR for estimating the soil relative permittivity, hereinafter
medium properties are consistent with Theta probe, electromag- referred to as permittivity, which is strongly related to the soil
netic resistivity tomography, and electromagnetic induction results
and enable the formulation of an empirical relationship between water content [11]–[13]. A commonly used method to estimate
soil texture and soil properties. The permittivities and conductivi- the dielectric permittivity is a velocity analysis of the elec-
ties increase with increasing clay and silt and decreasing skeleton tromagnetic waves using the common midpoint (CMP) or the
content. Moreover, with increasing permittivities and conductivi- wide-angle reflection–refraction (WARR) method [14], where
ties, the wavelet center frequency decreases, whereas the wavelet the midpoint or source position remains fixed. By increasing
amplitude increases, which is consistent with the radiation pattern
and the antenna coupling characteristics. the distance between the source and receiver antennas, the
different wave types such as direct airwaves (DAWs), direct
Index Terms—Antenna radiation patterns, conductivity, decon- ground waves (DGWs), and reflected and refracted waves can
volution, electromagnetic measurements, frequency-domain anal-
ysis, ground-penetrating radar (GPR), permittivity, wavelets. be distinguished. Here, the velocity of the subsurface can be
estimated using ray-based techniques such as ground-wave
I. I NTRODUCTION travel-time picking [15], [16], reflected-wave analysis [17]–
[22], refracted-wave analysis [23], or a combined analysis of

T HE characterization of the subsurface with various elec-


tromagnetic methods is of growing importance for a wide
range of applications, such as environmental and agricultural
ground, reflected, and refracted waves [11], [14]. The great
advantage of these methods is that the permittivity can then
be directly computed from the measured data. Estimates of the
engineering studies, where the medium property permittivity subsurface conductivity can be carried out using a far-field am-
and conductivity are determined. Since the dielectric permittiv- plitude decay function [24]. However, these estimates contain
ity is highly correlated to the soil water content [1], [2] and the large errors [25] and are not reliable for the characterization of
electric conductivity depends on soil properties such as water the near subsurface.
Estimates of the apparent electrical conductivity of the sub-
surface can also be carried out using low-frequency electrical
Manuscript received September 19, 2012; revised February 28, 2013 and resistivity tomography (ERT) and electromagnetic induction
July 7, 2013; accepted July 7, 2013. Date of publication October 25, 2013; date
of current version February 27, 2014. This work was supported in part by the (EMI) measurements. Many studies have used ERT and EMI
“TERrestrial ENvironmental Observatories” and in part by the “Transregional measurements to monitor transient hydrological processes [26],
Collaborative Research Centre 32” (TR32). to estimate soil properties such as water content and solute
The authors are with the Institute of Bio- and Geosciences: Agrosphere
(IBG-3), Forschungszentrum Jülich GmbH, 52425 Jülich, Germany (e-mail: concentrations [27], [28] and to obtain electrical conductivity
s.busch@fz-juelich.de; j.van.der.kruk@fz-juelich.de; h.vereecken@fz- profiles in different sensing depths [29]–[31].
juelich.de). Although the traditionally used methods analyzing ERT,
Color versions of one or more of the figures in this paper are available online
at http://ieeexplore.ieee.org. EMI, and GPR data enable to obtain estimates of the soil
Digital Object Identifier 10.1109/TGRS.2013.2278297 permittivity or conductivity, none of the currently available

0196-2892 © 2013 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.
See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
3948 IEEE TRANSACTIONS ON GEOSCIENCE AND REMOTE SENSING, VOL. 52, NO. 7, JULY 2014

methods is able to return two independent medium property the inversion process, which leads to an improvement of the
estimates for the same sensing volume. inversion result. For synthetic data, the proposed optimization
Recent developments in full-waveform inversion (FWI) of procedure worked well even when the initial model and, there-
cross-hole [32]–[35] and off-ground GPRs [36], [37], where fore, the effective wavelet differ strongly from the true model
significant portions of the measured waveform are used, in- parameters. Complex wave phenomena such as offset depen-
dicate the benefits of FWI approach to estimate quantitative dent amplitudes and dispersion due to a low-velocity waveguide
permittivity and conductivity values. All FWIs need knowledge were well reconstructed. In the case of measured waveguide
about the emitted field, which is mostly described by an effec- data, the FWI algorithm reduced the misfit of the initial model,
tive source wavelet. For off-ground GPR, the emitted electric derived from conventional kinematic dispersion inversion, by a
field can be considered as independent from the subsurface factor of four to eight [25], and obtained quantitative values for
[38]. For cross-hole GPR, the radiation characteristics for the the permittivities and conductivities.
antennas can be approximated by the radiation patterns in a In this paper, we 1) extend the on-ground GPR FWI for the
homogeneous space. Here, the exact radiation patterns are well analysis of the DGW; 2) include an automated time-domain
described by the far-field expressions [39] that are often used in filter to filter out interferences with the DAWs and reflections;
ray-based tomography inversion, e.g., [40]. Coupling variations 3) apply the FWI for fine-texture soils with a high variability
also occur, particularly if conditions change from saturated to in the soil water content; 4) compare the obtained results with
unsaturated zone [41], but usually, these effects are not included Theta probe, ERT, and EMI measurements; and 5) correlate
in the analysis since most borehole surveys only are employed the obtained medium permittivities and conductivities with the
in the saturated or unsaturated zone. soil texture (clay, silt, and skeleton) and the wavelet center fre-
For on-ground GPR radiation patterns, far-field expressions quency and amplitude by formulating a linear relationship. The
still differ from the exact radiation patterns at a distance of use of multichannel systems and optimized data acquisition will
seven wavelengths [42]. Moreover, the coupling of the anten- enable large-scale measurements with the presented approach
nas and, therefore, the shape and amplitude of the wavelet that returns quantitative permittivity and conductivity values of
strongly depend on the underlying medium [43]. Consequently, the same sensing volume, which can be applied for a wide range
the estimation of the effective source wavelet and, therefore, of applications.
the application of the FWI are less straightforward. Recently, a
frequency-domain FWI for synthetic and experimental on-
ground reflection GPR data has been introduced that returns II. M ETHODOLOGY
quantitative values of the subsurface permittivity and conduc- The FWI scheme for on-ground CMP GPR data is based
tivity [25]. The inversion approach requires an initial model of on a 3-D frequency-domain solution of Maxwell’s equations
the subsurface medium parameters: permittivity and conductiv- [25] assuming a layered model of the subsurface. Whereas
ity, which mainly influence the velocity and amplitudes, respec- the FWI inversion approach is implemented and validated for
tively. Travel-time inversion usually returns initial values for the reflected waves in [25], here, we adapt the on-ground GPR
relative permittivity. However, reliable conductivity values are inversion scheme for the analysis of the DGW which propagates
difficult to obtain particularly for conductive fine-texture soils. to a depth of up to ∼30 cm [47], [48]. An important part of
Erroneous conductivity values result in wrong amplitudes of the this inversion approach is the use of an exact forward model
effective source wavelet that is needed for the FWI, which, in assuming subsurface homogeneity, isotropy, and infinitesimal
turn, also results in erroneously inverted conductivity values, point-dipole antenna, a sequential optimization of the source
since the conductivity and wavelet amplitudes are coupled. wavelet and starting model parameters as well as a subsequent
To assure a good inversion result and, therefore, an accurate parameter optimization while keeping the wavelet fixed.
estimation of the unknown source wavelet, an essential part
of the inversion consists of a combined optimization of the
effective source wavelet and the medium parameters. A. Forward Modeling
Commonly, the center frequency of the antenna source To further enable the inversion of the DGW, the subsurface
wavelet is assumed to be specified by the manufacturer [44]– is approximated by a homogeneous half-space with permittivity
[46]. However, this is not representative for on-ground GPR ε1 and conductivity σ1 , respectively. Using the starting model
application since the center frequency decreases with increasing parameters m = [ε1 , σ1 ], the subsurface Green’s function is
permittivity values. Therefore, the FWI returns a data-driven calculated using an accurate 3-D frequency-domain forward
effective source wavelet. In this way, changes in the subsurface model of Maxwell’s equation [49]. The electrical field in the
properties and, therefore, the wavelet characteristics are explic- frequency domain is described as the multiplication of the
itly taken into account during the inversion process. source wavelet of a point-dipole source and Green’s function
The novel inversion approach was first applied to synthetic of the medium as follows:
single-layer and waveguide data as well as to measured GPR
data reflecting a single-layer low-velocity waveguide where Ê(fn , xm ) = Ĝ(fn , xm , m) · Ŵ (fn ) (1)
a thin layer of high-permittivity sandy silt is overlying low-
permittivity gravel. Focusing on the analysis of reflected waves, where ˆ indicates the frequency domain and xm and fn are
the effective source wavelets and the starting model of the the specific offsets and frequencies, respectively. Closed-form
subsurface properties were significantly improved throughout expressions for the airwaves and ground waves that travel
BUSCH et al.: IMPROVED CHARACTERIZATION OF FINE-TEXTURE SOILS 3949

directly from the point source toward the point receiver for a data independently. The filter essentially defines the mute zone
broadside source–receiver configuration are given by [24], [43] above and below the ground wave and sets all data points within
these areas to zero. This filter is applied on the measured and
∞ 
1 −ς0 ∂12 synthetic data. Since the forward model calculates the data in
Ĝ(fn , xm , m)= · + J0 (κxm )κdκ
2π Γ0 −Γ1 η1 Γ0 −η0 Γ1 the frequency domain, two additional fast Fourier transforms
κ=0 are needed to implement this time-domain filter in the FWI
(2)
algorithm. Because a limited number of offsets are used in
where ζ0 = jωn μ 0 , η0 = jωn ε0 , η1 = σ1 + jωn ε1 , ωn = the inversion, multichannel operations such as τ − p or ω − k
2πn f , and Γs = κ2 + ηs ς0 . Here, the subscripts s = 0, 1 filters to remove the airwaves and other reflections are less
indicate the properties of the air and soil, j is the imaginary appropriate.
unit, κ is the horizontal wavenumber, and J0 is the zeroth-order
Bessel function.
D. Wavelet Estimation
After filtering of the airwaves and possible reflections, we
B. Far-Field Expression for Ground Wave
have isolated the measured ground wave for a certain number of
Using a high-frequency and large-offset approximation, offsets that contain the same effective wavelet. To estimate this
Green’s function in (2) can be described as [24] effective source wavelet Ŵest (fn ), the overdetermined system
   in the following equation is solved for each separate frequency
Z0 xm
Ĝ(fn , xm , m) = exp −iω n using a least squares approach [25], [51]:
2π(1 − ε1 )x2m c0
   
√ xm σ1 Z1 Ĝ(fn , xm , m) · Ŵest (fn ) = Êobs (fn , xm ). (6)
− ε1 exp −iωn exp − xm (3)
c1 2

where c0 and Z0 are the velocity and electrical impedance of E. Sequential and Simultaneous Source Wavelet and
air, respectively; ε1 , σ1 , c1 , and Z1 are the relative permittivity, Model Optimization
conductivity, velocity, and electrical impedance of the subsur-
face, respectively; and xm is the offset. Rewriting (1) using (3) Since the wavelet amplitudes and medium properties are
returns separate expressions for the DAW Ê0 (fn , xm , m) and coupled, an important aspect for a successful inversion is the
the DGW Ê1 (fn , xm , m) estimation of the unknown source wavelet, which is addressed
  by an iterative sequential optimization process combining the
Z0 xm global and local optimizations of model parameters and source
Ê0 (fn , xm , m) = exp −iωn · Ŵ (fn )
2π(1−ε1 )x2m c0 wavelet. The local minimization is carried out using the sim-
(4) plex search algorithm [52], which belongs to the direct search
√   methods and enables the solving of nonlinear unconstrained op-
− ε1 Z 0 xm σ1 Z1
Ê1 (fn , xm , m) = exp −iωn − xm timization problems. In the case of a homogeneous half-space,
2π(1−ε1 )x2m c1 2
each iteration k = 1, . . . , K includes the following: 1) the
· Ŵ (fn ). optimization of the initial values for ε1 together with the phase
(5) of the effective source wavelet ϕ for each frequency and 2) the
optimization of the initial value for σ1 together with the ampli-
Using these approximated ray-based models that ignore the tude of the wavelet by minimizing the objective functions
vectorial radiation and wave propagation of the electromagnetic ⎛

waves, as well as the angle-dependent reflection coefficients, 1 
M  N

CP (mP ) = · ⎝ Êmod (fn , xm , mP )
relative large errors are obtained compared to exact field expres- M · N m=1 n=1 Ê
sions [42], [50], yielding inaccurate estimates of the subsurface mod (fn , xm , mP )

conductivity [25].
Êobs (fn , xm ) ⎠

(7)
Êobs (fn , xm )
C. Airwave Filtering ⎛


A crucial step for the inversion of the DGW is the effect of 1 
M  N
mod n m A
(f , x , m )
CA (mA ) = · ⎝
the DAW propagating between the transmitting and receiving M · N m=1 n=1 Ê max (xm )
antennas. In the case of near offsets in the CMP/WARR profiles, ⎞

the DAW might interfere with the DGW, and the analysis of the Êobs (fn , xm )
− ⎠ (8)
Ê max (xm )
DGW will become a challenging task since the DAW and DGW
cannot be clearly distinguished. Moreover, in the case of larger
offsets, the DAW may mask weaker subsurface signals and
make them difficult or impossible to see or interpret. Therefore, where Êobs (fn , xm ) is the measured electrical field, Êmod (fn ,
to minimize the effect of the DAW on the DGW FWI result, we xm , mA ) is the modeled electrical field for the parameters
introduce a time-domain filter which acts on each trace of the mP = [ε1 , ϕ] and mA = [σ1 , A], and M and N are the
3950 IEEE TRANSACTIONS ON GEOSCIENCE AND REMOTE SENSING, VOL. 52, NO. 7, JULY 2014

numbers of offsets and frequencies, respectively. Note that, to


avoid the irregular weighting of different offsets in the objective
function CA (mA ), (8) is trace-normalized by the maximum
spectral amplitude Êmax (xm ) at each offset. The last step of
each iteration contains the deconvolution of the effective source
wavelet (6) based on the optimization results mk = [εk , σ k ].
The process is repeated until the objective function
N

1 M 
Êmod (fn , xm , mk )
Fig. 1. Inversion results of the ERT showing lateral and vertical conductivity
variations at the Selhausen test site. The bright colors indicate low conductivi-
k
Cf x (m ) = ·
M · N m=1 n=1 max (x )
Êxm m
ties; the dark colors indicate high conductivities.
TABLE I
Êobs (fn , xm ) S TARTING M ODEL AND S IMULTANEOUS I NVERSION R ESULTS O BTAINED
− (9) F ROM THE FWI OF THE WARR M EASUREMENTS AT THE R ECEIVER
max (x )
Êxm m P OSITIONS AT 10, 30–50, AND 80–110 m AT THE S ELHAUSEN
T EST S ITE . I N M OST C ASES , THE S AME O FFSET R ANGE xFWI
WAS U SED TO D ETERMINE THE R AY-BASED (RB)
fulfills the termination criterion Cf x (mk ) > Cf x (mk−1 ) or the AND F ULL -WAVEFORM (FWI) R ESULTS
maximum iteration number k is reached.
To further address the cross-coupling between the permittiv-
ity and the conductivity, which is not explicitly taken into ac-
count during the sequential optimization process, the optimized
parameters mk serve as input for a simultaneous parameter
optimization while keeping the wavelet fixed, which returns the
final inversion result.

III. C OMBINED A NALYSIS OF THE S OIL P ROPERTIES


To explore the potential of the FWI for a fine-texture soil,
combined Theta probe, ERT, EMI, and GPR measurements
were carried out on a silty loam at our test site in Selhausen.

A. Selhausen Test Site B. Measurement Setup

The Selhausen test site of the Forschungszentrum Jülich Along a profile with a length of 120 m, we combined ERT
GmbH is located in the southern part of the Lower Rhine using the SYSCAL PRO system (IRIS Instruments, Orleans,
Embayment in Germany. Here, Eolian sediments of Pleistocene France) with 120 electrodes and an electrode spacing of 0.25 m
and Eocene ages, with a thickness of up to 1 m, cover Qua- in dipole–dipole configuration. Seven 30-m-long profiles were
ternary sediments, which are mostly fluvial deposits from the measured using the roll-along technique with 15-m overlap.
Rhine/Meuse river and the Rur river system [53]. According to Moreover, EMI measurements were carried out every 1 m in
the U.S. Department of Agriculture textural classification, the the horizontal coplanar (HCP) and vertical coplanar (VCP) ori-
major soil type is silty loam [53]. Since the surface topography entations with EM38 (Geonics, Mississauga, ON, Canada) and
of the test site is weakly inclined (< 4◦ ), colluvial sediments GSSI Profiler (Geophysical Survey Systems, Inc., Salem, NH,
eroded from the upper part of the test site can be found in USA) in VCP and HCP orientations. GPR WARR measure-
the lower part of the test site. The ground water depth shows ments with unshielded 200-MHz antennas (Sensors & Software
seasonal fluctuations between 3 and 5 m below the surface. Inc., Mississauga, ON, Canada) were conducted every 10 m,
A distinct gradient in soil texture is present with a consid- and measurements of the effective permittivity in the top soil
erably higher stone content at the upper part of the field [54]. were carried out using Theta Probe ML2 sensors (Macaulay
In the upper part of the test site, the surface soil contains up to Land Use Research Institute, Aberdeen, U.K., and Delta-T
60% stones with diameters of several centimeters and 10% at Devices, Cambridge, U.K.) every 1 m.
the lower part. Soil samples show that the top soil (0–30 cm)
contains 54% skeleton, 14%, sand, 25% silt, and 7% clay in IV. R ESULTS AND D ISCUSSION
the upper part of the field compared to 9% skeleton, 14% sand,
A. ERT and EMI Inversions
63% silt, and 14% clay in the lower part of the field [Fig. 5(a)].
Due to the geomorphology and soil texture variation, a high The ERT data were inverted with the RES2DINV software
variability in the surface soil water content is detectable. Previ- [56] (Geotomo Software, Penang, Malaysia) which gives a
ous studies in [53] and [55] using GPR, TDR, and volumetric 2-D model consisting of multiple rectangular blocks of different
soil samples showed that a gradient in soil water content and, discrete resistivity values. Optimizing the resistivity (inverse of
therefore, in the permittivity from the upper to the lower part of conductivity) by minimizing the misfit between the calculated
the field is partly related to changes in texture. and measured ERT data returns a resistivity image of the
BUSCH et al.: IMPROVED CHARACTERIZATION OF FINE-TEXTURE SOILS 3951

Fig. 2. Measured WARR after applying a gain function at the receiver positions at (a) 40 and (b) 90 m. The red and blue colors indicate trace-normalized positive
and negative, respectively. For the inversion, the airwave present in the data is muted out. The dashed black lines indicate the selected offset range xFWI for the
FWI; (c) and (d) show (black) the picked maximum amplitudes of the DGW, (dashed red) the fitted ray-based decay function, and (blue) the applied gain function.
Note that, for the inversion of the GPR data, the true amplitudes without applying a gain function are used.

subsurface [56]. A robust inversion scheme that uses the L1- the subsurface and, therefore, a decreasing penetration depth of
norm for data and model space was used because sharp layer the GPR with increasing conductivities. Since reflected waves
boundaries were expected due to the presence of a plowing zone. could not be clearly identified in the measured WARR profiles
The inverted profile shown in Fig. 1 has a root-mean-square by muting the airwaves and applying a gain function to the
error of 3.3%, was obtained after five iterations, and clearly in- data, the data are well suited to validate the extended FWI for
dicates the heterogeneity of the Selhausen test site. In the upper the analysis of the DGW. We estimated the initial permittivity
part of the test site (10–30 m horizontally), the ERT inversion and conductivity values for each WARR measurement using the
results indicate relative low conductivities of 5–10 mS/m and ray-based direct ground expression (2).
apparently no layering, whereas in the lower part of the test site, Next, the FWI is performed for each WARR measurement
a high conductivity layer (5–20 mS/m) of a thickness of 0.4 m along the 120-m-long profile. Table I gives an overview of the in-
is overlying an even more conductive layer (> 20 mS/m). version results and indicates the offset range xFWI and the
In contrast to ERT measurements, current EMI systems only number of traces nFWIx within xFWI used for the inversion, the
return qualitative values for the subsurface conductivity be- initial model for the subsurface permittivity εRB and conduc-
cause of instrument calibration difficulties [57]–[60]. To over- tivity σ RB obtained from ray-based methods, and the inverted
come these limitations, Lavoué et al. [61] proposed a method permittivity εFWI and conductivity σ FWI , as well as the corre-
to calibrate EMI measurements with the electrical conductivity sponding objective functions C RB and C FWI , respectively.
values obtained from ERT inversion. Therefore, the conductiv- The WARR profiles after applying a gain function at the
ity distribution obtained from RES2DINV inversion was used receiver positions at 40 and 90 m, representing GPR mea-
as input in an electromagnetic forward model for synthetic surements on relatively low and higher conductivity media,
EMI data with the same EMI configurations as used for the are shown in Fig. 2(a) and (c), respectively. In both data sets,
experimental measurements. The obtained synthetic EMI data a dominant airwave is presented, which interferes with the
were then used to calibrate the measured data following the ground wave. However, due to these interferences, clipping of
procedure described in [61]. After calibration, the field data the amplitudes for small offsets, the increasing attenuation of
were then inverted for a two-layer subsurface by minimizing the DGW, and a decreasing signal-to-noise ratio for increasing
the misfit between the measured and modeled magnetic fields offsets limits the number of offsets which can be used for the
following the procedure described in [62]. Note that, due to FWI (see also Table I).
metal objects at the surface, the area between 55 and 85 m is Fig. 2(b) and (d) shows the picked maximum amplitudes of
excluded from the calibration of the EMI data. the DGW (black) and the applied gain function (blue). The
fitted far-field decay function as given in (5) (dashed red) and
the slope of the ground wave return ray-based conductivity and
B. GPR FWI
permittivity, respectively, that are used as starting model for the
The relatively high conductivities at the Selhausen test site FWI. Note that the inversion of the data is carried out using true
cause significant attenuation of the electromagnetic waves in amplitudes and no gain function was applied. The subsurface
3952 IEEE TRANSACTIONS ON GEOSCIENCE AND REMOTE SENSING, VOL. 52, NO. 7, JULY 2014

Fig. 3. Time- and frequency-domain amplitudes and phase spectra of the initial and inverted effective wavelets for the WARR measurements at the receiver
positions at (a)–(c) 40 and (d)–(f) 90 m.

Fig. 4. Time-domain traces of the measured data and the initial and inverted models for the WARRs at (a) 40 and (b) 90 m. The amplitudes are trace normalized
to the measured data and show true amplitudes.

properties obtained from the ray-based approach and the full- 1.7 × 10−1 to 1.6 × 10−1 ), whereas the objective function C RB
waveform inverted subsurface properties are shown in Table I. for the WARR at 90 m decreases by 10% (from 2.1 × 10−1 to
The significant improvement of the subsurface model is indi- 1.9 × 10−1 ). The initial (green) and inverted effective wavelets
cated by the smaller FWI objective function C FWI , compared to (dashed red) are shown in Fig. 3. Small variations in the
the ray-based objective function C RB . For the WARR measure- phase of the wavelet [Fig. 3(c) and (f)] indicate a relatively
ment at the receiver position at 40 m, the objective functions good permittivity ε starting model obtained from conventional
C RB of the ray-based model in Table I decrease by 6% (from velocity analysis for the WARR measurements at the receiver
BUSCH et al.: IMPROVED CHARACTERIZATION OF FINE-TEXTURE SOILS 3953

TABLE II
S OIL T EXTURE C ONTENT AT THE WARR M EASUREMENTS
AT THE R ECEIVER P OSITIONS AT 10, 30–50, AND
80–110 m AT THE S ELHAUSEN T EST S ITE

TABLE III
P ERMITTIVITIES AND C ONDUCTIVITIES O BTAINED F ROM THE C OMBINED
T HETA P ROBE , ERT, EMI, AND GPR M EASUREMENTS
AT S ELHAUSEN T EST S ITE

Fig. 5. Results of the measurements. (a) Grain-size distribution at the


Selhausen test site for 0–30-cm depth shows increasing clay and silt contents,
a rather constant sand content, and a decreasing skeleton content for increasing
position. (b) Permittivity ε values obtained from (blue) Theta probe measure-
ments, (dashed black) GPRRB ray-based techniques, and (black) GPRFWI
FWI, as well as (c) the conductivity σ values obtained from ERT, EMI,
GPRRB , and GPRFWI , increase for increasing positions. Between 55 and
85 m, metal objects and cables at and in the subsurface influence the geophysi-
cal measurements.

positions at 40 and 90 m. Comparing the initial and inverted


wavelet amplitudes [Fig. 3(a), (b), (d), and (e)], particularly the increasing clay and silt contents and a decreasing skeleton
wavelet for the WARR at 90 m differs significantly, indicating content. The conductivity values obtained from ERT (green;
an erroneous initial conductivity σ model. 4.3–19.1 mS/m) and EMI (red; 5.2–19.3 mS/m) inversions,
The time-domain traces of the measured data and the initial as well as the values obtained from GPRRB (dashed black;
and inverted models for the WARRs at 40 and 90 m are 5.5–28.2 mS/m) and GPRFWI (black; 2.4–28.5 mS/m), are
presented in Fig. 4. Note that the data are trace normalized shown in Fig. 5(c). Since only the ground wave is used for the
to the measured data and therefore show true amplitudes. In FWI of the GPR data, which is traveling through the upper
contrast to the initial model (dashed green), the measured data ∼30 cm of the subsurface [47], [48], the GPR conductivities
(black) and the inverted model (dashed red) overlie each other in Fig. 5(c) are compared with the mean ERT and EMI
except for some phase mismatch on the first cycle at far offset, conductivities for a depth of 0–30 cm. Here, particularly at the
which is possibly due lateral inhomogeneity. WARR positions at 50, 80, and 90 m where the fine-texture
Fig. 5 and Tables II and III show the grain-size distribution soil content increases, conventional used ground-wave picking
at the Selhausen test site and the results of the combined Theta GPRRB returns either lower or higher conductivity values
probe, ERT, EMI, and GPR measurements. For a depth of than the ERT, EMI, and GPRFWI techniques. Although each
0–30 cm, Fig. 5(a) clearly indicates increasing clay (green; method has a different sensing depth and frequency range,
7%–14%) and silt (red; 25%–63%) contents, a constant and the results are therefore not directly comparable, these
sand content (blue; 13%–14%), and a decreasing skeleton methods indicate similar trends in the electric properties of
content (dashed black; 54%–9%) for increasing position. the soil.
The permittivity values obtained from Theta probe (blue; Between the 55- and 85-m positions along the profile, metal
11.4–17.4) measurements, the GPRRB ray-based technique objects at the surface and cables in the subsurface (60–70 m)
(dashed black; 10.6–17.4), and the GPRFWI FWI (black; influence the geophysical measurements. Therefore, ERT and
8.1–18.5) are shown in Fig. 5(b). The increasing permittivities EMI measurements within this offset range are not used for the
observed with increasing WARR positions are consistent with calibration of the EMI data.
3954 IEEE TRANSACTIONS ON GEOSCIENCE AND REMOTE SENSING, VOL. 52, NO. 7, JULY 2014

Fig. 7. Correlation between permittivity, conductivity, and texture of the soil


indicates a linear relationship for [(a) and (b)] fine-texture (clay and silt) and
(c) coarse-texture (skeleton) soils.

ing skeleton content, i.e., with increasing coarse-texture soil


[Fig. 6(d) and (h)]. The corresponding correlation coefficients
R2 between 0.61 and 0.92 indicate a strong linear relationship
between conductivity, permittivity, and soil texture (clay, silt,
and skeleton). Due to a relatively constant sand content, there is
no correlation with the increasing conductivity and permittivity
[Fig. 6(c)–(g)] for increasing position. These results are in
good agreement with the results of Theta probe, ERT, and EMI
measurements, which indicates the reliability of the on-ground
GPR FWI.
Fig. 7 shows the 3-D correlation between the permittivity
and conductivity values obtained from the GPR FWI and the
(a) clay, (b) silt, and (c) skeleton contents. The calculated
2 2
correlation coefficients are Rclay = 0.55, Rsilt = 0.81, and
2
Rskeleton = 0.44 and clearly indicate a linear relation between
Fig. 6. Correlations between the subsurface conductivities obtained from
measured with ERT, EMI, and GPR inversions, the permittivities obtained the clay, silt, and skeleton contents with the subsurface permit-
from Theta probe measurements and GPR inversion, and the soil texture at tivity and conductivity values.
the Selhausen test site, respectively. [(a), (b), (e), and (f)] Conductivity and This relationship can be parameterized by an orthogonal
permittivity increase with increasing clay and silt contents, whereas [(d) and
(h)] the σ and ε decrease with increasing skeleton content. [(c) and (g)] Due to distance regression
a constant sand constant, there is no correlation with increasing σ and ε from ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
the upper to the lower part of the test site. x x0 vx
⎝ y ⎠ = ⎝ y0 ⎠ + ⎝ v y ⎠ · d (10)
Fig. 6 shows the correlations between the subsurface conduc- z z0 vz
tivity (measured with GPR, ERT, and EMI), the permittivity
(measured with GPR and Theta probes), and the soil texture where (x, y, z) is a data point, (x0 , y0 , z0 ) is the point on the
at the Selhausen test site. As already seen in Fig. 5(c) and regression line L, (vx , vy , vz ) is the vector defining the direc-
(d), the GPR FWI conductivities and permittivities increase tion of L, and d is the distance whose value is varied to define
with increasing clay and silt contents, i.e., with increasing the point (x0 , y0 , z0 ) on L. Minimizing the sum of squared
content of fine-texture soil [Fig. 6(a), (b), (e), and (f)], whereas distances from (x0 , y0 , z0 ) to L returns the best fitting line with
the conductivities and permittivities decrease with increas- the direction (vx , vy , vz ). The orthogonal distance regression
BUSCH et al.: IMPROVED CHARACTERIZATION OF FINE-TEXTURE SOILS 3955

Fig. 9. (a) Normalized time-domain source wavelet for (black) the WARRs at
10, 30–50, and 80–110 m and (red) the calculated mean wavelet; (b) and (c)
indicate the corresponding maximum wavelet amplitude and center frequency.
Fig. 8. Data-driven time-domain (a) and frequency-domain amplitude spectra
(b) of the optimized wavelets for the WARR’s at 10 m, 30–50 m, and
80–110 m. Note that the wavelets are normalized to the wavelet with the
maximum amplitude (WARR at receiver position 110 m) and thus shifted
in time.

for the soil texture and the permittivity ε and conductivity σ


values obtained from the FWI is parameterized by
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
ε 15.13 0.27
⎝ σ ⎠= ⎝ 15.17 ⎠ − ⎝ 0.94 ⎠ · d (11)
Mclay 11.36 0.20
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
ε 15.13 0.17
⎝ σ ⎠= ⎝ 15.17 ⎠ − ⎝ 0.57 ⎠ · d (12)
Msilt 50.32 0.80
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
ε 15.13 0.16
⎝ σ ⎠= ⎝ 15.17 ⎠ − ⎝ 0.52 ⎠ · d (13)
Mskeleton 24.73 −0.84

where Mclay , Msilt , and Mskeleton are the clay, silt, and skeleton
contents, respectively.
Fig. 8 shows the inverted data-driven effective wavelets
for the WARRs at the receiver positions at 10, 30–50, and Fig. 10. Correlations between the wavelet center frequency and amplitude
and the subsurface permittivity and conductivity, respectively. The correlations
80–110 m in the time domain [Fig. 8(a)] and the corresponding indicate [(a) and (b)] an increasing wavelet amplitude WA with increasing
amplitude spectra [Fig. 8(b)]. Here, the zero crossings are conductivities σ and ε and [(c) and (d)] a decreasing wavelet center frequency
aligned to enable a comparison of the wavelet shape and ampli- Wf c with increasing σ and ε.
tude. Although the amplitude of the effective wavelets differs at
each midpoint position, the shapes of the wavelets are similar. For a half-wavelength dipole antenna in a homogeneous
Fig. 9(a) shows the normalized wavelets (black) from Fig. 8(a) half-space, the resonance frequency depends on the relative
as well as the calculated mean wavelet (red). With increasing permittivity of the material surrounding the antenna, which,
permittivity ε and conductivity σ values for increasing position in turn, depends on the permittivity of the material supporting
along the profile (see Fig. 5), the wavelet center frequency fc the metal antenna, the air, the ground material and roughness,
shifts to lower frequencies [Fig. 9(c)], whereas the wavelet and the antenna height [63]. For an increasing permittivity ε,
amplitude ŴA (fc ) increases [Fig. 9(b)]. This is also obvious in the antenna appears to be electrically longer, resulting in a
Fig. 10, where fc and ŴA (fc ) show a strong correlation with σ decrease of the antenna center frequency fc . In addition, due to
and ε. The correlation coefficients R2 in the range of 0.44–0.94 a decrease in fc , the antenna system emits more low-frequency
clearly indicate that the effective wavelet is influenced by the energy, which, in turn, results in an increasing wavelet ampli-
subsurface properties. tude ŴA (fc ).
3956 IEEE TRANSACTIONS ON GEOSCIENCE AND REMOTE SENSING, VOL. 52, NO. 7, JULY 2014

All inverted wavelet characteristics are consistent with the ACKNOWLEDGMENT


changing electrical length of the antenna and the radiation
The authors would like to thank the anonymous reviewers and
characteristics for ground coupled antennas. Note that, by
the Editor A. Plaza for their help in improving the manuscript.
carrying out FWI, also quantitative values for the wavelet
center frequency and amplitude are obtained that might also
be used for an improved characterization of the subsurface. In
R EFERENCES
addition, radiation patterns of horizontal electric dipoles show
that the wave energy emitted in the subsurface, characterized by [1] G. C. Topp, J. L. Davis, and A. P. Annan, “Electromagnetic determination
of soil water content, measurements in coaxial transmission lines,” Water
ŴA (fc ), is increasing with increasing permittivity [64], [65] Resour. Res., vol. 16, no. 3, pp. 574–582, Jun. 1980.
and conductivity [66], which seem to be dominant over the [2] A. Tabbagh, C. Camerlynck, and P. Cosenza, “Numerical modeling for
decreasing radiated power due to a possible deterioration of the investigating the physical meaning of the relationship between relative
dielectric permittivity and water content of soils,” Water Resour. Res.,
impedance matching condition. vol. 36, no. 9, pp. 2771–2776, Sep. 2000.
[3] J. D. Rhoades, P. A. C. Raats, and R. J. Prather, “Effects of liquid phase
electrical conductivity, water content, and surface conductivity on bulk
V. C ONCLUSION soil electrical conductivity,” Soil Sci. Soc. Amer. J., vol. 40, no. 5, pp. 651–
655, Sep. 1976.
A recently developed FWI approach for on-ground GPR has [4] F. N. Dalton, W. N. Herkelrath, D. S. Rawlines, and J. D. Rhoades,
been extended and validated for the inversion of the DGW “Time-domain reflectometry: Simultaneous measurement of soil water
content and electrical conductivity with a single probe,” Science, vol. 224,
to characterize the shallow part of the subsurface, assuming no. 4652, pp. 989–990, Jun. 1984.
a locally homogeneous half-space. In the case of on-ground [5] W. R. Whalley, “Considerations on the use of time-domain reflectome-
GPR CMP/WARR data, the proposed method indicates the try (TDR) for measuring soil-water content,” J. Soil Sci., vol. 44, no. 1,
pp. 1–9, Mar. 1993.
high information content present in GPR data. The permittivity [6] M. A. Malicki, R. Plagge, and C. H. Roth, “Improving the calibration of
and conductivity values obtained from the inversion of WARR dielectric TDR soil moisture determination taking into account the solid
data measured on a silty loam are in very good agreement soil,” Eur. J. Soil Sci., vol. 47, no. 3, pp. 357–366, Sep. 1996.
[7] G. C. L. Wyseure, M. A. Mojid, and M. A. Malik, “Measurement of
with the results obtained from conventional used Theta probe volumetric water content by TDR in saline soils,” Eur. J. Soil Sci., vol. 48,
(effective dielectric permittivity), ERT, and EMI measurements no. 2, pp. 347–354, Jun. 1997.
(electric conductivity), respectively. Here, in contrast to con- [8] G. J. Gaskin and J. D. Miller, “Measurement of soil water content using a
simplified impedance measuring technique,” J. Agric. Eng. Res., vol. 63,
ventional used GPR ray-based techniques, the FWI returns reli- no. 2, pp. 153–159, Feb. 1996.
able conductivity values particularly for conductive fine-texture [9] D. A. Robinson, C. M. K. Gardner, and J. D. Cooper, “Measurement of rel-
soils. ative permittivity in sandy soils using TDR, capacitance and theta probes,
comparison, including the effects of bulk soil electrical conductivity,”
Since the novel inversion approach enables to obtain one J. Hydrol., vol. 223, no. 3/4, pp. 198–211, Oct. 1999.
permittivity value and one conductivity value for the same [10] N. Sarani and P. Afrasjab, “Effect of soil texture on moisture measurement
sensing volume, this technique also allows the formulation of accuracy with Theta probe ML2 in Sistan region,” in Proc. Int. Conf.
Chem., Ecol. Environ. Sci., 2012, pp. 114–117.
an orthogonal distance regression parameterization between the [11] R. A. van Overmeeren, S. V. Sariowan, and J. C. Gehrels, “Ground pen-
soil electrical properties and texture (clay, silt, and skeleton). etrating radar for determining volumetric soil water content; Results of
Here, a combination of soil moisture measurements and on- comparative measurements at two test sites,” J. Hydrol., vol. 197, no. 1–4,
pp. 316–338, Oct. 1997.
ground GPR measurements will also attribute the obtained [12] K. W. Weiler, T. S. Steenhuis, J. Boll, and K. S. J. Kung, “Comparison
changes in the subsurface permittivity and conductivity to of ground penetrating radar and time domain reflectometry as soil water
changes in the soil moisture and texture, respectively. sensors,” Soil Sci. Soc. Amer. J., vol. 62, pp. 1237–1239, 1998.
[13] J. A. Huisman, S. Sperl, W. Bouten, and J. M. Verstraten, “Soil water
Moreover, strong correlations between the subsurface per- content measurements at different scales, accuracy of time domain re-
mittivity and conductivity and the wavelet amplitude and center flectometry and ground-penetrating radar,” J. Hydrol., vol. 245, no. 1–4,
frequency clearly show the benefits of the on-ground GPR pp. 48–58, May 2001.
[14] J. A. Huisman, S. S. Hubbard, J. D. Redman, and A. P. Annan, “Measuring
FWI for the improved characterization of the subsurface. This soil water content with ground penetrating radar, A review,” Vadose Zone
indicates the need for an accurate estimated source wavelet for J., vol. 2, no. 4, pp. 476–491, 2003.
ground coupled GPR antennas. [15] L. W. Galagedara, G. W. Parkin, and J. D. Redman, “An analysis of
the ground-penetrating radar direct ground wave method for soil water
Note that the inversion results characterizing the subsurface content measurement,” Hydrol. Process., vol. 17, no. 18, pp. 3615–3628,
and antenna properties are based on a limited number of WARR Dec. 2003.
measurements and the inversion of the DGW but can easily [16] C. M. Steelman and A. L. Endres, “An examination of direct ground wave
soil moisture monitoring over an annual cycle of soil conditions,” Water
be extended to a multilayer model. Furthermore, acquiring and Resources Res., vol. 46, no. 11, p. W11 533, Nov. 2010.
inverting quantitative medium properties on large scales offer [17] S. Tillard and J. C. Dubois, “Analysis of GPR data—Wave-propagation
considerable potential for a wide range of applications. As a velocity determination,” J. Appl. Geophys., vol. 33, no. 1–3, pp. 77–91,
Jan. 1995.
next step, the results obtained from the FWI can be used to [18] R. J. Greaves, D. P. Lesmes, J. M. Lee, and M. N. Toksoz, “Velocity vari-
estimate the soil moisture content using the permittivities and ations and water content estimated from multi-offset, ground-penetrating
Topp’s equation as well as the conductivities and Archie’s law. radar,” Geophysics, vol. 61, no. 3, pp. 683–695, May 1996.
[19] A. L. Endres, W. P. Clement, and D. L. Rudolph, “Ground penetrating
Moreover, inverting for the constitutive parameter in Archie’s radar imaging of an aquifer during a pumping test,” Groundwater, vol. 38,
law will probably improve the characterization of soils. no. 4, pp. 566–576, Jul. 2000.
As a next step, the proposed FWI approach can be extended [20] S. Garambois, P. Senechal, and H. Perroud, “On the use of combined
geophysical methods to assess water content and water conductivity
for multilayer models and can be used as a starting model for a of near-surface formations,” J. Hydrol., vol. 259, no. 1–4, pp. 32–48,
2.5-D or 3-D on-ground GPR FWI. Mar. 2002.
BUSCH et al.: IMPROVED CHARACTERIZATION OF FINE-TEXTURE SOILS 3957

[21] R. W. Jacob and J. F. Hermance, “Assessing the precision of GPR velocity [44] Z. Wu and C. Liu, “An image reconstruction method using GPR data,”
and vertical two-way travel time estimates,” J. Environ. Eng. Geophys., IEEE Trans. Geosci. Remote Sens., vol. 37, no. 1, pp. 327–334, Jan. 1999.
vol. 9, pp. 143–153, 2004. [45] J. Buchner, U. Wollschläger, and K. Roth, “Inverting surface GPR data
[22] J. H. Bradford, “Measuring water content heterogeneity using multifold using FDTD simulation and automatic detection of reflections to esti-
GPR with reflection tomography,” Vadose Zone J., vol. 7, no. 1, pp. 184– mate subsurface water content and geometry,” Geophysics, vol. 77, no. 4,
193, Feb. 2008. pp. H45–H55, 2012.
[23] R. N. Bohidar and J. F. Hermance, “The GPR refraction method,” [46] M. Bano, “Modelling of GPR waves for lossy media obeying a complex
Geophysics, vol. 67, no. 5, pp. 1474–1485, 2002. power law of frequency for dielectric permittivity,” Geophys. Prospect.,
[24] J. van der Kruk and E. C. Slob, “Effective source wavelet determination,” vol. 52, no. 1, pp. 11–26, Jan. 2004.
in Proc. 9th Int. Conf. Ground Penetrating Radar, Expanded Abstracts, [47] L. W. Galagedara, G. W. Parkin, J. D. Redman, P. von Bertoldi, and
2002, pp. 144–149. A. L. Endres, “Field studies of the GPR ground wave method for estimat-
[25] S. Busch, J. van der Kruk, J. Bikowski, and H. Vereecken, “Quantitative ing soil water content during irrigation and drainage,” J. Hydrol., vol. 301,
conductivity and permittivity estimation using full-waveform inversion of no. 1–4, pp. 182–197, Jan. 2005.
on-ground GPR data,” Geophysics, vol. 77, no. 5, pp. H79–H91, 2012. [48] L. W. Galagedara, J. D. Redman, G. W. Parkin, A. P. Annan, and
[26] K. Kemna, J. Vanderborght, B. Kulessa, and H. Vereecken, “Imaging and A. L. Endres, “Numerical modeling of GPR to determine the direct ground
characterisation of subsurface solute transport using electrical resistivity wave sampling depth,” Vadose Zone J., vol. 4, no. 4, pp. 1096–1106,
tomography (ERT) and equivalent transport models,” J. Hydrol., vol. 267, Nov. 2005.
no. 3/4, pp. 125–146, 2002. [49] J. van der Kruk, R. Streich, and A. G. Green, “Properties of surface wave-
[27] A. Ramirez, W. Daily, D. LaBrecque, E. Owen, and D. Chesnut, “Moni- guides derived from separate and joint inversion of dispersive TE and TM
toring an underground steam injection process using electrical resistance GPR data,” Geophysics, vol. 71, no. 1, pp. K19–K29, Jan./Feb. 2006.
tomography,” Water Resour. Res., vol. 29, no. 1, pp. 73–87, Jan. 1993. [50] J. van der Kruk, R. Streich, and M. Grasmueck, “Toward true-amplitude
[28] Q. Z. Zhou, J. Shimada, and A. Sato, “Three-dimensional spatial and vector migration of GPR data using exact radiation patterns,” in Advances
temporal monitoring of soil water content using electrical resistivity to- in Near-Surface Seismology and Ground-Penetrating Radar, Geophys.
mography,” Water Resour. Res., vol. 37, no. 2, pp. 273–285, Feb. 2001. Dev. Ser., vol. 15, R. D. Miller, J. H. Bradford, and K. Holliger, Eds.
[29] B. Borchers, T. Uram, and J. M. H. Hendrickx, “Tikhonov regularization Tulsa, OK, USA: Soc. Exploration Geophys., 2010, ch. 6, pp. 97–116.
of electrical conductivity depth profiles in field soils,” Soil Sci. Soc. Amer. [51] R. Streich and J. van der Kruk, “Characterizing a GPR antenna system
J., vol. 61, no. 4, pp. 1004–1009, Jul. 1997. by near-field measurements,” Geophysics, vol. 72, no. 5, pp. A51–A55,
[30] J. M. H. Hendrickx, B. Bochers, D. L. Corwin, S. M. Lesch, 2007.
A. C. Hilgendorf, and J. Schlue, “Inversion of soil conductivity profiles [52] J. C. Lagarias, J. A. Reeds, M. H. Wright, and P. E. Wright, “Convergence
from electromagnetic induction measurements, Theory and experimental properties of the Nelder–Mead simplex method in low dimensions,” Siam
verification,” Soil Sci. Soc. Amer. J., vol. 66, no. 3, pp. 673–685, 2002. J. Optim., vol. 9, no. 1, pp. 112–147, 1998.
[31] T. Saey, D. Simpson, K. Vermeersch, L. Cockx, and M. Van Meirvenne, [53] L. Weihermüller, J. S. Huisman, S. Lambot, M. Herbst, and H. Vereecken,
“Comparing the EM38DD and DUALEM-21S sensors for depth-to-clay “Mapping the spatial variation of soil water content at the field scale
mapping,” Soil Sci. Soc. Amer. J., vol. 73, no. 1, pp. 7–12, Jan. 2009. with different ground penetrating radar techniques,” J. Hydrol., vol. 340,
[32] J. R. Ernst, A. G. Green, H. Maurer, and K. Holliger, “Application of a no. 3/4, pp. 205–216, Jul. 2007.
new 2D time-domain full-waveform inversion scheme to crosshole radar [54] J. Vanderborght, A. Graf, C. Steenpass, B. Scharnagl, N. Prolingheuer,
data,” Geophysics, vol. 72, no. 5, pp. J53–J64, Sep. 2007. M. Herbst, H. J. Hendricks Franssen, and H. Vereecken, “Within-field
[33] G. A. Meles, J. van der Kruk, S. A. Greenhalgh, J. R. Ernst, H. Maurer, variability of bare soil evaporation derived from eddy covariance mea-
and A. G. Green, “A new vector waveform inversion algorithm for simul- surements,” Vadose Zone J., vol. 9, no. 4, pp. 943–954, Nov. 2010.
taneous updating of conductivity and permittivity parameters from com- [55] K. Z. Jadoon, S. Lambot, B. Scharnagl, J. van der Kruk, E. Slob, and
bination crosshole/borehole-to-surface GPR data,” IEEE Trans. Geosci. H. Vereecken, “Quantifying field-scale surface soil water content from
Remote Sens., vol. 48, no. 9, pp. 3391–3407, Sep. 2010. proximal GPR signal inversion in the time domain,” Near Surf. Geophys.,
[34] A. Klotzsche, J. van der Kruk, G. A. Meles, J. Doetsch, H. Maurer, and vol. 8, no. 6, pp. 483–491, Dec. 2010.
N. Linde, “Full-waveform inversion of cross-hole ground-penetrating [56] Manual RES2DINV, Geotomo Software, Penang, Malaysia, Feb. 2013.
radar data to characterize a gravel aquifer close to the Thur River, [57] J. Triantafilis, G. M. Laslett, and A. B. McBratney, “Calibrating an
Switzerland,” Near Surf. Geophys., vol. 8, no. 6, pp. 635–649, Dec. 2010. electromagnetic induction instrument to measure salinity in soil under
[35] A. Klotzsche, J. van der Kruk, G. A. Meles, and H. Vereecken, “Crosshole irrigated cotton,” Soil Sci. Soc. Amer. J., vol. 64, no. 3, pp. 1009–1017,
GPR full-waveform inversion of waveguides acting as preferential flow May 2000.
paths within aquifer systems,” Geophysics, vol. 77, no. 4, pp. H57–2H62, [58] K. A. Sudduth, S. T. Drummond, and N. R. Kitchen, “Accuracy issues in
Jul. 2012. electromagnetic sensing of soil electrical conductivity for precision agri-
[36] S. Lambot, E. C. Slob, I. van den Bosch, B. Stockbroeckx, B. Scheers, culture,” Comput. Electron. Agric., vol. 31, no. 3, pp. 239–264, May 2001.
and M. Vanclooster, “Estimating soil electric properties from monostatic [59] H. Abdu, D. A. Robinson, and S. B. Jones, “Comparing bulk soil elec-
ground-penetrating radar signal inversion in the frequency domain,” Water trical conductivity determination using the DUALEM-1S and EM38-DD
Resour. Res., vol. 40, p. W04 205, 2004. electromagnetic induction instruments,” Soil Sci. Soc. Amer. J., vol. 71,
[37] A. Kalogeropoulos, J. van der Kruk, J. Hugenschmidt, S. Busch, and no. 1, pp. 189–196, Jan. 2007.
K. Merz, “Chlorides and moisture assessment in concrete by GPR full [60] R. Gebbers, E. Lücke, M. Dabas, and H. Domsch, “Comparison of in-
waveform inversion,” Near Surf. Geophys., vol. 9, no. 3, pp. 277–285, struments for geoelectrical soil mapping at the field scale,” Near Surf.
Jun. 2011. Geophys., vol. 7, no. 3, pp. 179–190, Jun. 2009.
[38] E. Slob, M. Sato, and G. Olhoeft, “Surface and borehole ground- [61] F. Lavoué, J. van der Kruk, J. Rings, F. André, D. Moghadas,
penetrating-radar developments,” Geophysics, vol. 75, no. 5, pp. 75A103– L. A. Huisman, S. Lambot, L. Weihermüller, J. Vanderborght, and
75A120, Sep. 2010. H. Vereecken, “Electromagnetic induction calibration using apparent elec-
[39] J. van der Kruk, C. P. A. Wapenaar, J. T. Fokkema, and P. M. van den Berg, trical conductivity modelling based on electrical resistivity tomography,”
“Improved three-dimensional image reconstruction technique for multi- Near Surf. Geophys., vol. 8, no. 6, pp. 553–561, Dec. 2010.
component ground penetrating radar data,” Subsurf. Sens. Technol. Appl., [62] A. Mester, J. van der Kruk, E. Zimmerman, and H. Vereecken, “Quantita-
vol. 4, no. 1, pp. 61–99, Jan. 2003. tive two-layer conductivity inversion of multi-configuration electromag-
[40] K. Holliger, M. Musil, and H. Maurer, “Ray-based amplitude tomography netic induction measurements,” Vadose Zone J., vol. 10, no. 4, pp. 1319–
for crosshole georadar data: A numerical assessment,” J. Appl. Geophys., 1330, Nov. 2011.
vol. 47, no. 3/4, pp. 285–298, Jul. 2001. [63] O. Loeffler and M. Bano, “GPR measurements in a controlled vadose
[41] J. Tronicke and K. Holliger, “Effects of gas- and water-filled boreholes on zone: Influence of the water content,” Vadose Zone J., vol. 3, no. 4,
the amplitudes of crosshole georadar data as inferred from experimental pp. 1082–1092, Nov. 2004.
evidences,” Geophysics, vol. 69, no. 5, pp. 1255–1260, 2004. [64] Sensors & Software Inc., Subsurface Views, Jul. 2010. [Online]. Avail-
[42] R. Streich and J. van der Kruk, “Accurate imaging of multicomponent able: http://www.sensoft.ca/Resources/NewsletterArchive.aspx
GPR data based on exact radiation patterns,” IEEE Trans. Geosci. Remote [65] N. Engheta, C. H. Papas, and C. Elachi, “Radiation patterns of in-
Sens., vol. 45, no. 1, pp. 93–103, Jan. 2007. terfacial dipole antennas,” Radio Sci., vol. 17, no. 6, pp. 1557–1566,
[43] G. S. Smith, “Directive properties of antennas for transmission into a Nov./Dec. 1982.
material half-space,” IEEE Trans. Antennas Propag., vol. AP-32, no. 3, [66] E. Slob and J. Fokkema, “Coupling effects of two electric dipoles on an
pp. 232–246, Mar. 1984. interface,” Radio Sci., vol. 37, no. 5, p. 1073, 2002.
3958 IEEE TRANSACTIONS ON GEOSCIENCE AND REMOTE SENSING, VOL. 52, NO. 7, JULY 2014

Sebastian Busch received the M.Sc. degree in Harry Vereecken received the Eng. and M.Sc.
geosciences from the Ruhr-Universität Bochum, degrees in agricultural engineering and the Ph.D.
Bochum, Germany, in 2008. He is currently working degree in agricultural sciences from the Katholieke
toward the Ph.D. degree at the Institute of Bio- and Universiteit Leuven, Leuven, Belgium, 1982 and
Geosciences: Agrosphere (IBG-3), Forschungszen- 1988, respectively. His Ph.D. was on the develop-
trum Jülich GmbH, Jülich, Germany. ment of pedotransfer functions to estimate soil hy-
His current research focuses on quantitative draulic properties.
ground-penetrating radar (GPR) imaging using From 1988 to 1990, he was a Research Assistant
on-ground GPR full-waveform inversion and the on modeling nitrogen and water fluxes in soils and
coupled hydrogeophysical inversion of time-lapse groundwater. From 1990 to 1992, he was a Re-
surface GPR data to estimate the hydraulic properties searcher with the Institute of Petroleum and Organic
of a layered subsurface. Geochemistry, Forschungszentrum Jülich GmbH, Jülich, Germany, where he
became the Head of the Division “Behavior of Pollutants in Geological Sys-
tems” from 1992 to 2000 and has been the Director of the Institute of Bio- and
Jan van der Kruk (S’00–A’01–M’03) received the Geosciences: Agrosphere (IBG-3) since 2000. His current field of research is
M.Sc. degree in electrical engineering and the Ph.D. modeling of flow and transport processes in soils and hydrogeophysics.
degree in geophysics from the Delft University of
Technology, Delft, The Netherlands, in 1995 and
2001, respectively.
From 2001 to 2008, he was a Lecturer and a
Senior Researcher with the Applied and Environ-
mental Geophysics Group, Eidgenoessische Tech-
nische Hochschule Zurich, Zurich, Switzerland. His
research career has focused on the development of
imaging and inversion methodologies and their ap-
plications to solving applied problems. Since 2008, he has held a profes-
sorship position with Rheinisch-Westfälische Technische Hochschule Aachen
University, Aachen, Germany, and is currently a Research Group Leader
with Forschungszentrum Jülich GmbH, Jülich, Germany. His recent work has
focused on the inversion of dispersive ground-penetrating radar (GPR) waves
propagating in surface waveguides, GPR full-waveform inversion techniques,
and quantitative multiconfiguration inversion of electromagnetic induction data.
Dr. van der Kruk was the recipient of the Best Paper Award from the
Department of Applied Earth Sciences, Delft University of Technology, in 2000
and the Honorable Mention for Best Paper in Geophysics in 2006. In 2011, he
was the General Chair of the International Workshop on Advanced GPR in
Aachen. In 2012, he presented a Society of Exploration Geophysicists (SEG)
online virtual class about “Full-Waveform Inversion of Ground Penetrating
Radar Data,” and he was a Co-organizer of the SEG-American Geophysical
Union (AGU) Hydrogeophysics Workshop in Boise, ID, USA.

You might also like