Sei sulla pagina 1di 13

GEOPHYSICS, VOL. 80, NO. 2 (MARCH-APRIL 2015); P. G67G79, 12 FIGS.

10.1190/GEO2014-0194.1
Downloaded 03/23/15 to 187.175.244.100. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Structural joint inversion coupled with Euler deconvolution


of isolated gravity and magnetic anomalies

Emilia Fregoso1, Luis A. Gallardo2, and Juan Garca-Abdeslem2

match cross-model structures, only a few field examples have


ABSTRACT shown a noticeable improvement in resolution power (e.g., Doetsch
et al., 2010; Gallardo and Thebaud, 2012; Gallardo et al., 2012).
We generalized the Euler deconvolution method to a joint This is because, under the cross-gradient constraint, prevailing
scheme, which consists of locating the horizontal and ver- shapes of cross-model heterogeneities are jointly determined by the
tical positions of the top of potential-field 3D sources. These combined geophysical data at any given position, whereas the mag-
results were then used to constrain the depth to the top of the nitude of individual model properties and the magnitude of their
models obtained by cross-gradient joint 3D inversions, im- variations (i.e., model resolution) are not (e.g., Gallardo, 2007; Gal-
posing fixed known values in the a priori models. The cou- lardo et al., 2012). As in conventional, separate inversion, the res-
pling of both methods produced more realistic density and olution power depends only on the geophysical data resolution or
magnetization models for separate and joint inversions, rel- on direct measurements from borehole or surface exposures. It is
ative to those obtained by applying cross-gradient joint in- apparent that, in current algorithms, the model resolution has been
version only. This strategy was tested on a 3D synthetic increased by supplying direct property values or imposing alterna-
experiment, and on a real field data set from the northwest tive hypotheses such as a preferred homogeneity (smoothness) or
region of the Baja California Peninsula, Mexico. After locat- value-to-value crosscorrelations (e.g., Moorkamp et al., 2011; Le-
ing the vertical position of the source, the algorithm uses this lievre et al., 2012).
information to obtain density and magnetization models that It is well known that compared with other geophysical data, po-
enhanced their structural compatibility and reduces the am- tential-field data have limited resolution power, and therefore, data-
biguity on the interpretation of their structural characteristics independent restrictions are readily supplied for their inversion. In
laterally and at surface. this paper, instead of directly imposing model simplifications, we
assume that the application of data-driven regularizing constraints,
consistent with potential-field data, can help cross-gradient joint
geophysical inversion to obtain heterogeneous models with im-
INTRODUCTION proved resolution.
To our knowledge, cross-gradient joint inversion of potential-
The cross-gradient joint inversion strategy has been successfully field data alone, for 3D models, has only been attempted by Fregoso
applied to various combinations of geophysical data in the search and Gallardo (2009). In this paper, the authors conclude that jointly
for models with enhanced structural similarity (e.g., Gallardo and inverted models show better definition of structural features and
Meju, 2004; Linde et al., 2008; Hu et al., 2009; Doetsch et al., 2010; increased depth-resolution relative to separate inversions. However,
Moorkamp et al., 2011). When applied to field site data, the their models show the effects of the regularizing assumption of
resulting structurally similar models have generally facilitated the homogeneity, yielding underestimated density and magnetization
interpretation of geologically meaningful features of the subsurface values and little spurious heterogeneity at depth, inconsistent with
(e.g., Linde et al., 2008; Fregoso and Gallardo, 2009; Doetsch et al., their original test model.
2010; Gallardo and Thebaud, 2012; Gallardo et al., 2012). In an attempt to counterbalance the smoothness in the model, we
Despite most cross-gradient joint inversion algorithms having aim to introduce higher resolution a priori models derived by a strat-
consistently succeeded in accommodating model heterogeneities to egy known to provide redundant subsurface structures such as Euler

Manuscript received by the Editor 25 April 2014; revised manuscript received 18 November 2014; published online 4 February 2015.
1
Universidad de Guadalajara, CUCEI, Department of Mathematics, Guadalajara, Mexico. E-mail: emilia.fregoso@red.cucei.udg.mx.
2
CICESE, Earth Science Division, Ensenada, B.C., Mexico. E-mail: lgallard@cicese.mx; jgarcia@cicese.mx.
2015 Society of Exploration Geophysicists. All rights reserved.

G67
G68 Fregoso et al.

deconvolution (Thompson, 1982; Reid et al., 1990). The Euler de- is equal to zero in the whole mapped domain, while solving an ex-
convolution approach has been successfully applied to potential- panded least-squares optimization problem for two geophysical
field data to provide an ensemble of physically overlapping mono- models m1 and m2 simultaneously.
or dipolar point sources to outline the contact surfaces of actual In a close analogy to the pioneering formulation of Gallardo and
causative bodies (e.g., Thompson, 1982; Reid et al., 1990; El Dawi Meju (2003), Fregoso and Gallardo (2009) propose an iterative
Downloaded 03/23/15 to 187.175.244.100. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

et al., 2004; Beiki, 2010, 2013; Melo et al., 2013). In a combined scheme that solves this 3D optimization problem, under the frame-
approach, the surfaces determined by Euler deconvolution should work of the generalized least-squares formulation proposed by
also be acknowledged by cross-gradient joint inversion. There Tarantola and Valette (1982) by calculating
are recent variants of the Euler deconvolution method to ad-
equately obtain the source depth, horizontal position, and/or infor- ^ k1 m
m ^ k N1
1 n2
mation about the source shape (i.e., structural index [SI]) (e.g.,
Stavrev, 1997; Elawadi and Ushijima, 2002; Fedi et al., 2009; Florio N1 k T k 1 k T 1 k 1 ^ k ;
1 B B N1 B B N1 n2 gm m
and Fedi, 2014). (3)
Considering the model redundancy provided by Euler deconvo-
lution techniques, in our approach, we only select the most consis- where
tent structural information to provide a more detailed 3D a priori " #1
model suitable for cross-gradient joint inversion. A similar ap- AT1 C1 2 T 1
d1 A1 1 L1 L1 Cmapr 0
N1
1
1
proach that retrieves information from potential-field data analysis 0 AT2 C1 2 T 1
d2 A2 2 L2 L2 Cmapr
2
and then uses this information to constrain potential-field inversion
(4)
(self-constrained inversion) is proposed by Paoletti et al. (2013).
However, in our procedure, the a priori information, such as hori-
and
zontal and vertical position of the source, is obtained from a joint
Euler deconvolution scheme that involves the same gravity and " #
AT1 C1 ^ k1  21 LT1 L1 m
d1 d1 f 1 m ^ k1 C1
mapr
^ k1 mapr
m 1
magnetic data. After validating our strategy on a synthetic 3D ex- n2 1
:
periment, we evaluate its performance on a field-data example in AT2 C1 ^ k2  22 LT2 L2 m
d2 d2 f 2 m ^ k2 C1
mapr
^ k2 mapr
m 2
2

Sebastian Vizcaino Bay, Mexico. (5)

STRUCTURAL JOINT INVERSION AND EULER The two sets of model parameters after k iterations are m ^k
DECONVOLUTION m k T
^ 1 m k T T
^ 2  , A1 and A2 are the sensitivity matrices for each data
type, Bk is the Jacobian matrix for the cross-gradient function, and
Cross-gradient joint 3D inversion formulation
gm m^ k contains the three components of the cross-gradient vector.
In our coupled approach, we adopt the 3D cross-gradient joint Following Fregoso and Gallardo (2009), we compose our subsur-
inversion algorithm of Fregoso and Gallardo (2009). As described face model as an aggregate of rectangular prisms, and compute their
by Gallardo and Meju (2011), the minimizing functional behind gravity and magnetic responses using the equations developed by
this approach is given by Bhattacharyya (1966) and Bhaskara-Rao et al. (1990).

m1 ; m2 d1 f 1 m1 T C1
d1 d1 f 1 m1  Euler deconvolution with depth correction
T
d2 f 2 m2  C1
d2 d2 f 2 m2  Following the methodology described by Thompson (1982)
21 mT1 LT1 L1 m1 22 mT2 LT2 L2 m2 and Reid et al. (1990), we solve the 3D Eulers homogeneity equa-
tions as
m1 mapr T 1 apr
1  Cmapr m1 m1 
1
g g g
m2 mapr T 1 mapr x x0 y y0 z z0 g (6)
2  Cmapr m2 2 ; (1) x y z
2

where m1 and m2 are the two sets of model parameters, d1 and d2 for gravity data and
are the observed data, and f 1 m1 and f 2 m2 are the predicted data,
Cd1 and Cd2 correspond to the covariance matrices for each geo- T T T
x x0 y y0 z z0 NB T (7)
physical data set (which are assumed to be uncorrelated), L1 and x y z
L2 correspond to matrices that accommodate smoothness operators,
and 21 and 22 are weighting factors that regulate the level of for magnetic data, where x0 ; y0 are the horizontal coordinates of
smoothness required by the models. The terms mapr apr
1 and m2 are the source centroid and z0 is the depth to mono- or dipolar sources
a priori model parameters, and Cmapr and Cmapr are covariance ma- (assumed coincident to 3D source interfaces). The gradient compo-
1 2
trices for a priori model parameters. Central to this algorithm is the nents gx, gy, and gz and Tx, Ty, and Tz are
imposition of an exact structural resemblance constraint, as defined the first-order derivatives of the observed gravity g and magnetic T
by Gallardo and Meju (2011), requesting that the 3D cross-gradient anomalies, respectively. The constants and B are the regional val-
function ues of the gravity and magnetic total fields, respectively. Further-
more, the values and N are the SI for gravity and magnetics,
tx; y; z m1 x; y; z m2 x; y; z (2) respectively, which depend on the source geometry. A list of SIs
Coupled gravity and magnetic 3D models G69

 

 

 
 
for various geometries can be found in Reid et al. (1990), Elawadi g T  g T  g T 
and Ushijima (2002), and Saibi et al. (2006). Ai x x  y y  z z  N
xi ;yj xi ;yj xi ;yj
Closely following Reid et al. (1990) and Elawadi and Ushijima
(10)
(2002), we obtain common horizontal x0 ; y0 and depth z0 posi-
tions of the source as well as two SIs from gravity and magnetic data
Downloaded 03/23/15 to 187.175.244.100. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

simultaneously by applying the following strategies (Figure 1):


1) From gravity g and magnetic T anomalies, we calculate their re- for square moving windows, by assuming any SIs ; N.
spective gradient components. Horizontal derivatives gx, Therefore, A is a matrix with p p rows, for p
gy, Tx, and Ty are calculated in the space domain window size and five columns; and xi ; yj is a point in the win-
using differentiation, whereas vertical derivatives gy and dow, for i; j 1; 2; : : : ; p. For each window, the solution is
Tz are obtained using the fast Fourier transform technique. the five unknown components x x0 y0 z0 B T ob-
2) We assume that gravity and magnetic source locations are fully tained simultaneously from its gravity and magnetic data, and
coincident for each studied window, and we formulate a com- the components bi of vector b are given by
bined 3D homogeneity equation by adding equations 6 and 7 as
   
g T g T
x x0 y y0
x x y y
 
g T
z z0 g NB T;
z z
(8)

which can be rewritten as


   
g T g T
x y
x x 0 y y 0
 
g T
z NB
z z 0
   
g T g T
x y
x x y y
 
g T
z g NT: (9)
z z

It is noticeable that the summation of these


equations only makes sense if we are dealing
with nondimensional homogeneous fields.
In our case, the magnitude of the gravity and
magnetic data is normalized by their re-
spective maximum values, to make sense of
summation and meaningfully balance their
contributions. In our specific application, the
normalization is done for the whole data set,
assuming that there is only additive errors on
the data with zero mean and a constant stan-
dard deviation. This is fully equivalent to
assume that gravity and magnetic standard de-
viations are identical fractions of their maxi-
mum values and has the added advantage that
the resulting normalized gravity and magnetic
fields are also homogeneous (e.g., Blakely,
1995). Thus, they satisfy the Euler equation
as much as the original.
3) Applying singular value decomposition
(SVD), we solve the overdetermined system
of linear equation 9 given by Ax b, in
which the rows Ai of the matrix A are
given by Figure 1. Flowchart of our process for Euler deconvolution with depth correction.
G70 Fregoso et al.
    
g T  g T  points are selected as optimal. If no common intersection
b i xi y
x x xi ;yj j
y y xi ;yj is found, we select different values for x0 ; y 0 and recal-
   culate the intersecting lines.
g T 
0 gx ; y NTx i j :
; y 5) Based on these plots, we select adequate SIs and N, and we
z z xi ;yj i j
solve equation 9 for z0 to obtain a single depth value. Assuming
Downloaded 03/23/15 to 187.175.244.100. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

(11) that the source is below the subanomaly, we consider


x x0 y y0 0, B 0, and z 0 obtaining the fol-
The pseudoinverse of the matrix AT A in terms of the SVD of A lowing equation:
is expressed by
g NT

AT A V2 VT ; z0 . (17)
(12) gz Tz
where V is a 5 5 orthonormal matrix, and is the correspond-
ing singular value matrix. We find that the rank of the AT A ma- 6) The optimum depth value is obtained observing the statistical
trix is four, which can be explained by the rank of the linear distribution of the estimated depths and selecting the most fre-
systems in equations 6 and 7. Therefore, the nonnull singular quent value. The center of the bin in the histogram corresponds
values selected for the solution are 1 2 3 4 . to the discretization used in the cross-gradient joint inversion, in
As in Elawadi and Ushijima (2002), feasible solutions should the vertical direction. Similarly, the size of the bin corresponds
individually have small standard deviations and collectively to the blocks size.
group well. For this reason, we record the window solution
if the standard deviation of the estimated depth is less than a
specific percentage (e.g., 15%) of the depth itself. Cross-gradient joint 3D inversion with Euler
4) To estimate a single representative depth to the top of the source deconvolution constraint
and SIs for each lateral position, we track and select the asso-
The limited depth resolution of gravity and magnetic nonpara-
ciated subregions of the anomalies that resulted on close lateral
metric underdetermined problems is a major issue that has been dis-
x0 ; y0 positions from step 3 and apply the following method-
cussed by several authors (e.g., Fedi et al., 2005; Pilkington, 2012;
ology (Elawadi and Ushijima, 2002):
Paoletti et al., 2014). As described above, the cross-gradient con-
For each selected sector of gravity and magnetic data straint (that drives the joint 3D inversion of Fregoso and Gallardo,
(subanomalies), we reestimate depths for a range of SIs 2009) is independent of the properties magnitude or gradients and,
using the following equations: as such, does not solve the lack of depth resolution inherent in grav-
ity and magnetic data. To alleviate this difficulty, in this case, we
z z0 ag bg (13) assume that model heterogeneity occurs below the shallowest geo-
logic interface and that it is also the major source of the data gra-
for gravity data and dients used in Euler deconvolution. We thus incorporate the results
obtained by Euler deconvolution with the depth correction method-
z z0 am N bm (14) ology, in the density and magnetization models. This information is
incorporated by holding the parameters fixed, above the depth esti-
for magnetic data. In these expressions, ag , bg , am , and mated from Euler deconvolution. For this, we select small enough
bm are constant values given by values for the covariance matrices Cmapr and Cmapr (equations 35)
1 2
in their corresponding cells. It is worth mentioning that the compu-
g tational cost of our independent handling of Euler deconvolution is
ag g ;
z meaningless compared with the cost of a 3D cross-gradient joint
 
g g g inversion.
bg x x0 y y0
; (15)
x y z
TEST CASE
To test the proposed methodology, we consider a simple single-
T body model chosen to resemble the dipping dike used by Li and
am T ; and Oldenburg (1998). This dike is embedded in a homogeneous half-
z
  space (with zero density and induced magnetization contrasts), and
T T T we select a constant density and induce magnetization contrasts of
bm x x 0 y y 0 ; (16)
x y z 1 gcm3 and 1 Am, respectively (Figure 2).
The studied volume extends 1000 1000 m horizontally and
where x0 ; y 0 is a common central position applied for 500 m at depth and was discretized by 20 20 10 cubes with
all the selected subanomalies. 50-m sides. Within this volume, the dike is set to dip west from
The application of equations 13 and 14 results in several 50 m until it reaches 400-m depth. We simulate a 10-m-spaced grid
straight lines (one for each selected anomaly sector), and of gravity and magnetic data that covers exactly the studied volume.
because the z0 values must coincide for the whole ensem- Gaussian noise with standard deviation of 0.03 mGal and 2 nT
ble of gravity and magnetic subanomalies, intersection (approximately 1% of the maximum amplitude of their respective
Coupled gravity and magnetic 3D models G71

anomalies) was added to the data. For magnetic data, we consider a the largest estimated density values, within the volume of the
total magnetic intensity anomaly assuming a local geomagnetic test heterogeneity, are approximately 0.4 gcm3 (40% of the
field inclination I of 45 and declination D of 45. Figure 3 shows maximum values set in the test model). Certainly, these global
the resulting gravity and magnetic anomalies. measures indicate a marginal improvement on the average val-
ues of the jointly inverted models and an underestimation of the
Downloaded 03/23/15 to 187.175.244.100. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Separate and joint inversion unconstrained by Euler targets properties; however, they do not reflect the improve-
deconvolution ment on their spatial distribution.
3) Model resemblance In terms of the estimated distribution of
To evaluate the advantages of applying Euler deconvolution in subsurface density and magnetization, the unconstrained den-
advance to the full 3D cross-gradient joint inversion, we first per- sity model obtained after separate inversion (Figure 4a) shows
form separate and joint inversions with unconstrained a priori mod- a structure that, unlike the test heterogeneity, clusters near the
els, which are also used as starting models for the inversion process. surface. This is a well-known lateral effect of voxel-based in-
In our application, we choose 1 0.03 mGal and 2 2.0 nT, version of potential-field data. Alternatively, the unconstrained
correspondingly, to furnish Cd1 21 I and Cd2 22 I. The param- density model obtained after joint inversion (Figure 4d) adjusts
eters 21 and 22 are set to 21 1 1011 and 22 1 107 for grav- more according to the original test structure. Despite the im-
ity and magnetics, respectively, and they are chosen after several proved density distribution of Figure 4d over Figure 4a, we
experiments for separate inversions observing the trade-off between may observe that both distributions still spread down from
the level of smoothness and data misfit in both models. The same the surface rather than start at the expected depth as their mag-
values are used in the cross-gradient joint inversion procedure to netic counterparts (Figure 4b and 4e). A strategy to evaluate the
obtain models that are structurally compatible.
Then, we increase or decrease these regulariza-
tion values in logarithmic steps until satisfactory
data misfit and model convergence are achieved.
The values are chosen according to the order of
magnitude of the parameters. Covariance matri-
ces of a priori models Cmapr1
and Cmapr
2
are selected
large enough to pose fully unconstrained ex-
periments.
Once the appropriate processing parameters
are chosen, the inversion program is set to run
independently in separate and joint inversion
modes. Figure 4 shows selected vertical slices
of density and magnetization models obtained
for both unconstrained experiments. Beyond a
visual comparison, several aspects of the
resulting models are analyzed to prove their suit-
ability.
A first element to analyze is the suitability of
the models in terms of the actual geophysical
data. For this, first, we calculate the data misfit
using equations 30 and 31 from Fregoso and Gal- Figure 2. (a) Vertical and (b-d) horizontal cross sections of a 3D dipping dike-like struc-
lardo (2009). Second, we gauge the accuracy of ture (in red) embedded in a homogeneous medium (in blue) with null density and mag-
the models by calculating the proximity to the netization. Vertical cross section through the studied volume runs at north 500 m.
actual test model. Third, we analyze their Horizontal slices run at (b) 50-, (c) 200-, and (d) 350-m depth. The source body has
structural resemblance as measured by the cross- a density of 1.0 gcm3 and a magnetization of 1.0 Am.
gradient values:
1) Data misfits The normalized data misfit attained for these
models are rmsg 1.14 and rmsm 0.94 for separate inver-
sion. For joint inversion, the values attained are rmsg 1.72
and rmsm 1.35. For both cases, the data misfits validate
the geophysical suitability of both pairs of models. Further-
more, it also shows that the cross-gradient constraint only in-
creases slightly the data misfit.
2) Model accuracy As a global measure, we may compute
the rms difference between the estimated and test models for
the separate and joint inversion. This results in average
Figure 3. Noise-corrupted (a) gravity and (b) magnetic anomalies
differences of 0.20 gcm3 and 0.29 Am for separate inversion, produced by the dike of Figure 2. In both data sets, standard devi-
and 0.21 gcm3 and 0.32 Am for joint inversion. We also note ation of the added noise is 1% of the maximum amplitude of the
that, analogous to the findings of Fregoso and Gallardo (2009), anomaly.
G72 Fregoso et al.

structural resemblance among the two models is by using the vertical locations of the top of the anomalous source based on both
norm of the cross-gradient vector (see Fregoso and Gallardo, geophysical data simultaneously. To achieve some statistical signifi-
2009). Figure 4c shows major structural dissimilarities in the cance in our estimates, we first perform several deconvolution ex-
region of the dike, whereas joint inversion models improve periments using various window sizes and SIs on noise-free data,
not only in the dikes region but also in the entire domain, and we obtain the results shown in Figure 5.
Downloaded 03/23/15 to 187.175.244.100. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

as shown in Figure 4f. It is well known that a smaller window may target finer structures
suitable for more complex geometries at the cost of capturing data
noise (e.g., Figure 5a and 5b), whereas a larger window may prevent
Considering that, despite the improvement in cross-structure
the assimilation of such noise at the cost of oversimplifying the es-
matching, the cross-gradient joint inversion requires additional con-
timates of the subsurface bounding interfaces (e.g., Figure 5c and
straints that affect their physical properties more directly and
5d). It is noticeable that a larger window size is more suitable for our
improve further the model accuracy. Even though this can be sup-
particular test experiments given the simplicity of our model geom-
plied by direct observations independent of the original geophysical
etry, resulting in a selected source contour (delimited in black),
data, we propose that the same geophysical data can still provide
chosen empirically based on the cumulative points that define a
constraints of this type. In the following experiment, we supply
structure and suggest the horizontal source location. In this case,
the inversion with constraints estimated by the more conventional
this selected region coincides satisfactorily with the actual lateral
Euler deconvolution of the same geophysical data. For this, we
position of the top of the dike (delimited in red).
regularize the cross-gradient joint inversion by holding the
The SI values and N provide different configurations of the
subset of model parameters of the shallowest layers fixed, and ac-
commodating all density and magnetization heterogeneities within cumulative points. In this case, the selection of 0 and N 1
the lower layers. in conjunction with a 100 100 m moving window (Figure 5c)
provide not only the best shape index (top of a vertical structure
for the gravity and the magnetic source) for the actual dike-like
Preparation of a constrained model: Euler deconvolu-
heterogeneity, but it also better encloses the lateral extension of
tion strategy
the test dikes top.
Following the procedure of Euler deconvolution with depth As proposed by Elawadi and Ushijima (2002), to jointly support
correction described previously, we determine the horizontal and our selection of suitable SIs and obtain a more robust estimate of the

Figure 4. Vertical slices of (a and d) density and (b and e) magnetization models. The magnitude of cross-gradient vectors using equation 2 is
shown in panel (c) for separate inversion and in panel (f) for joint inversion. Top panels are after unconstrained separate inversion, and the
bottom panels are after unconstrained cross-gradient joint inversion. The true position of the dike is outlined by a black solid line.
Coupled gravity and magnetic 3D models G73

sources depth, we perform several experiments that should follow distribution of the shallowest estimated sources, and in general,
steps 4 through 6 of Euler deconvolution with depth correction. an adequate selection of the SIs yields a clear delimitation of the
For this, we select different test models showing the same dike-like horizontal and vertical positions of the source.
structure, but buried at different depths. Figure 6a and 6b shows the
intersecting lines for the case of our dike buried at a 50-m depth for Cross-gradient joint inversion constrained by Euler
Downloaded 03/23/15 to 187.175.244.100. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

the noise-free data. Observing these graphs, we select 0.5 and deconvolution
N 0.7, which are used in equation 17 to obtain the depth value z0 .
As shown in Figure 6c, the most recurrent depth is 50 m, which To gauge our full methodology of constrained cross-gradient
coincides with the actual depth to the top of the dike. Furthermore, joint inversion, we perform two comparative experiments using sep-
the SI values and N can be selected close to the values given by the arate and joint inversion algorithms. To assimilate the information
intersecting lines with satisfactory results, and according to the ex- provided by Euler deconvolution, we extend the density and mag-
pected values from the theory of the Euler deconvolution method, netization values observed at the surface (0 gcm3 and 0 Am) up
that preserve a difference of one between SI values when interpret- to 50-m depth, as suggested by our results on Euler deconvolution.
ing gravity and magnetic fields. Figure 6d shows the 50-m depth These parameter values are preset in the a priori models mapr 1 and
using 0 and N 1. It is worth noting that we select depth in- mapr
2 and held fixed by assigning small a priori covariance values in
tervals according to the vertical size of our model discretization. their corresponding places in the covariance matrices Cmapr 1
and
C apr . In this case, we impose density and magnetization real values
Figure 6e through 6h corresponds to the case of the dike buried at m2
a 200-m depth. In this experiment, we select 1.2 and N 1.6 at the surface. However, in field examples, we must select these
where the intersection of the lines occurs (Figure 6e and 6f). Using values from observations on the surface geology as well as from
these values for the SIs, we obtain the z0 value observing the most the results of other geophysical studies. The rest of the values of
recurrent value of 200 m (Figure 6g), which, again, coincides with the a priori models are taken from the unconstrained joint inversion
the test depth. Using 1.0 and N 2.0 (Figure 6h), we obtain a
model (see Figure 4d and 4e) but assigning larger covariance values,
different configuration of the frequency diagram but statistically the comparable with those imposed at surface. Beyond this adaptation,
real depth. the selection of covariance matrices for geophysical data sets and
Figure 6i through 6k corresponds to the estimates of the bottom smoothness parameters is similar to the case of separate and joint
layer of the dike structure shown in Figure 2. In this case, the in- inversion with no Euler deconvolution for a full comparison. The
tersection point occurs at 1.5 and N 2.5, and the most fre-
parameters 21 and 22 are set as 21 1 108 and 22 1 109 for
quent depth estimate occurs at 350 m and coincides with the test gravity and magnetics, respectively.
model.
We may note that even though the independent
plots of intersecting lines for gravity and mag-
netics show suitable depth estimations, we prefer
the most robust depth estimated from joint inver-
sions, which reduces the ambiguity by reconcil-
ing discrepancies of separated calculations. We
then use this information to restrict the cross-gra-
dient joint inversion to search only for deeper
heterogeneities.
In our more realistic framework, with noise-
added data (Figure 3), we obtain the results shown
in Figure 7. Applying the Euler deconvolution
with depth correction described above, we obtain
the horizontal location of the top of the dike using
SIs 0 and N 1 (red rectangle in Figure 7a)
and we select a suitable subregion to estimate the
depth value in the subsequent process (black rec-
tangle in Figure 7a). Similarly to the selection
of SI values for the noise-free data case, we use
0 and N 1 in equation 17. These values
are selected close to the corresponding values pro-
vided by the intersecting lines, shown in Figure 7b
and 7c, preserving a difference of one. Remark-
ably, we observe that the largest frequency for
z0 matches the depth of the true model at 50 m
(Figure 7d). Figure 5. Horizontal location of the top of the dike using SIs (a) 0, N 1;
In this experiment, we observe that Euler de- (b) 1, N 2; and window size of 3 3 in both cases. (c) SIs 0, N 1; (d)
1, N 2; and window size of 10 10 in both cases. A black solid line outlines
convolution with depth correction provides satis- the selected horizontal position of the top of the source. The real horizontal position
factory results for noise-corrupted data because of the dike is delimited in red. The intervals of depths are plotted in different colors
noise leads to an evident random horizontal as shown in panel (c).
G74 Fregoso et al.
Downloaded 03/23/15 to 187.175.244.100. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 6. Ensemble of linear relationships between z0 and SI for (a) gravity data and (b) magnetic data, for the case

of the dipping
dike located
at 50-m depth

(see Figure
2). The diagram of frequencies of z0 values jointly estimated using equation 17 for 0.5 and N 0.7
is shown
(c) and 0 and N 1.0 is shown in (d). Panels (e-h) show a minor dike model, located at 200-m depth using SI values
in
1.2 and
N 1.6 in panel (g) and 1.0 and N 2.0 in panel (h). Panels (i-k) show the bottom block of the dike at 350 m, selecting 1.5 and
N 2.5.

Sections of constrained separate inversion


models are shown in Figure 8a and 8b, and those
for constrained joint inversion are shown in Fig-
ure 8d and 8e. Following our comparison to the
previous unconstrained experiments, we may
comment on the following aspects:

1) Data misfits It is remarkable that despite


the differences in the shapes defined by the
models, the data misfits obtained for both
experiments are as suitable as those obtained
in their corresponding unconstrained experi-
ments with rmsg 1.0 and rmsm 1.09 for
separate inversion, and rmsg 1.0, rmsm
1.11 for joint inversion.
2) Model accuracy In separate and joint
inversion models, we may remark that the
estimated property values are larger, i.e.,
closer to the values set in the original test
model. The average rms differences are
Figure 7. (a) Horizontal position of the sources estimated from noise-corrupted data 0.19 gcm3 and 0.25 Am for constrained
using SIs 0 and N 1.0. Linear relationships between z0 and SIs for (b) gravity separate inversion, and 0.17 gcm3 and
and (c) magnetic data. (d) Diagram of frequencies of z0 for 0 and N 1.0. The 0.25 Am for constrained joint inversion.
black solid line in panel (a) outlines the subregion selected to estimate the depth to
the top of the source, whereas the red solid line outlines the actual horizontal position These values are smaller than the average
of the dike. Estimated source positions for different depth intervals are shown in differ- rms obtained for separate and joint inversion
ent colors. experiments unconstrained by Euler decon-
Coupled gravity and magnetic 3D models G75

volution (Figure 4). In both cases, the sole use of Euler decon- suggest that the shape of structures depicted in constrained inver-
volution-derived constraint does not suffice to improve model sion actually better resembles the test structure. A detailed in-
resolution. However, reducing the solution space of the esti- spection of the actual values clearly marks the superiority of the
mated models by imposing constraints on the surface layers new constrained joint estimates that are correspondingly distrib-
through the assigned covariance values of the a priori models uted and closer to the actual test values. These results suggest
Downloaded 03/23/15 to 187.175.244.100. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

favors significantly the resolution of the estimated models (e.g., that the combined constraints produced the desired effects in
Figures 8d versus 4d). This provides similar joint inversion the estimated models. Moreover, we can observe that although
models, which can be observed by the structural characteristics the magnetic models obtained after unconstrained inversion
and by the same model accuracies. Figure 8a (after depth con- (Figure 4b and 4e) located at the top of the dike and the top layer
strained separate inversion) shows that the sole use of depth satisfactorily, providing relatively homogeneous magnetization
constraint prevented the accumulation of mass near the surface. values (close to zero), they provided no structure to be propa-
Figure 8b shows that magnetization was not largely favored by gated into the density models for this same zone. However, the
this constraint because the outcropping slab was already well corresponding constrained inversion models (Figure 8b and 8e)
resolved in the unconstrained inversion (Figure 4b). imposed a homogeneity layer that led to an improved density
3) Model resemblance The strategy of fixing parameter values models and allowed the structural resemblance to focus on the
has favored the cross-gradient constraint and thus our con- structure-rich part of the model. Additionally, we can observe that
strained joint inversion, which lead to noticeable improvements the separate inversion models show enhanced structural resem-
in the distribution of the density and magnetization hetero- blance (Figure 8c) comparable with Figure 4c, and the norm
geneities (Figure 8d and 8e). We may note that this improve- of the cross-gradient vector, obtained from jointly inverted mod-
ment was limited because the gravity model was somehow els, shows an improvement of this characteristic, as shown in
restrained to the magnetic model. In this case, we select a cutoff Figure 8f.
level of singular value that preserves the structural similarity
(see Fregoso and Gallardo, 2009) at the same level as the joint
inversion experiment unconstrained by Euler deconvolution, for FIELD EXAMPLE
which the gravity model resulted coupled to the magnetic model.
A structurewise comparison of unconstrained and constrained To fully test the methodology in actual field data, we select an
joint inversion (Figure 4d and 4e versus Figure 8d and 8e) may area located in Sebastian Vizcaino Bay, Mexico (282820N, 115

Figure 8. Vertical slices of (a and d) density and (b and e) magnetization models. (c) The magnitude of cross-gradient vectors using equation 2
for constrained separate inversion, and (f) for constrained joint inversion. Panels (a-c) result after Euler-constrained separate inversion and
panels (d-f) after Euler-constrained cross-gradient joint inversion. The true position of the dike is outlined by a solid line. Note that the Euler-
determined 50-m top layer was held fixed during both inversions.
G76 Fregoso et al.

1411450W) (Figure 9a). A geologic map of the study area is turbidite sequence of Cretaceous age and a Miocene-Pliocene silt-
shown in Figure 9b. stone sequence that discordantly overlaid the upper-plate Meso-
zoic rocks.
Geology of the area Toward the east, along the Baja California Peninsula, there is a
suite of volcano-plutonic rocks that related with the subduction of
The Vizcaino Peninsula and Cedros Island are composed of a
Downloaded 03/23/15 to 187.175.244.100. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

oceanic lithosphere into the underlying mantle. This region, defined


suite of Triassic to Cretaceous rocks, typical of oceanic crust, that by Sedlock et al. (1993) as Yuma terrane, includes Paleozoic-
are defined by Sedlock et al. (1993) as the Cochim terrane. Accord- Mesozoic metamorphic rocks, as well as Cretaceous rocks typical
ing to these authors, this terrane was accreted to western North of a volcanic arc. The Yuma terrane was intruded by granitoids in
America approximately 140 Ma ago, in early Cretaceous times. the late Cretaceous (100 Ma), and these granitoids are exposed
The compressive margin built by the Cochim terrane can be inter- from the eastern portion of the study area to the Gulf of California.
preted as an upper plate, constituted by Triassic-Jurassic ophiolitic Relevant to the present study is the serpentine-matrix mlange.
rocks, overriding a lower plate consisting of a subduction complex, The reactions during the process of serpentinization of ultramafic
deformed and metamorphosed to blueschist facies conditions, ap- rocks yield a large amount of magnetite that may produce strong
proximately 115104 Ma ago. Both plates are structurally separated magnetic anomalies, and the density may change from 3300 to
by an intervening serpentinite-matrix mlange containing a variety 2700 kgm3 with a concurrent increase in volume of approximately
of mafic and ultramafic rocks as well as exotic blocks of blueschists, 30%40%, causing a decrease in the magnitude of the gravity
eclogite, and amphibolite. It had been postulated that extensional anomalies. These geophysical observables characterize the gravity
deformation of Mesozoic rocks and blueschist exhumation occurred and magnetic anomalies as shown in Figure 10a and 10b.
during synsubduction extension of the North American forearc
between 95 and 2030 Ma (Baldwin, 1996; Sedlock, 1996). Uncon- Geophysical data of the area
formably overlaying the upper-plate rocks, there is a siliciclastic
The gravity, magnetic, and bathymetric data
used in this study are part of a compilation for
northwest Mexico. On land, this compilation in-
cludes proprietary survey data and data obtained
from INEGI (2008). At sea, this compilation in-
cludes gravity, magnetic and bathymetric data
from the western continental margin of Mexico
(Secretara de Marina, 1987), free-air gravity
data derived from satellite altimetry (Andersen
et al., 2010), magnetic data from the digital
magnetic map of North America (North Ameri-
can Magnetic Anomaly Group) (Bankey et al.,
2002), bathymetry from the GEBCO (2009),
and topography from CGIAR-CSI (2006).
For this compilation, a complete Bouguer
anomaly was calculated using SRTM data up
to 100 km surrounding the study area, using
Figure 9. (a) Location of the Sebastin Vizcano Bay in the Baja California Peninsula, variable density 2 gcm3 for the regions blan-
Mexico. (b) Geologic map that outlines the studied region (blue square) in Sebastin keted with sediments and soil and a density of
Vizcano Bay. The red dashed line in panel (b) represents the boundary between the
Yuma and Cochim terrains as proposed by Sedlock et al. (1993). Qs = Quaternary sedi- 2.67 gcm3 for the volcanic, metamorphic, and
ments, Qv = Quaternary volcanics, Ts = Tertiary sedimentary, Mivs = Miocene volcano plutonic terrains, accordingly with the digital geo-
sedimentary, and PMmu = Paleozoic-Mesozoic metamorphic undifferentiated. logic chart of Mexico of INEGI. The Bouguer
anomaly at sea was obtained assuming seawater
density of 1.03 gcm3 and adding the gravity ef-
fect caused by a body constrained using the bathy-
metric data from the GEBCO, with a density of
0.97 gcm3 , to simulate that density across the
sea-land interface has a density of 2 gcm3 .

Analytic continuation and induced


magnetization of the area
The gravity and magnetic data were analyti-
cally continued upward to homogenize the data.
The rationale for this practice follows an analysis
Figure 10. (a) Bouguer gravity and (b) total magnetic intensity anomalies for the studied made by Bhattacharyya (1980) that has been fol-
region (blue rectangle in Figure 9b) in UTM coordinates. The black dotted line indicates lowed by other authors (e.g., Li and Oldenburg,
the boundary of the Cedros Island. 1996). Bhattacharyya (1980) points out that the
Coupled gravity and magnetic 3D models G77

height of the observational surface above the source region pri- Cross-gradient joint inversion constrained by Euler
marily determines the critical dimension of the smallest inhomoge- deconvolution
neity in magnetization that can be resolved from the data.
Furthermore, when the horizontal dimensions of a rectangular block Once we estimate this depth, we fix the upper layers of the region
are smaller in size than this critical dimension, the block appears from surface to 3-km depth, around the source delimited by the
black rectangle in Figure 11a. In this case, we consider fixed values
Downloaded 03/23/15 to 187.175.244.100. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

homogeneously magnetized in the observed magnetic field. There-


fore, considering the size of the rectangular prism used in the in- of density contrast and magnetization from 390 to 417 km in the
verse modeling, the gravity and magnetic data were analytically east direction, and from 3115 to 3137.5 km in the north direction.
continued upward 1 km. The a priori parameter values were obtained from previous knowl-
Following commonly used assumptions and the present limita- edge on regional geology as follows: (1) For density, we assign a
tions of the algorithm, we assume that the magnetic anomalies linear depth-decreasing density contrast ranging from 0.6 gcm3
are due to a magnetization parallel to the current geomagnetic field on surface level up to 0.2 gcm3 at 3-km depth. (This accounts for
only and neglect any perpendicular component of remanent mag- a superficial water layer filled with gradually compacted sediments
netization. We bear in mind that this component of the remanency at depth.) (2) For magnetization, the corresponding values were
assumption may be present in the case of magnetite generated simply set to 0 Am assuming that no magnetic anomaly originates
during serpentinization because this event may have occurred from either sediment or water layers. The a priori parameter
during Middle Miocene (approximately 12 Ma ago) at a time values at the bottom layer are set as the mean value of the resulting
when the geomagnetic polarity changed frequently, and the Farallon models from unconstrained joint inversion, assigning 0 gcm3 and
plate was being consumed under North America. The serpen- 0.5 Am for density and magnetization models, respectively. The
tinization process may possibly be related with the opening of model was discretized using 26 19 10 rectangular blocks of
an asthenospheric window on the underlying Farallon plate, due 1.5 km wide, 1.5 km long, and variable widths in the vertical di-
to a slab tearing following a ridge-trench collision (Pallares et al., rection (1 km thick from surface to a 4-km depth, 2 km thick from
2007). a 4- to 12-km depth, and 3 km thick from a 12- to 18-km depth). The
assumed direction of the inducing magnetic field was D 11 and
Euler deconvolution I 54. In this case, the smoothness factors were set as 21 1
103 and 22 1 102 for gravity and magnetics, respectively.
The results of Euler deconvolution of the joint gravity and mag- The data misfit for gravity is rmsg 1.73 and rmsm 1.86 for
netic data are shown in Figure 11a. We can observe that the major magnetic data. Vertical sections (Figure 12a and 12b) show the
dispersion of the sources corresponds to the shallowest depths, parameter distribution that satisfies the imposed surface and bottom
whereas the clustered position of the sources, which delineates conditions, whereas horizontal sections of the 3D models show
the largest magnetic anomaly, corresponds to depths from 3 km. the distribution of the structure of the main body that produces the
To obtain these results, we follow steps 1 through 3 of the Euler
anomalies at 3.5-(Figure 12c and 12d), 7.0-(Figure 12e and 12f) and
deconvolution with depth correction strategy described above.
11.0-km depth (Figure 12g and 12h).
We use the initial SI values N 0.5 and a
moving window size of 10 10 that implies
horizontal distances approximately x 7.2 km
and y 5.4 km.
The results of our experiments to determine
the depth to the top of the source are shown in
Figure 11b11d and point to a depth of approx-
imately 3 km. To obtain these results, we use the
normalized gravity and magnetic anomalies of
the subregion delimited in black in Figure 11a.
Following equations 1316 and selecting a cen-
tral point x0 ; y0 in the subregion, we obtain the
straight lines shown in Figure 11b and 11c that
provide the adequate SIs and their corresponding
depth values. In this case, we select 0 and
N 0.5, and the obtained depth is approxi-
mately 5 km. These differences in SIs for gravity
and magnetics are smaller than one unit. By se-
lecting these values, we may argue that the meth-
od shows some robustness when handling the
geometric complexity of a source that departs
significantly from the single pole-dipole forms.
Using these SIs and applying steps five and six Figure 11. (a) Map of the source positions determined by joint Euler deconvolution
of the Euler deconvolution with depth correction showing the depth intervals in different colors. The red solid lines outline the studied
area whereas the black solid lines bound the estimated sources. (b and c) Ensemble of
strategy, we obtain the depth histogram shown in linear relationships between z0 and SI for various sectors of (b) gravity and (c) magnetic
Figure 11d and determine a suitable depth to the data. (d) Diagram of frequencies that provides the z0 value obtained jointly using equa-
source of 3 km. tion 17 with 0 and N 0.5.
G78 Fregoso et al.

Model interpretation by an approximately 12-km-thick ophiolite suit. At the bottom


of the model (Figure 12g and 12h) is a region with negative
The result of the inverse modeling is consistent with the expected density contrast that is interpreted as the lower plate of the Cochim
consequences of serpentinization and with the accretion of the terrane, which possibly includes poorly consolidated sediments,
Cochim terrane over western North America, as proposed by Sed- accumulated over the continental margin in a graben developed
lock (1996). Due to the serpentinization, we expect the coexistence
Downloaded 03/23/15 to 187.175.244.100. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

on the forearc region prior to the accretion over the continental


of low-density and highly magnetized rocks. This scenery is clearly margin.
observed on the eastern down-portion of the model (Figure 12a and
12b) as well as on the plan views shown at 3.5, 7.0, and 11.0 km of
depth (Figure 12c12h). Over the eastern margin of the study area, a CONCLUSIONS
prominent feature in the model is a positive density feature that ex- We developed a methodology to locate the horizontal and vertical
tends from 3.0 to 12.0 km of depth (Figure 12c12h), which we position to the top of isolated sources by using a joint strategy of
interpret as the upper plate of the Cochim terrane that is constituted Euler deconvolution for gravity and magnetic data. Given that the
joint Euler deconvolution strategy does not in-
volve additional weighting functions nor addi-
tional control parameters in the joint inversion
process, its implementation is a low-cost re-
source. This information is used to constrain
the surface layers of 3D density and magnetiza-
tion models in a cross-gradient joint 3D inversion
process and tested on synthetic and real field
data. Incorporating information on the top of the
models is essential to recover the models at
depth.
The results of the synthetic experiment show
that the estimated source is better located and that
their property values are closer to the actual test
model when the cross-gradient joint inversion is
constrained by using minimum depths derived
from joint Euler deconvolution.
In the field example, the cross-gradient joint
inversion constrained by depths estimated from
the joint Euler deconvolution, led to density and
magnetization models with realistic property val-
ues with a common and geologically meaningful
distribution.
We may thus conclude that all our experiments
suggest that the combination of both methodol-
ogies reduces the ambiguity in depth resolution
that occurs when potential-field data are jointly
inverted.

ACKNOWLEDGMENTS
The authors acknowledge CONACYT for the
financial support to develop this work through
project no. CB-2012-01-177655. We appreciate
the constructive comments and suggestions made
by assistant editor, J. Shragge, and reviewers V.
Paoletti, R. Pasteka, and F. Caratori Tontini,
which improved the readability of the manuscript
as well as enhanced the methodology and its ap-
plication to the field data.
Figure 12. Vertical cross sections of (a) density and (b) magnetization models at north =
3125 km. Note the fixed values from the surface up to a 3-km depth, which were derived
from our joint Euler deconvolution strategy (Figure 1). Note the length of each block in REFERENCES
the vertical direction, represented by equidistant blocks. Horizontal sections of (c) den-
sity and (d) magnetization at a 3.5-km depth of the studied region. Correspondingly, Andersen, O. B., P. Knudsen, and P. A. M. Berry, 2010,
horizontal sections (e and f) at a 7.0-km depth and (g and h) at a 11.0-km depth. The DNSC08GRA global marine gravity field from
All figures show the inverted models from the anomalies in Figure 10. The black dotted double retracked satellite altimetry: Journal of Geod-
line indicates the boundary of the Cedros Island as a reference. esy, 84, 191199, doi: 10.1007/s00190-009-0355-9.
Coupled gravity and magnetic 3D models G79

Baldwin, S. L., 1996, Contrasting P-T-t histories for blueschists from the Gallardo, L. A., and N. Thebaud, 2012, New insights into Archean granite-
western Baja terrane and Aegean: Effects of synsubduction exhumation greenstone architecture through joint gravity and magnetic inversion:
and backarc extension, in G. E. Bebout, D. Scholl, S. Kirby, and J. P. Platt, Geology, 40, 215218, doi: 10.1130/G32817.1.
eds., Subduction top to bottom: American Geophysical Union Geophysi- General Bathymetric Chart of the Oceans (GEBCO), 2009, GEBCO_08
cal Monographs 96, 135141, doi: 10.1029/GM096p0135. grid, http://www.gebco.net/, accessed 19 January 2009.
Bankey, V., A. Cuevas, D. Daniels, C. A. Finn, I. Hernandez, P. Hill, R. Hu, W. Y., A. Abubakar, and T. M. Habashy, 2009, Joint electromagnetic
Kucks, W. Miles, M. Pilkington, C. Roberts, W. Roest, V. Rystrom, S. and seismic inversion using structural constraints: Geophysics, 74, no. 6,
Downloaded 03/23/15 to 187.175.244.100. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Shearer, S. Snyder, R. Sweeney, J. Velez, J. D. Phillips, and D. Ravat, R99R109, doi: 10.1190/1.3246586.
NAMAG (North American Magnetic Anomaly Group), 2002, Digital data Instituto Nacional de Estadstica y Geografa (INEGI), 2008, Modelos y Car-
grids for the magnetic anomaly map of North America, http://pubs.usgs tas Gravimtricas, http://www.gebco.net/, accessed 26 June 2008.
.gov/of/2002/ofr-02-414/, accessed 29 January 2004. Lelievre, P. G., C. G. Farquharson, and C. A. Hurich, 2012, Joint inversion
Beiki, M., 2010, Analytic signals of gravity gradient tensor and their appli- of seismic traveltimes and gravity data on unstructured grids with appli-
cation to estimate source location: Geophysics, 75, no. 6, I59I74, doi: 10 cation to mineral exploration: Geophysics, 77, no. 1, K1K15, doi: 10
.1190/1.3493639. .1190/geo2011-0154.1.
Beiki, M., 2013, TSVD analysis of Euler deconvolution to improve estimat- Li, Y., and D. W. Oldenburg, 1996, 3-D inversion of magnetic data: Geo-
ing magnetic source parameters: An example from the sele area, Swe- physics, 61, 394408, doi: 10.1190/1.1443968.
den: Journal of Applied Geophysics, 90, 8291, doi: 10.1016/j.jappgeo Li, Y., and D. W. Oldenburg, 1998, 3-D inversion of gravity data: Geophys-
.2013.01.002. ics, 63, 109119, doi: 10.1190/1.1444302.
Bhaskara-Rao, D., M. J. Prakash, and N. Ramesh-Babu, 1990, 3D and 2 D Linde, N., A. Tryggvason, J. E. Peterson, and S. S. Hubbard, 2008, Joint
modeling of gravity anomalies with variable density contrast: Geophysi- inversion of crosshole radar and seismic traveltimes acquired at the South
cal Prospecting, 38, 411422, doi: 10.1111/j.1365-2478.1990.tb01854.x. Oyster bacterial transport site: Geophysics, 73, no. 4, G29G37, doi: 10
Bhattacharyya, B. K., 1966, A method for computing the total magnetization .1190/1.2937467.
vector and the dimensions of a rectangular block-shaped body from mag- Melo, F. F., V. C. F. Barbosa, L. Uieda, V. C. Oliveira, Jr., and J. B. C. Silva,
netic anomalies: Geophysics, 31, 7496, doi: 10.1190/1.1439765. 2013, Estimating the nature and the horizontal and vertical positions of 3D
Bhattacharyya, B. K., 1980, A generalized multibody model for inversion of magnetic sources using Euler deconvolution: Geophysics, 78, no. 6, J87
magnetic anomalies: Geophysics, 45, 255270, doi: 10.1190/1.1441081. J98, doi: 10.1190/geo2012-0515.1.
Blakely, R. J., 1995, Potential theory in gravity and magnetic applications: Moorkamp, M., B. Heincke, M. Jegen, A. W. Roberts, and R. W. Hobbs,
Cambridge University Press. 2011, A framework for 3-D joint inversion of MT, gravity and seismic
Consortium for Spatial Information (CGIAR-CSI), 2006, SRTM 90 m dig- refraction data: Geophysical Journal International, 184, 477493, doi:
ital elevation data, http://www.cgiar-csi.org/, accessed 20 April 2006. 10.1111/j.1365-246X.2010.04856.x.
Doetsch, J., N. Linde, I. Coscia, S. A. Greenhalgh, and A. G. Green, 2010, Pallares, C., R. C. Maury, H. Bellon, J-Y. Royer, T. Calmus, A. Aguilln-Ro-
Zonation for 3D aquifer characterization based on joint inversions of mul- bles, J. Cotten, M. Benoit, F. Michaud, and J. Bourgois, 2007, Slab-tearing
timethod crosshole geophysical data: Geophysics, 75, no. 6, G53G64, following ridge-trench collision: Evidence from Miocene volcanism in Baja
doi: 10.1190/1.3496476. California, Mxico: Journal of Volcanology and Geothermal Research, 161,
Elawadi, E., and K. Ushijima, 2002, Interpretation of isolated gravity 95117, doi: 10.1016/j.jvolgeores.2006.11.002.
anomalies using Euler deconvolution: 64th Annual International Confer- Paoletti, V., P. C. Hansen, M. F. Hansen, and M. Fedi, 2014, A computa-
ence and Exhibition, EAGE, Extended Abstracts, P114. tionally efficient tool for assessing the depth resolution in large-scale po-
El Dawi, M. G., L. Tianyou, S. Hui, and L. Dapeng, 2004, Depth estimation of tential-field inversion: Geophysics, 79, no. 4, A33A38, doi: 10.1190/
2-D magnetic anomalous sources by using Euler deconvolution method: geo2014-0017.1.
American Journal of Applied Sciences, 1, 209214., doi: 10.3844/ajassp Paoletti, V., S. Ialongo, G. Florio, M. Fedi, and F. Cella, 2013, Self-
.2004.209.214. constrained inversion of potential fields: Geophysical Journal Interna-
Fedi, M., G. Florio, and T. A. M. Quarta, 2009, Multiridge analysis of po- tional, 195, 854869, doi: 10.1093/gji/ggt313.
tential fields: Geometric method and reduced Euler deconvolution: Geo- Pilkington, M., 2012, Analysis of gravity gradiometer inverse problems us-
physics, 74, no. 4, L53L65, doi: 10.1190/1.3142722. ing optimal design measures: Geophysics, 77, no. 2, G25G31, doi: 10
Fedi, M., P. C. Hansen, and V. Paoletti, 2005, Tutorial: Analysis of depth .1190/geo2011-0317.1.
resolution in potential-field inversion: Geophysics, 70, no. 6, A1A11, Reid, A. B., J. M. Allsop, H. Granser, A. J. Millett, and I. W. Somerton,
doi: 10.1190/1.2122408. 1990, Magnetic interpretation in three dimensions using Euler deconvo-
Florio, G., and M. Fedi, 2014, Multiridge Euler deconvolution: Geophysical lution: Geophysics, 55, 8091, doi: 10.1190/1.1442774.
Prospecting, 62, 333351, doi: 10.1111/1365-2478.12078. Saibi, H., J. Nishijima, E. Aboud, and S. Ehara, 2006, Euler deconvolution
Fregoso, E., and L. A. Gallardo, 2009, Cross-gradients joint 3D inversion of gravity data in geothermal reconnaissance; The Obama geothermal
with applications to gravity and magnetic data: Geophysics, 74, no. 4, area, Japan: Butsuri-Tansa, 59, 275282, doi: 10.3124/segj.59.275.
L31L42, doi: 10.1190/1.3119263. Sedlock, R. L., 1996, Syn-subduction forearc extension and blueschist ex-
Gallardo, L. A., 2007, Multiple cross-gradient joint inversion for geospectral humation in Baja California, Mxico, in G. E. Bebout, D. W. Scholl, S. H.
imaging: Geophysical Research Letters, 34, L19301, doi: 10.1029/ Kirby, and J. P. Platt, eds., Subduction top to bottom: American Geophysi-
2007GL030409. cal Union Geophysical Monographs 96, 155162.
Gallardo, L. A., S. L. Fontes, M. A. Meju, M. P. Buonora, and P. P. de Lugao, Sedlock, R. L., F. Ortega-Gutirrez, and R. C. Speed, 1993, Tectonostrati-
2012, Robust geophysical integration through structure-coupled joint in- graphic terranes and tectonic evolution of Mxico: Geological Society of
version and multispectral fusion of seismic reflection, magnetotelluric, America Special Papers 278.
magnetic, and gravity images: Example from Santos Basin, offshore Bra- Stavrev, P. Y., 1997, Euler deconvolution using differential similarity trans-
zil: Geophysics, 77, no. 5, B237B251, doi: 10.1190/geo2011-0394.1. formations of gravity or magnetic anomalies: Geophysical Prospecting,
Gallardo, L. A., and M. A. Meju, 2003, Characterization of heterogeneous 45, 207246, doi: 10.1046/j.1365-2478.1997.00331.x.
near-surface materials by joint 2D inversion of DC resistivity and Secretara de Marina, 1987, Atlas/Memoria del levantamiento geofsico de la
seismic data: Geophysical Research Letters, 30, 16581661, 10.1029/ zona econmica exclusiva y margen continental oeste de Mxico: report
2003GL017370. SMPO8710.
Gallardo, L. A., and M. A. Meju, 2004, Joint twodimensional DC resistivity Tarantola, A., and B. Valette, 1982, Generalized non-linear inverse problems
and seismic travel time inversion with crossgradients constraints: Journal solved using the least-squares criterion: Reviews of Geophysics and
of Geophysical Research, 109, B03311, doi: 10.1029/2003JB002716. Space Physics, 20, 219232, doi: 10.1029/RG020i002p00219.
Gallardo, L. A., and M. A. Meju, 2011, Structure-coupled multiphysics im- Thompson, D. T., 1982, EULDPH: A new technique for making computer-
aging in geophysical sciences: Reviews of Geophysics, 49, RG1003, doi: assisted depth estimates from magnetic data: Geophysics, 47, 3137, doi:
10.1029/2010RG000330. 10.1190/1.1441278.

Potrebbero piacerti anche