You are on page 1of 15

Tectonophysics 609 (2013) 82–96

Contents lists available at ScienceDirect

Tectonophysics
journal homepage: www.elsevier.com/locate/tecto

Review Article

Models of crustal thickness for South America from seismic refraction,


receiver functions and surface wave tomography
Marcelo Assumpção a,⁎, Mei Feng b, Andrés Tassara c, Jordi Julià d
a
Institute of Astronomy, Geophysics and Atmospheric Sciences, University of São Paulo, Rua do Matão 1226, 05508-090, São Paulo, SP, Brazil
b
Institute of Geomechanics, Chinese Academy of Geological Sciences, Beijing, 100081, China
c
Department of Earth Sciences, Universidad de Concepción, Victor Lamas 1290, Concepción, Chile
d
Departamento de Geofísica and Programa de Pós-Graduação em Geodinâmica e Geofísica, Universidade Federal do Rio Grande do Norte, Natal, 59078-970 RN, Brazil

a r t i c l e i n f o a b s t r a c t

Article history: An extensive compilation of crustal thicknesses is used to develop crustal models in continental South
Received 22 April 2012 America. We consider point crustal thicknesses from seismic refraction experiments, receiver function anal-
Received in revised form 5 November 2012 yses, and surface-wave dispersion. Estimates of crustal thickness derived from gravity anomalies were only
Accepted 15 November 2012
included along the continental shelf and in some areas of the Andes to fill large gaps in seismic coverage.
Available online 28 November 2012
Two crustal models were developed: A) by simple interpolation of the point estimates, and B) our preferred
Keywords:
model, based on the same point estimates, interpolated with surface-wave tomography. Despite gaps in con-
Crust tinental coverage, both models reveal interesting crustal thickness variations. In the Andean range, the crust
Moho reaches 75 km in Southern Peru and the Bolivian Altiplano, while crustal thicknesses seem to be close to the
Andes global continental average (~ 40 km) in Ecuador and southern Colombia (despite high elevations), and along
Tomography the southern Andes of Chile–Argentina (elevation lower than 2000 m). In the stable continental platform the
average thickness is 38 ± 5 km (1-st. deviation) and no systematic differences are observed among Archean–
Paleoproterozoic cratons, NeoProterozoic fold belts, and low-altitude intracratonic sedimentary basins. An
exception is the Borborema Province (NE Brazil) with crust ~30–35 km thick. Narrow belts surrounding
the cratons are suggested in central Brazil, parallel to the eastern and southern border of the Amazon craton,
and possibly along the TransBrasiliano Lineament continuing into the Chaco basin, where crust thinner than
35 km is observed. In the sub-Andean region, between the mid-plate cratons and the Andean cordillera, the
crust tends to be thinner (~35 km) than the average crust in the stable platform, a feature possibly inherited
from the old pre-Cambrian history of the continent. We expect that these crustal models will be useful for
studies of isostasy, dynamic topography, and crustal evolution of the continent.
© 2012 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
2. Compilation of published crustal thickness data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
3. Assessing gravity-based and surface-wave based models of crustal thickness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
4. Crustal model A — point constraints and gravity-based studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
4.1. Data and method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
4.2. Model results and data misfit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
5. Crustal model B — point constraints and surface-wave tomography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
5.1. Data and method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
5.2. Model results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
5.3. Data misfit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
6. Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
6.1. Stable continental interior . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
6.2. Andean range . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
6.3. Sub-Andean and Chaco basins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

⁎ Corresponding author.
E-mail addresses: marcelo@iag.usp.br (M. Assumpção), mei_feng_cn@yahoo.com.cn (M. Feng), andrestassara@udec.cl (A. Tassara), jordi@geofisica.ufrn.br (J. Julià).

0040-1951/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.tecto.2012.11.014
M. Assumpção et al. / Tectonophysics 609 (2013) 82–96 83

7. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

1. Introduction America and/or focused on a single data type. Models CRUST5.1 and
CRUST2.0, for instance, were based entirely on crustal thickness esti-
Mapping variations of crustal thickness in the continents have many mates from active-source profiling and ignored constraints on crustal
important applications for the study of the continental crust and litho- thickness from passive-source studies. Soller et al. (1982) used both
sphere. Besides giving information on the crustal evolution, degree of active-source profiling and surface-wave studies, but included just a
isostatic compensation (e.g., Sacek and Ussami, 2009), and intraplate few studies in the Andean region for South America leaving crustal
stress patterns (e.g., Lithgow-Bertelloni and Guynn, 2004), crustal thickness estimates for the majority of the continent as mere extrap-
thickness estimates are essential for modeling wave-propagation in olations (Tanimoto, 1995). Our dataset largely improves and updates
global and regional seismic studies (e.g., Hjörleifsdóttir and Ekstrom, the constraints provided in those earlier compilations.
2010; Lebedev and van der Hilst, 2008), monitoring regional-scale Comparing the newly compiled set of seismic crustal thickness
seismicity (e.g. Ferreira et al., 2008), and for source discrimination in with the Bouguer Anomaly we derived an empirical relationship
the framework of the Comprehensive Nuclear Test-Ban Treaty that is then used for predicting crustal thickness in areas where no
(e.g. Cormier and Anderson, 2004; Fan and Lay, 1998). In addition, seismic data is available (such as parts of the northern Andes and
models of crustal thickness variation can serve for developing sur- the continental shelf). We developed two types of models of crustal
face corrections to investigate upper mantle structure through ei- thickness variation for South America. The first type consists of an
ther body-wave or surface-wave tomography studies (e.g. Bastow interpolation of the seismically-constrained and gravity-predicted
et al., 2008; Park et al., 2008) and/or through the study of the crustal thicknesses, while the second type consists of an interpolation
Earth's normal-modes of vibration (e.g., Justowski et al., 2007; Mooney based on surface-wave studies. As expected, the addition of new
and Kaban, 2010). Also, the increasing use of shorter wavelengths in point constraints improves the resolution of crustal thickness esti-
global seismic modeling requires correspondingly more accurate mates in previously unsampled areas of the continent, such as North-
models of crustal thickness variation (e.g., Fichtner and Igel, ern Andes (Ecuador, Colombia and Venezuela) and Northeastern
2008; Fichtner et al., 2009). Brazil (Borborema Province and northern part of the São Francisco
In spite of its importance, crustal thickness is still among the least craton) and the southern part of the Paraná basin. However, large
known crustal properties of South America. Large areas of the conti- portions of stable South America remain poorly sampled and predic-
nent, such as the Amazon craton and the Chaco basin in NE Argentina, tions from our models within those regions could have significant er-
are sparsely sampled and detailed information on crustal structure is rors (around 5–10 km). We hope that this study will motivate the
lacking (e.g. van der Lee et al., 2001). The best known area is the deployment of temporary broadband experiments to fill in the gaps
Andean region, for which a number of passive and active seismic ex- in our knowledge. Despite the largely unsampled areas, we show
periments have been carried out (e.g. ANCORP, 2003; Gans et al., that derived maps of crustal thickness correlate with first order tec-
2011; McGlashan et al., 2008; Yuan et al., 2002) and detailed models tonic features of the continent and give new insights into the old geo-
of crustal structure have been developed by combining seismic data logical control on the present-day crustal structure.
and gravity modeling in Venezuela (Niu et al., 2007) and South and Finally, the point constraints compiled in this study have been uti-
Central Andes (Tassara and Echaurren, 2012). In addition, several lized in the validation of an independent model of crustal thickness
temporary seismic experiments in NW Argentina and Chile have variation for South America based on satellite gravity (van der
also mapped crustal structure over the region of flat-slab subduction Meijde et al., 2013–submitted for publication).
(e.g. Calkins et al., 2006; Gans et al., 2011) and the Bolivian altiplano
(e.g. Beck and Zandt, 2002; Beck et al., 1996; Swenson et al., 2000). In 2. Compilation of published crustal thickness data
the tectonically stable part of the continent, the most comprehensive
crustal thickness maps were produced by Feng et al. (2007) and Lloyd We expanded the recent compilation of seismic crustal thicknesses for
et al. (2010), through constrained tomographic inversion of surface- the Brazilian shield and adjacent regions (Assumpção et al., in press),
wave data. Those studies used both average epicenter-station 1D which included 229 previous point constraints from Feng et al. (2007),
models and group velocities and incorporated local constraints on 244 from Lloyd et al. (2010), 183 from Tassara and Echaurren (2012),
crustal thickness from independent receiver function studies and seis- and 200 from Pavão et al. (2012), to a total of 920 point constraints. Addi-
mic refraction profiles. Some additional information based on isostatic tional point constraints were incorporated for Venezuela (Niu et al., 2007;
assumptions was also included. Nonetheless, in spite of fitting seismic Schmitz et al., 2005), Central and Northern Andes (Dorbath et al., 1993;
point constraints with a root mean square (RMS) deviation of about Robalino, 1977 (apud Feininger and Seguin, 1983; Ocola et al., 1975);
3–4 km, these crustal thickness models still had errors around 10 km Pacific offshore margin (Agudelo et al., 2009; Hussong et al., 1976;
in areas without point constraints. Meyer et al., 1976); Western Argentina (Gans et al., 2011), NE Argentina
We have built on a previous compilation of seismic point (Rosa et al., 2010), South Central Peru (Phillips et al., 2012), southern
constraints for crustal thickness in the Brazilian shield and adjacent Puna (Bianchi et al., 2013), and Patagonia (Lawrence and Wiens, 2004).
regions (Assumpção et al., in press) to produce an enlarged compila- Point constraints developed in Assumpção et al. (in press) from receiver
tion of 920 point constraints, 730 onshore and 190 offshore, for the functions for 19 newly installed seismic stations in Brazil were also
whole of South America. To our knowledge, this is the most compre- included.
hensive compilation of seismic point constraints on crustal thick- The point constraints come from two different seismic data types:
nesses for the continent. Previous continental-scale compilations of active source experiments (deep seismic refraction lines, or deep seis-
crustal thickness were performed during the development of global mic reflection surveys) and receiver functions (see Fig. 1). Logically,
crustal thickness models, such as CRUST2.0 (Bassin et al., 2000), point constraints from active-source profiling required sampling the
CRUST5.1 (Mooney et al., 1998), or the global compilation of Soller seismic line at regular intervals. For 2D seismic refraction models,
et al. (1982), but those studies had very sparse coverage in South points were selected at every 50 km, on average; for old, 1D models,
84 M. Assumpção et al. / Tectonophysics 609 (2013) 82–96

only a few points were selected with larger spacing. Receiver func- Andes (Mooney et al., 1979; Ocola et al., 1975), the assigned uncertainty
tions, on the other hand, are sensitive to the structure around the re- could reach 10 km. For receiver function point constraints based on the
cording station and naturally provide point constraints (e.g. Ammon, hk-stacking technique of Zhu and Kanamori (2000), uncertainties in
1991). Also, in the offshore areas, the seismic constraints were crustal thickness were generally reported during the analysis; for re-
complemented with published crustal thickness estimates based on ceiver function point constraints involving modeling of the waveforms,
Bouguer gravity anomalies constrained by seismic data, such as confidence bounds equal to ±one layer thickness were assigned to the
Mohriak et al. (2000) and Zalán et al. (2011) for the Atlantic margin estimates (usually ±2 km).
and Schmitz et al. (2005) for Venezuela. More details of the compila- It is important to realize that, for the same point, several estimates
tion can be found in Assumpção et al. (in press). of crustal thickness might be available. For instance, analysis of re-
Our compilation also includes uncertainties for each point con- ceiver functions at some stations was performed by several authors
straint. If uncertainties were not provided by the source, a rough esti- utilizing different techniques or, sometimes, even the same tech-
mate was obtained based on the quality of the processed data. For nique. Not surprisingly, a range of estimates for crustal thickness is
example, for recent seismic refraction experiments interpreted with available at those stations and the estimates vary with the choice of
ray tracing techniques in 2D models (such as Schmitz et al., 2005; filter bandwidth, event selection, and assumed parameters during
Soares et al., 2006), we assigned an uncertainty of 2 km for crustal the data processing. Assumpção et al. (in press) compiled all available
thicknesses. For old, unreversed seismic experiments interpreted with estimates of crustal thickness for the Brazilian shield and adjacent
1D models, such as the Nariño Project in the Ecuadorean/Colombian areas and calculated the standard deviations (σH) of the mean crustal

Fig. 1. Distribution of data points from seismic refraction/reflection profiles and stations with Receiver Function. Color scale indicates crustal thickness (from surface to Moho). Major geo-
logical provinces are indicated as follows: GS and CBS=Guyana and Central Brazil shields of the Amazon craton; SFC=São Francisco craton; BB, TT and MQ are the Borborema, Tocantins
and Mantiqueira NeoProterozoic foldbelt provinces, respectively; intracratonic basins: Am=Solimões and Amazon, Pn=Parnaíba, Pc=Parecis, Pt=Pantanal, Pr=Paraná, Ch=Chaco;
Or=Oriente Basin; PTG=Patagonia province (northern limit from Ramos, 2004). Blue solid line is the 3000 m altitude.
M. Assumpção et al. / Tectonophysics 609 (2013) 82–96 85

thickness (H) taking into account the uncertainties of each individual


measurement. Fig. 2 shows the distribution of the standard deviations 50
(σH) of the crustal thickness (H). Most measurements agree to within
± 2 km, and scatter (σH) less than 1 km is rare. In the final dataset, a
minimum uncertainty of 1 km was assigned to all stations with only
one reported value. 40
It has been argued that refraction data and receiver functions may
not sample exactly the same Moho depth (e.g., Stratford et al., 2009).
The high frequency (5 to 10 Hz) Pn head-waves, used in seismic re-
fraction studies, sample the topmost upper mantle. Lower frequency 30
receiver functions (around ~ 1 Hz), in areas of gradational Moho,
may sample roughly the middle of the Moho transition zone, thereby
giving slightly lower values for the crustal thickness. However, within
20
the uncertainties of our compiled data (roughly 2 to 3 km for both re-
ceiver functions and seismic refraction) the two types of data should
be consistent. In addition, most of the onshore seismic refraction lines
are close to seismic stations with receiver function data; the 1 degree 10
spatial resolution of our model smears out possible systematic differ-
ences between the two types of data.
All compiled point constraints in South America with crustal thick-
nesses derived from seismic methods (both deep seismic refraction/ 0
reflection experiments and receiver functions) are shown in the 10 20 30 40 50 60 70 80
map of Fig. 1, and a histogram of crustal thicknesses for the complete Crustal thickness (km)
dataset is shown in Fig. 3. Thicknesses less than 20 km are mostly
from seismic refraction lines in oceanic areas and have smaller uncer- Fig. 3. Distribution of the 920 compiled crustal thicknesses for the whole area
(RF + refraction/reflection data), in both continental and oceanic areas. Dark shade
tainties. Thicknesses larger than 55 km refer to the Andean range. The shows data with uncertainty ≤3 km; light shade denotes data with larger uncer-
mean crustal thickness in the stable part of the continent for the com- tainties. The three groups reflect the preferential distribution of the experiments in
piled data is 38 km with a standard deviation of 4.8 km (Assumpção the compiled literature.
et al., in press). We should note that this is not the mean thickness of
the stable continent because the sampling is not geographically uniform
throughout the continental interior. gravity-derived crustal models, for instance, assume that the anoma-
lies are due mainly to Moho topography with a constant density
3. Assessing gravity-based and surface-wave based models of contrast between lower crust and upper mantle, which may not be
crustal thickness accurate enough, especially in the stable part of the continent.
Surface-wave tomography, on the other hand, does not depend on as-
To construct models of crustal thickness for South America using sumptions about isostatic mechanisms, but is affected by uneven dis-
the point constraints displayed in Fig. 1, we need to interpolate the tribution of epicenter-station paths. In addition, there is a trade-off
dataset within poorly sampled areas of the continent. To accomplish between crustal thickness and upper mantle velocity, which usually
this goal we explore two possibilities: 1) using crustal models derived prevents a resolution better than about 5 km or so when inverting
from gravity anomalies, and 2) using surface wave tomography. Both for crustal thickness. Before interpolating our point constraints,
approaches, however, have advantages and disadvantages. Many therefore, we first make an assessment of previously gravity-based
and dispersion velocity-based models of crustal thickness for South
America.
30
We start by assessing the accuracy of two previous surface-wave
based crustal thickness maps for South America (Feng et al., 2007;
25 Lloyd et al., 2010). Both maps were obtained by joint inversion of
surface-wave group velocities, waveform modeling and sparse point
constraints on crustal thickness. Both models fit the crustal thickness
20 data used in their joint inversion with an rms misfit of about 3 km. To
assess the real accuracy of those models, we use 56 of our newly
compiled point estimates of crustal thicknesses (not utilized in their
15 surface-wave studies) as test points for the surface-wave based
models. Fig. 4 shows the two models and the 56 test points, all located
more than 100 km away from any of the point constraints of their
10 joint inversion. In the Andean region, Fig. 4 shows that the model of
Feng et al. (2007) fits 17 test points with an rms residual of 9.4 km,
while the model of Lloyd et al. (2010) fits the test points with an
5 rms residual of 10.5 km. In the stable part of the continent, Fig. 4
shows that 39 test points have an rms misfit of 4.7 and 5.8 km, re-
spectively. Recall that for the stable part of the continent (Fig. 1),
0 the compiled data for crustal thickness had an average value of
0 1 2 3 4 5 6 7 8
~ 38 km and a standard deviation of 4.9 km. This means that neither
st.dev.(km) Feng et al. (2007) nor Lloyd et al. (2010) predict Moho depths in
Fig. 2. Distribution of standard deviations of crustal thickness for all stations with three
the unsampled regions much better than the expected variability.
or more different estimates (i.e., different authors or different techniques). That is, a model with a constant thickness of 38 km in the unsampled
From Assumpção et al. (in press) areas would fit the data almost as well as the models obtained with
86 M. Assumpção et al. / Tectonophysics 609 (2013) 82–96

A B

Fig. 4. (A,B) Misfits of previous models of Moho depth (top: Feng et al., 2007; bottom: Lloyd et al., 2010) to 56 new point constraints far from the ones used to derive the models.
Red circles mean the new measurements indicate a Moho 4 km or more deeper than the model; blue circles indicate the Moho is more than 4 km shallower than the model.

the tomographic inversions. This clearly shows the need for an im- model of Tassara and Echaurren (2012) for Central and Southern
provement in the model derived from the surface wave tomography. Andes to guide the development of our crustal thicknesses in areas
Next, we assess the performance of a gravity-based model for the of poor seismic coverage, such as the sub-Andean region. This gravity
Andean region. Tassara and Echaurren (2012) developed a detailed 3D model has errors larger than ±8 km near the Amazon, far from the
model of Central and Southern Andes by forward modeling of gravity Andean seismic constraints (see Fig. 5). For this reason, points in
data constrained by 183 seismically derived Moho estimates (from re- the Amazon too far from the Andes were excluded from our model A.
ceiver functions and active experiments) and other seismic and thermal In the northern Andes, where not much data was found in the liter-
information to fix the subducted slab and lithosphere–astenosphere ature, we use point constraints derived from the following relationship
boundary. Their model fits these 183 points with an rms misfit of
5.8 km. This misfit is due to scatter in the compiled point constraints HðkmÞ ¼ 32:96 − 0:1090  Boug þ 5:00e−5  Bougˆ2 þ 2:90e−7
(uncertainties in crustal thickness estimates from receiver functions)  Bougˆ3; ð1Þ
but also due to smoothing of the model in regions with strong lateral var-
iations. Fig. 5 shows 44 newly compiled points, which were not used in which was developed by fitting those point constraints from our compi-
their gravity modeling. These points have an rms misfit of 5.9 km, very lation (Fig. 1) that are located in the Andes and the adjacent Pacific mar-
similar to the misfit of the actual data. With exception of the northeastern gin to Bouguer anomalies taken from the continental-scale map of Sá
area (Western Amazon, Brazil, Fig. 5) the model of Tassara and Echaurren (2004) (see Fig. 6). The relationship fits the point constraints with an
(2012) seems to make a useful prediction of Moho depths for the rms residual of 7.4 km. This means that crustal thicknesses estimated
sub-Andean area, with an expected error of about 6 km. from gravity data should have an expected standard error of about
Given the result of the assessments, we have opted for developing 7 km. This is less than the rms misfit of the tomographic models of
two crustal models, both including gravity-derived estimates in the Feng et al. (2007) and Lloyd et al.(2010) in the Andes (9.4 and
Central and Southern Andes (model of Tassara and Echaurren, 2012) 10.5 km, respectively, as seen in Fig. 4). Given that crustal thicknesses
and in northern Andes (empirically estimated from Bouguer anomalies) across the Pacific margin and the Andean cordillera vary widely (from
to fill large gaps in seismic data: model A — using the seismic point 10 to 20 km in the ocean, to more than 60 km in the high Andean
constraints shown in Fig. 1, with simple interpolation in the stable part range, and back to ~35–45 km in the sub-Andean basins), we think
of the continent, and model B — using seismic point constraints, and that this empirical relation is useful to fill large gaps in seismic data cov-
surface-wave tomography. erage, as in the Northern Andes. For the Atlantic margin, Assumpção et
al. (in press) derived the following quadratical empirical relation using
4. Crustal model A — point constraints and gravity-based studies only data in the oceanic areas:

4.1. Data and method HðkmÞ ¼ 31:85 − 0:1291  Boug þ 0:0002098  Bougˆ2: ð2Þ

In this model, we used the crustal thicknesses from the 762 seis- The rms residual of 3.1 km is similar to the uncertainties assigned
mic constraints (Fig. 1) complemented with the gravity-based crustal to the compiled thickness data in the Atlantic margin. This relation
M. Assumpção et al. / Tectonophysics 609 (2013) 82–96 87

80° 75° 70° 65° 60° −85°−80° −75°−70°−65° −60°−55° −50°−45°−40° −35°
5° 5°

10°


10° 10°

−5°

15° 15° −10°

−15°

−20°
20° 20°
−25°

−30°

25° 25° −35°

80 −40°
70
30° 30° Pacific Margin + Andes
moho, km

60
50 80
40
30 70
Crustal thickness, km

35° 20 35° 60
10

50

40

40° 40°
30

20

10

45° 45° −400 −200 0 200


80° 75° 70° 65° 60° Bouguer, mGal
Fig. 5. Moho depths from gravity model of Tassara and Echaurren (2012). Open circles Fig. 6. Relationship between crustal thickness and Bouguer anomaly in the Pacific mar-
are the seismic constraints used in the gravity modeling. Larger circles are new seismic gin and Andean region. Red circles denote points used to establish a 3rd order
points not used in the gravity modeling. Grey circles denote misfits less than ±4 km; polinomial (blue line in the bottom plot). Blue crosses in the top show the points
red circles denote observations are 4–8 km deeper than the model; blue circles denote with estimated crustal thicknesses used to complement the seismic data of Fig. 1 and
Moho depths 4–8 km shallower than the model; larger circles denote misfits larger make the contour map of crustal thicknesses.
than ±8 km, such as in the northern area. The data used in the modeling (open circles)
have an rms misfit of 5.8 km. The rms misfit of the 49 new data points is also 5.8 km.

continent is not very useful and we preferred to interpolate


was also used here to fill data gaps in the oceanic areas of the Atlantic between widely spaced data points (as done by Assumpção et al., in
margin. press) rather than introduce bias from Bouguer anomalies in the con-
For the stable continental interior (onshore, East of the Andes), tinental interior.
Assumpção et al. (in press) showed that the correlation between
elevation and crustal thickness is poor (only 0.20). The correlation 4.2. Model results and data misfit
between crustal thickness and Bouguer anomalies is slightly better
(correlation coefficient of − 0.38). However, the regression of Fig. 7 shows the crustal model constructed by interpolating seismic
crustal thicknesses and Bouguer anomalies has an rms misfit of point constraints (receiver functions and active seismics) complemented
4.4 km (Assumpção et al., in press). The 260 crustal thicknesses with thicknesses estimated from seismically-constrained gravity model-
compiled for the stable part of the continent have a mean value ing in the Andes (Tassara and Echaurren, 2012), and gravity-predicted
of 37.7 km and a standard deviation of 4.8 km. For this reason, estimates in the northern Andes and oceanic areas (one point every
using gravity to fill large data gaps in the stable part of the 2.5°). We interpolated the data with the minimum curvature technique
88 M. Assumpção et al. / Tectonophysics 609 (2013) 82–96

(Smith and Wessel, 1990; Wessel and Smith, 1998) after getting weight- 5. Crustal model B — point constraints and surface-wave tomography
ed mean values every 0.5°. A surface tension factor of 0.5 was used to
avoid large maxima and minima outside the data range. The rms fit to 5.1. Data and method
the 920 seismic points is 3.5 km. The rms fit to the 260 seismic points
within the stable continental area is 1.6 km, close to the average accura- Here we use the tomographic inversion method of Feng et al.
cy of the seismic data (Fig. 2). (2007) to derive a crustal model based on joint inversion of different
Along the Andean chain, crust thicker than 60 km is seen from kinds of data: waveform modeling, group velocity dispersion, and
central Peru (8°S) to the Puna plateau (30°S). In central Ecuador seismic point constraints. The waveform data was the same as used
and northern Peru (3°S to 7°S) the crust does not seem to be by Feng et al. (2007): 1537 regional wave trains analyzed with the
more than 50 km thick, although no direct seismic data is avail- partitioned waveform inversion (PWI) method of van der Lee and
able in this region. In the stable continental interior, two main Nolet (1997) and van der Lee et al. (2001). Each waveform modeling
areas of thick crust (40 km or more) can be seen, one in the (S+ Rayleigh waves) produces an average 1D model for the epicenter-
Amazon craton and the other in the São Francisco and Paraná station path. To the 5700 group velocity measurements of Feng et al.
basin. The northern part of the intracratonic Paraná basin shows (2004, 2007), we added 1031 new measurements (Collaço et al.,
the thickest crust, reaching more than 45 km. In the southern 2012; Rosa et al., 2012) including inter-station paths using ambient
part of the continent, east of the Andes, the crust is generally thin- noise cross-correlation in the Paraná and Chaco basins (Fig. 8). The
ner than 40 km. A belt of thin crust (reaching 35 km or less) is same seismic point constraints used in model A were also used here,
seen along the sub-Andean basins, especially in Southern Colom- with large weighting, to constrain model B. Additionally, point con-
bia, Peru and Bolivia. This is based mainly on gravity data, but straints from the gravity model of Tassara and Echaurren (2012), and
the few available seismic points seem to be consistent with this from the empirical Bouguer relations for unsampled areas in the North-
feature. ern Andes and oceans, were also used, but with lower weights.

20
10°
40

30
50

40


20

−5°
40
70
30
−10°
60

−15° 50

40
−20°

30
−25°
40

20

−30° 10
30

−35°

−40°

−45°
−80° −75° −70° −65° −60° −55° −50° −45° −40° −35°

Fig. 7. Contour map of crustal thickness of MODEL A (contour interval 5 km). Red dots show seismic data (RF and active seismics). Crosses in the oceanic areas denote thicknesses
inferred from Bouguer anomalies. White dots denote thickness data given by seismically-constrained gravity modeling of Tassara and Echaurren (2012). Green lines are boundaries
of the main geological provinces (see Fig. 1 for legend). Values of crustal thickness in the ocean include the water layer.
M. Assumpção et al. / Tectonophysics 609 (2013) 82–96 89

First, a 2D group velocity tomography is carried out on a 1° by 1° geo- for the Andean region from Tassara and Echaurren (2012) were added.
graphical grid, for each period, with the ~(5700 +1031) paths using the Fig. 9 shows the resulting model based only on waveform modeling
ray-theory tomography method of Feng et al. (2004). Ray-theory to- and Rayleigh-wave group velocity measurements (model B1). Thick
mography can offer nearly identical result as finite-frequency tomogra- crust can be seen in Southern Peru, Southern Bolivia and Northern
phy (Ritzwoller et al., 2002; Sieminski et al., 2004) in continental-scale Chile. Interestingly, the thick crust in the Bolivian Altiplano (between
studies if appropriate smoothing is applied and the path distribution is 15oS and 20oS) is not retrieved by the surface-wave data only. This is
good. The regionalized dispersion curves are then combined with the probably due to irregular distribution of ray paths and few ray crossings
1D path-averaged models (obtained from the PWI inversions) to obtain over the Bolivian Altiplano. Similarly, in the Ecuadorian/N.Peruvian area,
perturbations of a 3D initial S-velocity structure. The initial reference 3D the surface-wave data does not indicate a thick crust. In the stable part of
model, called SA40, is the IASP91 upper mantle model with a 40 km the continent, the surface-wave data indicate thicker crust (>40 km) in
crust in the stable part of South America and 45 to 70 km in the Andes the Amazon and thinner in the Atlantic shield.
(van der Lee et al., 2001). The two data sets are then jointly inverted Adding the seismic depth constraints we get model B2 (Fig. 10). In
with the following set of equations (Feng et al., 2007; Lloyd et al., 2010): this model extra points were also used in the ocean away from the
seismically determined point constraints (based on the empirical re-
0 1 0 1
Hw qw lationship with Bouguer anomaly, Eqs. (1) and (2)), as well as in
B λd Hd C B λd qd C the northern Andes. The thick crust of the Bolivian altiplano is now
B C B C
@ λRF HRF Am ¼ @ λRF qRF A ð3Þ
better defined with the inclusion of receiver function data. In the At-
λI 0 lantic shield, the thick crust (> 40 km) spans a larger area, especially
in the Paraná Basin and near the São Francisco craton. A large region
where Hw, Hd and HRF are the sensitivity matrix for the waveform, group of thin crust (b35 km) is seen in the Chaco Basin.
velocity data and seismic point constraints, respectively; m is the model Fig. 11A shows our preferred and final tomographic model (B3)
parameters, i.e., 3D S-wave velocity perturbations (Δβ) and Moho depth where, in addition to the data used in model B2, we added the
perturbations (ΔH) from the reference model SA40; qw is the data vector gravity-derived crustal thicknesses of Tassara and Echaurren (2012),
related to the PWI data; qd is the data vector of group velocity difference which had been tightly constrained by seismic observations. Model
between the regionalized and predicted group velocities (ΔU); qRF is the B3 fits the 730 on-shore seismic point constraints with an rms misfit
vector with crustal thickness data; λ is a damping factor, and λd and λRF of 4.7 km (Fig. 11B. For the stable continental interior (260 points
control the relative weights between waveform, group velocity and onshore), the misfit is only 1.7 km. The misfit in the Andes (Fig. 11B)
crustal thickness data. In addition, we weight all the data by their re- is larger partly because of smoothing of the model in the inversion pro-
spective inverse estimated uncertainty. cess and partly because of larger variability of the compiled data. The
weights in the tomographic inversion were chosen so that the seismic
5.2. Model results point constraints have relatively more influence than waveform or dis-
persion data. Model B3 improves the definition of the thick Andean
We present three versions of model B, obtained with different combi- crust. However, in the region of Southern Ecuador and Northern Peru
nations of seismic data types, to investigate their contribution to the final (where no seismic data have been compiled), the crust remains rela-
model: model B1 — based on waveform modeling and Rayleigh-wave tively thin, about 40 km or less.
group dispersion; model B2 — where the point constraints compiled in The map of Fig. 12 shows the S-wave velocity anomalies at 100 km
this study were added, including those derived from the empirical rela- depth, which is not the main topic of this work but is briefly described
tionships (Eqs. (1) and (2)); and model B3 – where gravity constraints here as an interesting by-product of the tomographic inversion. The

BRASIS+4BRZ evts LPA ANT


−90° −80° −70° −60° −50° −40° −30° −20° −90° −80° −70° −60° −50° −40° −30° −20° −90° −80° −70° −60° −50° −40° −30° −20°
20° 20° 20°

10° 10° 10°

0° 0° 0°

−10° −10° −10°

−20° −20° −20°

−30° −30° −30°

−40° −40° −40°

−50° −50° −50°

−60° −60° −60°


−90° −80° −70° −60° −50° −40° −30° −20° −90° −80° −70° −60° −50° −40° −30° −20° −90° −80° −70° −60° −50° −40° −30° −20°

Fig. 8. Additional 1031 paths with group velocity measurements used to complement the ~5700 paths previously used by Feng et al. (2004, 2007). Left) Earthquake data from re-
cently installed Brazilian stations and some GSN stations between 2009 and 2011. Center) Paths measured at LPA station (University of La Plata; María Laura Rosa 2012). Right)
Interstation paths measured with ambient noise cross-correlation (Bruno Collaço, USP, 2012).
90 M. Assumpção et al. / Tectonophysics 609 (2013) 82–96

Fig. 9. Model of Moho depths (tomographic inversion method of Feng et al., 2007) using only waveform modeling and group velocity dispersion (Model_B1_WF + DSP). Same
waveform data as in the 2007 model; dispersion data from 2007 with additional paths from Fig. 8. Blue line denotes 3000 m elevation of the Andean plateau. Note generally
thick crust in the Andes, except at the Southern Bolivia plateau (near 18°S), probably due to insufficient ray-crossings in that area. No deep crust is suggested in the Ecuadorean/
N. Peru area (0° to 5°S). Deep crust in the northern and southern Paraná basin is suggested, but not near the São Francisco craton.

results are very similar to those of Feng et al. (2007) except in the the model inverted with only the specific data. Interestingly, the WF
southern part of the Paraná Basin where higher velocities are found. data improves the group velocity fit, relative to the initial model
The inclusion of more dispersion paths (Fig. 8) improved the veloci- (Fig. 13B), but using only the DSP data one cannot improve the wave-
ties of the lithospheric lid in Eastern Argentina. The region of the form fit of the seismograms (Fig. 13D). The amplitude and phase in-
Chaco Basin is characterized by low velocities in the upper-mantle, formation contained in the PWI modeling are very important to better
which correlate with the generally thin crust (Fig. 11A), as discussed constrain the final model.
below. In the Andes, the velocities at 100 km depth are generally low
reflecting the shallow asthenosphere, except for the high velocities
in the two areas of flat subduction of the Nazca slab: in southern 6. Discussion
Ecuador/northern Peru near 5°S, and in central Chile/Argentina near
30°S. 6.1. Stable continental interior

Both models A and B3 show that the average crustal thickness in


5.3. Data misfit the stable part of the continent is about 40 km. Two regions with
crust generally thicker than 40 km can be seen in the Amazon and
Examples of fitting of the surface-wave data for the final model B3 in the Atlantic shield. This is seen both in model A (Fig. 7, based on
are shown in Fig. 13, compared with the initial reference model SA40. seismic constraints only) and model B3 (Fig. 11A) including
Models inverted with only one type of data (DSP = group velocities surface-wave tomography. The more detailed features in the final
only, or WF = PWI waveform data only) are also shown for compari- model B3 indicate that the thick crust in the Amazon correlates
son. The joint model B3 fits the DSP and the WF data almost as well as roughly with the oldest Archean provinces as defined by Tassinari
M. Assumpção et al. / Tectonophysics 609 (2013) 82–96 91

Fig. 10. Model B2 of Moho depths (tomographic method of Feng et al., 2007) using waveform modeling and group velocity dispersion (same data as in Fig. 9) plus 762 seismic point
constraints. In the oceans and Northern Andes, gravity-derived additional points were used to fill large gaps, based on Eqs. (1) and (2) (shown in Fig. 7).

and Macambira (1999). In the Atlantic shield (i.e., SE of the Crustal thicknesses of the Rio de la Plata craton of eastern Argen-
TransBrasilian Lineament, Fig. 11A) there is evidence of thick crust tina also seem to be larger than 40 km The crust in the Patagonian
associated with the São Francisco craton but the deepest Moho province, in the southern part of the continent, appears thinner
(reaching more than 45 km) is found in the intracratonic Paraná than 40 km, but the coverage of our data set is poor in the southern
Basin based mainly on receiver functions (Assumpção et al., 2002; tip of South America and model B3 has poor resolution in this area.
Julià et al., 2008). The basin has generally lower elevations than the
surrounding areas, which led Assumpção et al. (2002) to propose 6.2. Andean range
that the deep crust in this basin could be compensated by high densi-
ties in the lithospheric lid. This hypothesis remains consistent with The Moho beneath the Andes can reach more than 70 km depth in
high S-wave velocities at 100 km depth shown in Fig. 12. the eastern cordillera of Bolivia, and is generally deeper than 50 km in
A SW–NE continental-scale lineament, the TransBrasiliano Linea- areas with elevations higher than 3000 m. A notable exception to this
ment (“TBL” in Fig. 11A), is believed to indicate a major Neo- is observed below the northern Puna at 25°S where the crust is thinner
Proterozoic suture between the Amazon craton and the other provinces than 55 km for elevations higher than 4000 m. This result reinforce the
to the SE (São Francisco craton and other old crustal units). In model B3, hypothesis as to that a recent (b 4 Ma) event of lithospheric delamina-
the TBL seems to be characterized by relatively thin crust, especially be- tion has affected this region and removed part of the dense lower
neath the Parnaiba basin. Another pronounced belt of thin crust is seen crust (Kay and Kay, 1993; Whitman et al., 1996). Interestingly, and
in model B3 along the eastern and southern borders of the Amazon cra- not noted by previous models (Feng et al., 2007; Lloyd et al., 2010;
ton. This feature could indicate an extension of a narrow belt of thin Tassara et al., 2006), this region of thin crust seems to be oriented in a
crust found by Assumpção et al. (2004a, 2004b) in the Tocantins prov- ENE direction that could be continued eastward into the TransBrasiliano
ince of central Brazil roughly coincident with the Neo-Proterozoic Lineament. The SW limit of the TBL is presently characterized by a me-
Goiás magmatic arc. chanically weak (Pérez-Gussinyé et al., 2007), thinned lithosphere
92 M. Assumpção et al. / Tectonophysics 609 (2013) 82–96

A Feng2012SET_Moho_3 B Feng2012SET_Moho_3
Caribbean gravity
10˚ 10˚
40 empirical
Tassara
5˚ 5˚

40 40
0˚ 0˚

L
TB
Or
−5˚ −5˚
40 40
40
40

40
−10˚ −10˚

−15˚ −15˚

70
30 40 30
−20˚ 60 −20˚
M depth, km

60 30 30
50
−25˚ −25˚
40

30 40 40
−30˚ −30˚ residual, km
(seismic)
40

20

10 > +4
−35˚ −35˚
−4 to +4
< −4
−40˚ −40˚
−80˚ −75˚ −70˚ −65˚ −60˚ −55˚ −50˚ −45˚ −40˚ −35˚ −80˚ −75˚ −70˚ −65˚ −60˚ −55˚ −50˚ −45˚ −40˚ −35˚

Fig. 11. A) Model B3 of Moho depths using waveform modeling, group velocity dispersion, all point constraints as model B2 with additional depths from the model of
Tassara and Echaurren (2012). Control points are shown in panel B. Contour interval 5 km. Blue solid line denotes the 3000 m altitude in the Andes. Dark green lines
are boundaries of major geological provinces. Gray solid lines are the limits of the South American plate. Red line in the Amazon is the oldest Archean geochronological
province (Tassinari and Macambira, 1999). TBL = TransBrasiliano Lineament; Ch = Chaco Basin, Or = Oriente basin. Note better definition of the Andean thick crust. Note
that in South Ecuador and Northern Peru (~ 3°S), the crust is still thin. B) Fit to MODEL B3. Crustal thickness contours with all point-constraints used in the continent:
colored circles are the seismic constraints; rms misfit to all 728 continental data points = 4.7 km; fit to 259 points in stable continent = 1.7 km). Small open circles denote
control points from Tassara and Echaurren (2012) gravity-derived model; crosses show locations of thicknesses derived from the empirical relationship with Bouguer
anomaly (Eq. (1), Fig. 6).

(Assumpção et al. (2004b) as suggested by the low S-wave veloc- (>40 km) for the Ecuador–Peru border region compared to the rest of
ities in Fig. 12). Although highly speculative, this hypothesis the Andes (b20 km).
could suggest that delamination below the Puna occurred along
a pre-weakened, thinned lithospheric zone inherited since the 6.3. Sub-Andean and Chaco basins
Neo-Proterozoic.
In southern Ecuador and northern Peru (around 3°S), at the limit Another interesting feature of the Moho models is a large region of
between central and northern Andes, another region of relatively thin crust (shallower than b 35 km) between the Andean range and
high elevation (near 3000 m) is underlain by a Moho that seems to the cratonic areas. In some areas of the Chaco basin, model B3
be only about 40 km or less. In this area there are no direct seismic suggests a Moho depth even shallower than 30 km (Fig. 11A).
constraints and the thin crust is controlled by the surface wave data Although the Moho model in the sub-Andean region (Fig. 11A is
and some gravity-predicted estimates of crustal thickness (Fig. 11B). mainly controlled by the surface-wave data, the few direct seismic
Geoid anomalies are mainly positive and higher than 20 m along measurements in these areas (Fig. 7) and the relatively high gravity
the whole Andean chain, reaching more than 40 m in the Altiplano. anomalies (Tassara and Echaurren, 2012; Tassara et al., 2006) confirm
However, in the south Ecuatorian Andes this anomaly almost van- the generally thin crust east of the Altiplano–Puna plateau of the
ishes and geoid heights are less than 10 m (Blitzkov et al., 2009; central Andes. Interestingly, this feature has been independently
King, 2002). Although long wavelength geoid anomalies are mainly found from the inversion of satellite gravity data (van der Meijde
controlled by deeper structures (subducted plates and density varia- et al., 2013–submitted for publication).
tions at the mid-lower mantle), the lack of a short wavelength positive Using surface-wave data to invert for both crustal thickness and
geoid anomaly can be connected to the lack of a deep crustal root in this S-wave velocity of the lithospheric lid can be a difficult problem be-
region. In this case, thin cordilleran crust can be the consequences of a cause of the trade-off between these two parameters. A solution
regional, flexural-controlled compensation of surface topography in with a thin crust could be compensated by lower velocities in the
contrast to the local, Airy-type compensation that seems to dominate upper mantle and produce similar dispersion curves. Snoke and
along the rest of the Andean cordillera as suggested by the correlation James (1997) analyzed phase and group velocity data in the Chaco
between topography and crustal thickness. Moreover, estimates of the basin and showed that this region has very low S-wave velocities in
elastic thickness that parametrizes the flexural rigidity of the lithosphere the upper mantle, which was later confirmed by Feng et al. (2007).
(Pérez-Gussinyé et al., 2008; Tassara et al., 2007) are anomalously high In addition, checker board resolution tests by Feng et al. (2007)
M. Assumpção et al. / Tectonophysics 609 (2013) 82–96 93

7. Conclusions

A new model of Moho depth was derived for continental South


America based on an increased data set of more than 900 crustal
thickness constraints from receiver functions and active seismic
experiments. The surface-wave data originally used by Feng et al.
(2007) was increased by ~ 1000 paths with new group velocity mea-
surements improving the coverage in the Paraná and Chaco basins. A
joint inversion of all these data sets, including extra (weak) con-
straints from gravity-derived depths in poorly covered areas, pro-
duced a model (model B3, Fig. 11A) fitting all used data reasonably
well, i.e. within their uncertainties.
The preferred model shows two main regions of thick crust in the
stable continental interior, one in the Amazon craton (mainly in the
oldest geochronological province) and the other in the Atlantic shield
(including the São Francisco craton and the Paraná intracratonic
basin) separated by a zone of thin crust which could be correlated
with the Transbrasilian Lineament.
The Moho in the Andean chain is usually deeper than 50 km in areas
with elevations higher than 3000 m, reaching more than 70 km in the
Bolivian Eastern Cordillera.
A shallowing of the Andean Moho seems to occur below the Puna of
north-western Argentina and in the border region between Ecuador
and Peru. Causes for the lack of a deep crustal root compensating high
elevations for these regions vary from a possible delamination event
and then a thermal compensation for the Puna, while for the southern
Ecuatorian Andes it seems likely that compensation is dominated by
the flexural support of a rigid lithosphere.
A belt of thin crust (b35 km thick) that is seen below most of the
Fig. 12. S-wave velocity anomalies at 100 km depth from the tomography model B3. sub-Andean zone between the Andean orogeny and the stable craton-
Red line in the Amazon denotes the oldest geochronological province of Tassinari ic shield, especially beneath the Chaco basin, is probably an old
and Macambira (1999). TBL is the TransBrasiliano Lineament.
inherited feature left after the rifting and drifting episodes related
to the separation of Rodinia during the Neo-Proterozoic and subse-
quent closure of ocean basins connected to the first orogenic cycles
showed that the low mantle velocities in the Chaco basin are well de- that formed the Phanerozoic Gondwana continent and then the mod-
termined. This means that this feature is robust and should not cause ern South America.
an artificially thin crust in our tomographic inversion.
We do not have a definitive explanation for the origin of this zone
of thin crust along the sub-Andean belt, but some points are worth Acknowledgments
discussing. Flexural uplift in the sub-Andean crust caused by Andean
load can only account for a few meters, even when 3D amplification This work is supported by CNPq grant 309724/2009-0 (MA), NSFC
effects due to the bend of the Andean chain are taken into account grant 41174039 (MF) and Fondecyt project 1101034 (AT). Part of the
(Sacek and Ussami, 2009). Dynamic topography of other possible data used in this paper was collected in the BRASIS project (IAG-USP)
origin (subducted slabs, lower crustal flow, etc.) normally accounts supported by “Rede de Estudos Geotectônicos” of Petrobras. We thank
for at most a couple of kilometers (Clark et al., 2005; Dávila et al., María Laura Rosa (UNLP), Bruno Collaço (IAG-USP) and Gerardo Sánchez
2010; Pérez-Gussinyé et al., 2008) and also seems unlikely to explain (INPRES) for dispersion data from their stations. We also thank Eric
a shallowing of the Moho by 5–10 km in a large area as observed for Sandvol (Missouri Univ.) for providing unpublished crustal thicknesses
the Chaco basin here. We note that crust of 30–35 km thickness is from the Southern Puna experiment, Kristin Phillips (Caltech) for
also observed in our models A and B3 along the eastern passive mar- unpublished data from the Southern Peru experiment, and Stéphane
gin of the continent where it is of course a consequence of the crustal Drouet (ON-Brazil) for unpublished data from RSIS network.
extension and thinning accompanying the rifting of Gondwana. The
area of thin crust bordering the Amazon and Rio de la Plata craton References
in the west, e.g. lying to the east of the Andes, coincides with what
has been proposed as the main arm of the intracontinental rift zone Agudelo, W.A., Ribodetti, J.-Y., Collot, S. Operto, 2009. Joint inversion of multichannel
seismic reflection and wide-angle seismic data: improved imaging and refined ve-
that separated Laurasia from proto-Gondwana during the disaggrega- locity model of the crustal structure of the north Ecuador–South Colombia conver-
tion of the supercontinent Rodinia and the aperture of the Iapetus Ocean gent margin. Journal of Geophysical Research 114, B02306 http://dx.doi.org/
at the Neo-Proterozoic (Cawood, 2005; Cordani et al., 2009; Rapela et 10.1029/2008JB005690.
Ammon, C.J., 1991. The isolation of receiver effects from teleseismic P waveforms. Bul-
al., 2007). The region of anomalously thin crust (b30 km) below the
letin of the Seismological Society of America 81 (6), 2504–2510.
Chaco basin occurs at the intersection of this old rifted margin ANCORP Working Group, 2003. Seismic imaging of a convergent continental margin
and the TransBrasiliano Lineament, which also coincide with the and plateau in the central Andes (Andean Continental Research Project 1996
(ANCORP'96)). Journal of Geophysical Research 108 (B7), 2328 http://dx.doi.org/
suture left after the closure during the Cambrian of the Clymene
10.1029/2002JB001771.
Ocean that separated Amazonia from Sao Francisco and Rio de la Assumpção, M., James, D., Snoke, A., 2002. Crustal thicknesses in SE Brazilian shield
Plata cratons (Escayola et al., 2011; Tohver et al., 2012). Thus, it by receiver function analysis: implications for isostatic compensation. Journal
seems very likely that the present-day crustal thickness configura- of Geophysical Research 107 (B1) http://dx.doi.org/10.1029/2001JB000422
(ESE2-1—ESE2-14, 2006).
tion at the interior of the South American continent reflects an old Assumpção, M., An, M., Bianchi, M., França, G.S.L., Rocha, M., Barbosa, J.R., Berrocal, J.,
inheritance left after previous steps of the Wilson cycle. 2004b. Seismic studies of the Brasília Fold Belt at the western border of the São
94 M. Assumpção et al. / Tectonophysics 609 (2013) 82–96

Group velocity (km/s) Group velocity (km/s)


4.0 1 2

3.5

3.0
4
2 1
2.5 3

4.0 3 4

3.5

3.0

2.5
20 40 60 80 100 120 140 20 40 60 80 100 120 140
Period (s) Period (s)
B
RMS Misfit of Group Velocity (km/s)

0.20

0.16

0.12

0.08
SA40
DSP
0.04 WF
JOINT
0.00
20 40 60 80 100 120 140
Period (s)

C
27.0 7
Epicenter distance (degrees)

2001/03/15 dep=37km mb=6.1

25.1 3 6
7
6

24.1 24 5 5

20.8 4

1
20.8 3

Obs. 2
19.2
SA40

8.4 JOINT 1

200 300 400 500 600 700 800 900 1000


Time (s)

D
200 1.629 (a) 200 1.632 (b)
SA40 DSP
Number

0 0
0.0 0.6 1.2 1.8 2.4 3.0 3.6 4.2 4.8 5.4 0.0 0.6 1.2 1.8 2.4 3.0 3.6 4.2 4.8 5.4

0.929 (c) 1.046 (d)


WF JOINT
Number

200
200

0 0
0.0 0.6 1.2 1.8 2.4 3.0 3.6 4.2 4.8 5.4 0.0 0.6 1.2 1.8 2.4 3.0 3.6 4.2 4.8 5.4
RMS Misfit RMS Misfit
M. Assumpção et al. / Tectonophysics 609 (2013) 82–96 95

Francisco craton, central Brazil, using receiver function, surface wave dispersion, Fichtner, A., Kennett, B.L.N., Igel, H., Bunge, H.-P., 2009. Full seismic waveform tomog-
and teleseismic tomography. Tectonophysics 388, 173–185. raphy for upper-mantle structure in the Australasian region using adjoint
Assumpção, M., Schimmel, M., Escalante, C., Rocha, M., Barbosa, J.R., Barros, L.V., 2004a. methods. Geophysical Journal International 179, 1703–1725 http://dx.doi.org/
Intraplate seismicity in SE Brazil: stress concentration in lithospheric thin spots. 10.1111/j.1365-246X.2009.04368.x.
Geophysical Journal International 159, 390–399. Gans, C.R., Beck, S.L., Zandt, G., Gilbert, H., Alvarado, P., Anderson, M., Linkimer, L.,
Assumpção, M., Bianchi, M.B., Julià, J., Dias, F.L., França, G.S., Nascimento, R.M., Drouet, S., 2011. Continental and oceanic crustal structure of the Pampean flat slab
Pavão, C.G., Albuquerque, D.F., Lopes, A.E.V., in press. Crustal thickness map of Brazil: region, western Argentina, using receiver function analysis: new high-resolution
data compilation and main features. Journal of South American Earth Sciences. results. Geophysical Journal International 186, 45–58 http://dx.doi.org/10.1111/
Bassin, C., Laske, G., Masters, G., 2000. The current limits of resolution for surface wave j.1365-246X.2011.05023.x.
tomography in North America. EOS Transactions of the American Geophysical Hjörleifsdóttir, V., Ekstrom, G., 2010. Effects of three-dimensional Earth structure on
Union 81, F897. CMT earthquake parameters. Physics of the Earth and Planetary Interiors 179,
Bastow, I.D., Nyblade, A.A., Stuart, G.W., Rooney, T.O., Benoit, M.H., 2008. Upper mantle 178–190 http://dx.doi.org/10.1016/j.pepi.2009.11.003.
seismic structure beneath the Ethiopian hot spot: rifting at the edge of the African Hussong, D.M., Edwards, P.B., Johnson, S.H., Campbell, J.F., Sutton, G.H., 1976. Crustal
low-velocity anomaly. Geochemistry, Geophysics, Geosystems 9, Q12022 http:// structure of the Peru–Chile trench: 8°–12°S latitude. The geophysics of the Pacific
dx.doi.org/10.1029/2008GC002107. Ocean Basin and its margins. AGU Geophysical Monograph 19, 71–85.
Beck, S.L., Zandt, G., 2002. The nature of orogenic crust in the central Andes. Journal of Julià, J., Assumpção, M., Rocha, M.P., 2008. Deep crustal structure of the Paraná Basin
Geophysical Research 107, 2230 http://dx.doi.org/10.1029/2000JB000124. from receiver functions and Rayleigh-wave dispersion: evidence for a fragmented
Beck, S.L., Zandt, G., Myers, S.C., Wallace, T.C., Silver, P.G., Drake, L., 1996. Crustal-thickness cratonic root. Journal of Geophysical Research 113, B08318 http://dx.doi.org/
variations in the central Andes. Geology 24, 407–410. 10.1029/2007JB005374.
Bianchi, M., Heit, B., Jakovlev, A., Yuan, A., Kay, S.M., Sandvol, E., Alonso, R.N., Coira, B., Justowski, B., Dziewonski, A.M., Ekstrom, G., 2007. Nonlinear crustal corrections for
Brown, L., Kind, R., Comte, D., 2013. Teleseismic tomography of the southern Puna normal-mode seismograms. Bulletin of the Seismological Society of America 97,
plateau in Argentina and adjacent regions. Tectonophysics 586, 65–83. 1756–1762 http://dx.doi.org/10.1785/0120070041.
Blitzkov, D., Matos, A.C.O.C., Lobianco, M.C.B., Campos, I.O., 2009. The progress of the Kay, R.W., Kay, S.M., 1993. Delamination and delamination magmatism. Tectonophysics 219
geoid in South America under GRACE and EGM08 models. IAG Assembly, Buenos (1–3), 177–189 http://dx.doi.org/10.1016/0040-1951(93)90295-U.
Aires, 2009 (Abstracts). King, S.D., 2002. Geoid and topography over subduction zones: the effect of phase
Calkins, J.A., Zandt, G., Gilbert, H.J., Beck, S., 2006. Crustal images from San Juan, Argentina, transformations. Journal of Geophysical Research 107 (B1), 2013 http://dx.doi.org/
obtained using high frequency local event receiver functions. Geophysical Research 10.1029/2000JB000141.
Letters 33, L07309 http://dx.doi.org/10.1029/2005GL025516. Lawrence, J.F., Wiens, D.A., 2004. Combined receiver-function and surface wave phase-
Cawood, P.A., 2005. Terra Australis Orogen: Rodinia breakup and development of velocity inversion using a Niching Genetic Algorithm: application to Patagonia.
the Pacific and Iapetus margins of Gondwana during the Neoproterozoic and Pa- Bulletin of the Seismological Society of America 94 (3), 977–987.
leozoic. Earth-Science Reviews 69 (3–4), 249–279 http://dx.doi.org/10.1016/ Lebedev, S., Van Der Hilst, Rob D., 2008. Global upper-mantle tomography with the au-
j.earscirev.2004.09.001. tomated multimode inversion of surface and S-wave forms. Geophysical Journal
Clark, M.K., Bush, J.W.M., Royden, L.H., 2005. Dynamic topography produced by lower International 173, 505–518 http://dx.doi.org/10.1111/j.1365-246X.2008.03721.x.
crustal flow against rheological strength heterogeneities bordering the Tibetan Lithgow-Bertelloni, C., Guynn, J.H., 2004. Origin of the lithospheric stress field. Journal
Plateau. Geophysical Journal International 162, 575–590 http://dx.doi.org/10.1111/ of Geophysical Research 109, B01408 http://dx.doi.org/10.1029/2003JB002467.
j.1365-246X.2005.02580.x. Lloyd, S., van der Lee, S., França, G.S., Assumpção, M., Feng, M., 2010. Moho map of
Collaço, B., Sánchez, G., Rosa, M.].L., Assumpção, M., Sabbione, N., Araujo, M., Rocha, M.P., South America from receiver functions and surface waves. Journal of Geophysical
2012. Crustal structure of the Paraná Basin from ambient noise tomography. XVI Research 115, B11315 http://dx.doi.org/10.1029/2009JB006829.
Congr. Geological Soc. of Peru, Lima, 23–26 September 2012, extended abstract. McGlashan, N., Brown, L., Kay, S., 2008. Crustal thickness in the central Andes from
Cordani, U.G., Teixeira, W., D'Agrella-Filho, M.S., Trindade, R.I., 2009. The position of the teleseismically recorded depth phase precursors. Geophysical Journal International
Amazonian Craton in supercontinents. Gondwana Research 15 (3–4), 396–407 175, 1013–1022 http://dx.doi.org/10.1111/j.1365-246X.2008.03897.x.
http://dx.doi.org/10.1016/j.gr.2008.12.005. Meyer, R.P., Mooney, W.D., Hales, A.L., Helsey, C.E., Woollard, G.P., Hussong, D.M.,
Cormier, V.F., Anderson, T.S., 2004. Efficiency of Lg propagation from SmS dynamic ray Kroenke, L.W., Ramirez, J.E., 1976. Project Nariño III: refraction observation across
tracing in three-dimensionally varying crustal waveguides. Pure and Applied Geo- a leading edge, Malpelo Island to the Colombian Cordillera Occidental. The geo-
physics 8, 1613–1633. physics of the Pacific Ocean basin and its margins. AGU Geophysical Monograph
Dávila, F.M., Lithgow-Bertelloni, C., Giménez, M., 2010. Tectonic and dynamic controls 19, 105–132.
on the topography and subsidence of the Argentine Pampas: the role of the flat Mohriak, W.U., Mello, M.R., Bassetto, M., Vieira, I.S., Koutsoukos, E.A.M., 2000. In: Mello,
slab. Earth and Planetary Science Letters 295 (1–2), 187–194 http://dx.doi.org/ M.R., Katz, B.J. (Eds.), Crustal architecture, sedimentation, and petroleum systems
10.1016/j.epsl.2010.03.039. in the Sergipe–Alagoas Basin, northeastern Brazil: Petroleum Systems of South
Dorbath, C., Granet, M., Poupinet, G., Martinez, C., 1993. A teleseismic study of the Atlantic Margins: AAPG Memoir, 73, pp. 273–300.
Altiplano and Eastern Cordillera in Northern Bolivia: new constraints on a litho- Mooney, W.D., Kaban, M.K., 2010. The North American upper mantle: density,
spheric model. Journal of Geophysical Research 98 (B6), 9825–9844. composition, and evolution. Journal of Geophysical Research 115, B12424 http://
Escayola, M.P., van Staal, Cees R., Davis, William J., 2011. The age and tectonic setting of dx.doi.org/10.1029/2010JB000866.
the Puncoviscana Formation in northwestern Argentina: an accretionary complex Mooney, W.D., Meyer, R.P., Laurence, J.P., Meyer, H., Ramírez, J.E., 1979. Seismic refraction
related to Early Cambrian closure of the Puncoviscana Ocean and accretion of the studies of the western cordillera, Colombia. Bulletin of the Seismological Society of
Arequipa–Antofalla block. Journal of South American Earth Sciences 32 (4), America 69, 1745–1761.
438–459 http://dx.doi.org/10.1016/j.jsames.2011.04.013. Mooney, W.D., Laske, G., Masters, T.G., 1998. CRUST 5.1: a global crustal model at 5°x5°.
Fan, G.W., Lay, T., 1998. Statistical analysis of irregular wave-guide influences on re- Journal of Geophysical Research 103 (B1), 727–747.
gional seismic discriminants in China. Bulletin of the Seismological Society of Niu, F., Bravo, T., Pavlis, G., Vernon, F., Rendon, H., Bezada, M., Levander, A., 2007. Re-
America 88, 74–88. ceiver function study of the crustal structure of the southeastern Caribbean plate
Feininger, T., Seguin, M.K., 1983. Simple Bouguer gravity anomaly field and the inferred boundary and Venezuela. Journal of Geophysical Research 112, B11308 http://
crustal structure of continental Ecuador. Geology 11, 40–44 http://dx.doi.org/ dx.doi.org/10.1029/2006JB004802.
10.1130/0091-7613(1983) 11b40:SBGAFA>2.0.CO;2. Ocola, L.C., Aldrich, L.T., Gettrust, J.F., Meyer, R.P., Ramírez, J.E., 1975. Project Nariño
Feng, M., Assumpção, M., Van der Lee, S., 2004. Group-velocity tomography and litho- I: crustal structure under southern Colombian–northern Ecuador Andes from
spheric S-velocity structure of the South American continent. Physics of the Earth seismic refraction data. Bulletin of the Seismological Society of America 65,
and Planetary Interiors 147, 315–331. 1681–1695.
Feng, M., Van der Lee, S., Assumpção, M., 2007. Upper mantle structure of South Amer- Park, Y., Nyblade, A.A., Rodgers, A.J., Al-Amri, A., 2008. S wave velocity structure of
ica from joint inversion of waveforms and fundamental-mode group velocities of the Arabian Shield upper mantle from Rayleigh wave tomography. Geochemistry,
Rayleigh waves. Journal of Geophysical Research 112, B04312 http://dx.doi.org/ Geophysics, Geosystems 9, Q07020 http://dx.doi.org/10.1029/2007GC001895.
10.1029/2006JB004449. Pavão, C.G., França, G.S., Marotta, G.S., Menezes, P.H.B.J., Neto, G.B.S., Roig, H.L.,
Ferreira, J.M., França, G.S., Vilar, C.S., do Nascimento, A.F., Bezerra, F.H.R., Assumpção, 2012. Spatial interpolation applied to crustal thickness in Brazil. Journal of
M., 2008. Induced seismicity in the Castanhão reservoir, NE Brazil — preliminary Geographic Information System 4 (2), 142–152 http://dx.doi.org/10.4236/
results. Tectonophysics 456, 103–110. jgis.2012.42019.
Fichtner, A., Igel, H., 2008. Efficient numerical surface wave propagation through the Pérez-Gussinyé, M., Lowry, A.R., Watts, A.B., 2007. Effective elastic thickness of
optimization of discrete crustal models—a technique based on non-linear disper- South America and its implications for intra-continental deformation. Geo-
sion curve matching (DCM). Geophysical Journal International 173, 519–533 chemistry, Geophysics, Geosystems 8, Q05009 http://dx.doi.org/10.1029/
http://dx.doi.org/10.1111/j.1365-246X.2008.03746.x. 2006GC001511.

Fig. 13. Data misfit. A) Example of group velocity misfit for one event in the Atlantic. Dotted line is the curve for the SA40 initial model. Solid curve is the prediction with the model
B3 (joint inversion). B) Average misfit of the up to 7000 measurements. SA40 is misfit to the initial reference model; DSP and WF are misfits to models inverted with group velocity
or waveform data only, respectively; JOINT is misfit using all data including seismic constraints. JOINT is similar to DSP, and is better than SA40 and WF. C) Example of waveform fit
for the S and Rayleigh waves of an event in Central Chile. Solid line is the observed vertical component, dotted line is the calculated trace for the SA40 initial model, and dashed line
is the calculated trace for the final joint model B3. Note better fit for the shorter periods at the end of the Rayleigh wave train. D) Distribution of all waveform fits (normalized rms fit
to the traces) for the initial reference model (SA40), model obtained with group velocity only (DSP), model obtained with waveform only (WF), and final model B3 with joint in-
version of all data sets (JOINT). Number at the top of each plot is the average misfit. JOINT fits are close to WF, and are better than DSP and the initial model SA40. The misfit for the
example in (C) is 1.91 and 0.64 for the SA40 and JOINT model, respectively.
96 M. Assumpção et al. / Tectonophysics 609 (2013) 82–96

Pérez-Gussinyé, M., Lowry, A.R., Phipps Morgan, J., Tassara, A., 2008. Effective elastic study. Journal of Geophysical Research 111, B12302 http://dx.doi.org/10.1029/
thickness variations along the Andean margin and their relationship to subduction 2005JB003769.
geometry. Geochemistry, Geophysics, Geosystems 9, Q02003 http://dx.doi.org/ Stratford, W., Thybo, H., Faleide, J.I., Olesen, O., Tryggvason, A., 2009. New Moho map
10.1029/2007GC001786. for onshore southern Norway. Geophysical Journal International 178, 1755–1765
Phillips, K., Clayton, R.W., Davis, P., Tavera, H., Guy, R., Skinner, S., Stubailo, I., Audin, L., http://dx.doi.org/10.1111/j.1365-246X.2009.04240.x.
Aguilar, V., 2012. Structure of the subduction system in southern Peru from seismic Swenson, J.L., Beck, S.L., Zandt, G., 2000. Crustal structure of the Altiplano from broad-
array data. Journal of Geophysical Research 117, B11306 http://dx.doi.org/10.1029/ band regional waveform modeling: implications for the composition of thick con-
2012JB009540. tinental crust. Journal of Geophysical Research 105, 607–621.
Ramos, V.A., 2004. La plataforma patagónica y sus relaciones con la plataforma Tanimoto, T., 1995. Crustal structure of the Earth. Global Earth Physics, A Handbook of
brasilera. In: Matess-Neto, V., et al. (Ed.), Geologia do Continente Sul-Americano: Physical Constants. In: Ahrens, T.J. (Ed.), AGU reference shelf Series, vol 1. American
Evolução e Obra de Fernando Flávio Marques de Almeida. Editora Beca, São Geophysical Union, Washington, DC, pp. 214–224 (ISBN 0-87590-851-9).
Paulo, Brazil, pp. 371–381. Tassara, A., Echaurren, A., 2012. Anatomy of the Andean subduction zone: three-
Rapela, C.W., Pankhurst, R.J., Casquet, C., Fanning, C.M., Baldo, E.G., González-Casado, dimensional density model upgraded and compared against global-scale models.
J.M., Galindo, C., Dahlquist, J., 2007. The Río de la Plata craton and the assembly Geophysical Journal International 189 (1), 161–168 http://dx.doi.org/10.1111/
of SW Gondwana. Earth-Science Reviews 83 (1–2), 49–82 http://dx.doi.org/ j.1365-246X.2012.05397.x.
10.1016/j.earscirev.2007.03.004. Tassara, A., Götze, H.-J., Schmidt, S., Hackney, R., 2006. Three-dimensional density
Ritzwoller, M.H., Shapiro, N.M., Barmin, M.P., Levshin, A.L., 2002. Global surface wave model of the Nazca plate and the Andean continental margin. Journal of Geophys-
diffraction tomography. Journal of Geophysical Research 107 (B12), 2335 http:// ical Research 111, B09404 http://dx.doi.org/10.1029/2005JB003976.
dx.doi.org/10.1029/2002JB001777. Tassara, A., Swain, C., Hackney, R., Kirby, J., 2007. Elastic thickness structure of South
Robalino, F., 1977. Espesor de la corteza en Quito mediante al análisis del espectro de America estimated using wavelets and satellite-derived gravity data. Earth and Plan-
las ondas P de periodo largo. XI Asamblea General, Instituto Panamericano de etary Science Letters 253 (1–2), 17–36 http://dx.doi.org/10.1016/j.epsl.2006.10.008.
Geografia e Historia, Quito (Abstract). Tassinari, C.C.G., Macambira, M.J.B., 1999. Geochronological provinces of the Amazo-
Rosa, M.L., Assumpção, M., Sabbione, N., 2010. Crustal structure in the eastern margin nian craton. Episodes 22 (3), 174–182.
of Argentina with receiver functions at LPA and TRQA: evidence for crustal thin- Tohver, E., Cawood, P.A., Rossello, E.A., Jourdan, F., 2012. Closure of the Clymene Ocean
ning near Río de La Plata. AGU Meeting of the Americas, Brazil. Abstract. and formation of West Gondwana in the Cambrian: evidence from the Sierras
Rosa, M.L., Collaço, B., Sánchez, G., Assumpção, M., Sabbione, N., Araujo, M., 2012. Crust- Australes of the southernmost Rio de la Plata craton, Argentina. Gondwana
al thickness beneath the Chaco–Paraná Basin. NE Argentina. Using surface-waves Research 21 (2–3), 394–405 http://dx.doi.org/10.1016/j.gr.2011.04.001.
and ambient noise tomography. XVI Congr. Geological Soc. of Peru, Lima, 23–26 van der Lee, S., Nolet, G., 1997. Upper mantle S velocity structure of North America.
September 2012, extended abstract. Journal of Geophysical Research 102, 22,815–22,838.
Sá, N.C., 2004. O campo de gravidade, o geóide e a estrutura crustal na Amé́rica do Sul: van der Lee, S., James, D., Silver, P., 2001. Upper mantle S velocity structure of
Novas estratégias de representação. Tese de Livre-Docência, Universidade de São central and western South America. Journal of Geophysical Research 106,
Paulo. 121pp. 30821–30834.
Sacek, V., Ussami, N., 2009. Reappraisal of the effective elastic thickness for the sub- van der Meijde, M., Julià, J., Assumpção, M., 2013. Gravity derived Moho for South
Andes using 3-D finite element flexural modelling, gravity and geological con- America. Tectonophysics, Special issue 609, 456–467 (submitted for publication).
straints. Geophysical Journal International 179, 778–786 http://dx.doi.org/ Wessel, P., Smith, W.H.F., 1998. New, improved version of Generic Mapping Tools re-
10.1111/j.1365-246X.2009.04334.x. leased. EOS Transactions of the American Geophysical Union 79 (47), 579.
Schmitz, M., Martins, A., Izarra, C., Jácome, M.I., Sánchez, J., Rocabado, V., 2005. The major Whitman, D., Isacks, B.L., Mahlburg Kay, S., 1996. Lithospheric structure and along-
features of the crustal structure in north-eastern Venezuela from deep wide-angle strike segmentation of the Central Andean Plateau: seismic Q, magmatism, flexure,
seismic observations and gravity modelling. Tectonophysics 399, 109–124. topography and tectonics. Tectonophysics 259 (1–3), 29–40 http://dx.doi.org/
Sieminski, A., Lévêque, J.-J., Debayle, E., 2004. Can finite-frequency effects be accounted 10.1016/0040-1951(95)00130-1.
for in ray theory surface wave tomography? Geophysical Research Letters 31, Yuan, X., Sobolev, S.V., Kind, R., 2002. Moho topography in the central Andes and its
L24614 http://dx.doi.org/10.1029/2004GL021402. geodynamic implications. Earth Planetary Science Letters 199, 389–402 http://
Smith, W.H.F., Wessel, P., 1990. Gridding with continuous curvature splines in tension. dx.doi.org/10.1016/S0012-821X(02)00589-7.
Geophysics 55 (3), 293–305. Zalán, P.V., Severino, M.C.G., Rigoti, C.A., Magnavita, L.P., Oliveira, J.A.B., Vianna, A.R.,
Snoke, J.A., James, D.E., 1997. Lithospheric structure of the Chaco and Paraná Basins of 2011. An entirely new 3D-view of the crustal and mantle structure of a South
South America from surface-wave inversion. Journal of Geophysical Research 102, Atlantic passive margin — Santos, Campos and Espírito Santo Basins, Brazil. AAPG
2939–2951. Annual Convention, Houston TX, 2011, Extended Abstracts, Search and Discovery
Soller, D.R., Ray, R.D., Brown, R.D., 1982. A new global crustal thickness map. Tectonics article #30177 (2011).
1 (2), 125–149. Zhu, L., Kanamori, H., 2000. Moho depth variation in southern California from teleseismic
Soares, J.E., Berrocal, J., Fuck, R.A., Mooney, W.D., Ventura, D.B.R., 2006. Seismic receiver functions. Journal of Geophysical Research 105 (B2), 2969–2980.
characteristics of central Brazil crust and upper mantle: a deep seismic refraction

You might also like