Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Journal of South American Earth Sciences 71 (2016) 278e295

Contents lists available at ScienceDirect

Journal of South American Earth Sciences


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

Intraplate stress field in South America from earthquake focal


mechanisms
~o a, *, Fa
Marcelo Assumpça bio L. Dias a, Ivan Zevallos b, 1, John B. Naliboff c
a ~o Paulo, Rua do Mata
IAG, Universidade de Sa ~o 1226, 05508-090, Sa ~o Paulo, SP, Brasil
b , Santar
Universidade Federal do Oeste do Para em, PA, Brasil
c
Earth and Planetary Sciences Department, University of California at Davis, CA, USA

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

Article history: We present an updated compilation of earthquake focal mechanisms in Brazil together with focal
Received 12 December 2014 mechanisms from the sub-Andean region (mainly from global CMT catalogs). All earthquakes in the sub-
Received in revised form Andean region show reverse (majority) or strike-slip faulting mechanisms. Focal mechanisms in Brazil
15 June 2016
show reverse, strike-slip and normal faulting. Focal mechanisms of nearby earthquakes in the same
Accepted 3 July 2016
Available online 5 July 2016
tectonic environment were grouped and inverted for the stress tensor. In the sub-Andean region, stresses
are compressional, as expected, with the principal major compression (S1) roughly E-W, on average. A
slight rotation of S1 can be observed and is controlled by the orientation of the Andean plateau. In the
sub-Andean region, the intermediate principal stress (S2) is also compressional (i.e., larger than the
lithostatic pressure, Sv), a feature that is not always reproduced in numerical models published in the
literature. In mid-plate South America stresses seem to vary in nature and orientation. In SE Brazil and
the Chaco-Pantanal basins, S1 tends to be oriented roughly E-W with S2 approximately equal to S3. This
stress pattern changes to purely compressional (both SHmax and Shmin larger than Sv) in the Sa ~o
Francisco craton. A rotation of SHmax from E-W to SE-NW is suggested towards the Amazon region.
Along the Atlantic margin, the regional stresses are very much affected by coastal effects (due to
continent/ocean spreading stresses as well as flexural effects from sediment load at the continental
margin). This coastal effect tends to make SHmax parallel to the coastline and Shmin (usually S3)
perpendicular to the coastline. Few breakout data and in-situ measurements are available in Brazil and
are generally consistent with the pattern derived from the earthquake focal mechanisms. Although
numerical models of global lithospheric stresses tend to reproduce the main large-scale features in most
mid-plate areas, the S1 rotation from ~E-W in SE Brazil to SE-NW in the Amazon region are not well
explained by the current numerical models. This means that the observed stress pattern in mid-plate
South America should provide new insights into upper mantle dynamics, distinct from current global
convection models.
© 2016 Elsevier Ltd. All rights reserved.

1. Introduction models of plate stresses (e.g., Mendiguren and Richter, 1978;


Coblentz and Richardson, 1996; Meijer, 1995; Meijer et al., 1997;
Mapping the stress field in intraplate regions is important to Heidbach et al., 2008 for the case of South America) showed that
help evaluate models of the plate driving forces and geodynamic the first order patterns of intraplate stresses (i.e., continental scale
properties of the upper mantle (e.g., Bird, 1998; Lithgow-Berteloni field) could be explained by four major sources: 1) ridge-push
and Guynn, 2004; Naliboff et al., 2012). Early work on the compi- forces, 2) spreading stresses from lateral density variations in the
lation of observed stress measurements, such as Assumpça ~o (1992) lithosphere (such as between the Andean plateau and the stable
for South America, Zoback (1992a) for the world and numerical South American platform, or between continental and oceanic
crusts), 3) plate interaction forces (such as in the Nazca-South
America convergent margin), and 4) lithosphere/asthenosphere
* Corresponding author. basal drag. These early compilations of intraplate stress measure-
~o).
E-mail address: marcelo@iag.usp.br (M. Assumpça ments (World Stress Map Project-WSM; Zoback, 1992a,b) indicated
1 nica, Chile.
Now at Imagineria-Eletro

http://dx.doi.org/10.1016/j.jsames.2016.07.005
0895-9811/© 2016 Elsevier Ltd. All rights reserved.
~o et al. / Journal of South American Earth Sciences 71 (2016) 278e295
M. Assumpça 279

that most plates had a predominant, long wavelength, relatively published in the literature, and c) new focal mechanisms for recent
uniform stress field, which was attributed to large-scale plate-wide Brazilian events. All compiled focal mechanism solutions are
forces. Local sources of stress were thought to have a relatively included in the last version of the WSM (Heidbach et al., 2016).
minor effect in mid-plate regions. More recent compilations They are shown in Fig. 1a as an approximate estimate of the
(Tingay et al., 2006; Heidbach et al., 2007, 2010), however, indicate orientation of the maximum horizontal compression (“SHmax”):
a higher degree of stress variability with local forces having the the P axes are plotted for reverse and strike-slip mechanisms, and
same importance as plate-wide forces. While uniform stress fields the B (Null) axes for normal faulting events. We used the criteria of
up to ~2000 km can be observed, short wavelength patterns (down Zoback (1992a) for assigning the stress regime. Fig. 1b shows all
to less than 200 km length) are just as common, even in mid-plate geographical names mentioned in this paper
regions such as Scandinavia (Heidbach et al., 2010) or some areas of All data were ranked for quality. The present WSM criteria are
the Canadian east coast (Reiter et al., 2014). based on the reliability of a focal mechanism to indicate the SHmax
While stresses arising from lateral density variations (including orientation. Because the major compressional axis (S1) can be
ridge-push) can be calculated from the known density structure of oriented anywhere in the P-wave dilatational quadrant of the focal
the lithosphere, plate collision and asthenospheric drag are more mechanism (e.g. McKenzie, 1969), the best WSM quality for a single
difficult to evaluate. The vertically averaged differential stresses solution is “C”, corresponding to SHmax uncertainty of ±20 to 25 .
from lateral density variations (spreading stresses and ridge-push) Here we adapted these criteria, and ranked the focal mechanisms as
tend to lie in the range 10e50 MPa (e.g., Turcotte and Schubert, qualities C1, C2, and D with the following criteria:
1982). Numerical modeling of intraplate stresses shows that this C1: the focal mechanism is well constrained by two indepen-
is the level of stresses at continental scales (e.g., Naliboff et al., dent data sets. For example, good quality first motion data (clear
2009, 2012). and well distributed polarities) and independent solution from
Global numerical models of intraplate stresses (e.g., Bird, 1998; waveform fitting; or good distribution of polarity data and the fault
Lithgow-Berteloni and Guynn, 2004; Naliboff et al., 2009, 2012) plane identified by the distribution of aftershocks. Identification of
automatically take into account plate interactions and lateral den- the fault plane helps restrain the possible orientation of SHmax
sity variations, but still depend on global tomography models with (Zoback, 1992b).
poorly constrained upper mantle velocities and relatively uncertain C2: Well constrained by one set of data only, such as first-motion
lithosphere thickness projections to estimate the stress contribu- polarities or waveform fitting at several stations. Fault plane not
tion from basal drag forces. In fact, comparing the calculated with usually identified.
the observed crustal stress field in mid-plate regions is a key tool to D: Just one set of data, and nodal planes not too tightly con-
investigate mantle convection patterns and rheological properties strained. For example, a regional moment tensor inversion with few
(e.g., Naliboff et al., 2009). stations, or a composite focal mechanism with uncertainties in the
In addition to constraining geodynamic models, knowledge of nodal planes between 20 and 30 .
the crustal stress field is fundamental to better understand the The use of composite focal mechanisms is regarded as a poor
causes of intraplate seismicity. Several models of intraplate seis- stress indicator, for which the WSM criteria recommend a D-quality
micity have been proposed (e.g., Mazzotti, 2007; Talwani, 2014), assignment. The majority of composite focal mechanisms compiled
which broadly include identification of crustal weak zones or here for Brazil, however, were obtained from aftershock studies
mechanisms of stress concentration in the upper crust. The using local networks. It has long been observed that intraplate
distinction between models of crustal weaknesses and models of aftershock sequences in Brazil almost always occur in the same
stress concentration can only be made by mapping the intraplate rupture plane of the mainshock (e.g., Ferreira et al., 1998;
stress field in detail. Moreover, definition of seismic zones for Chimpliganond et al., 2010; Barros et al., 2009, 2015; Carvalho
hazard analysis can benefit from maps of crustal stresses. et al., 2016). In several cases, the space distribution of the after-
Assumpça ~o et al. (2014) suggested that seismic zones in Brazil seem shocks used in the focal mechanism solution clearly identifies
to be characterized by different stress regime or orientation. which one of the nodal planes is the fault plane. For these cases we
Improved mapping of the stress field will help better delineate believe that composite mechanisms with local aftershocks can be
intraplate seismic zones. as valuable as a single solution for a large event, and can be ranked
Here we update the early compilation of intraplate focal as C-quality.
mechanism solutions in South America (Assumpça ~o, 1992), Poorly determined focal mechanisms were ranked quality “E”
including recently published focal mechanisms in Brazil. We also and were not used, but were retained in the database for
present some new, unpublished focal mechanisms of small Brazil- completeness sake.
ian earthquakes. Ferreira et al. (1998) and Assumpça ~o (1998a,b)
estimated stress orientations by inverting focal mechanisms of 2.1.1. Moment-tensor solutions
groups of earthquakes and compared with stress models of the For the sub-Andean region, Zevallos (2001) analysed 62 CMT
South American plate available at the time (Meijer, 1995; Coblentz solutions from 1990 to 1998. Different solutions (Harvard or USGS)
and Richardson, 1996; Meijer et al., 1997). Here, we updated the for the same earthquake were checked and the one most consistent
stress estimates with new groups in Brazil and in the Andean region with P-wave first motions at regional stations was chosen. The
and discuss the main patterns of the observed stress field. We also solutions were ranked for quality (D, C2 and C1), as above, ac-
compare the observed data with an updated numerical model for cording to the available independent first motion data. Quality C1
South America. was assigned to solutions where P-wave first motion data could
confirm at least one of the nodal planes (i.e., a change of
2. Focal mechanisms compressional to dilatational first motion was consistent with the
CMT solution). Quality C2 was assigned to solutions consistent with
2.1. Data compilation first-motions but no nodal plane could be checked. Quality D was
assigned when there were not enough clear first motions to test the
We updated the database used by Assumpça ~o (1992) and CMT solution. From 1999 to 2014, we used the GCMT solutions, as
Assumpça~o and Araujo (1993) mainly with a) Centroid Moment compiled by the ISC, all of them ranked as quality D (not yet
Tensor solutions compiled from the ISC, b) focal mechanisms checked for consistency with polarity data).
280 ~o et al. / Journal of South American Earth Sciences 71 (2016) 278e295
M. Assumpça

Fig. 1. a. Focal mechanism data for intraplate South America. Blue and green bars denote direction of the P axes of thrust and strike-slip events; red bars are B axes of normal
faulting events. Open bars are in-situ measurements (overcoring and hydrofrac). Bar size indicates data quality, according to Table S1 (Suplementary Material). Major geological
provinces are: GS ¼ Guyana shield, CBS ¼ Central Brazil shield, SFC ¼ S~ao Francisco craton; AmB ¼ Amazon basin, PnB ¼ Parnaiba Basin, PcB ¼ Parecis Basin, PrB ¼ Parana  Basin,
ChB ¼ Chaco Basin, Pt ¼ Pantanal Basin, TP ¼ Tocantins foldbelt Province. Thick solid lines are plate boundaries (Coffin et al., 1998): barbed green line ¼ trenches; red ¼ Mid-Atlantic
ridge; orange ¼ transform faults of the Caribbean plate and Andean Block (AB) including several “slivers”. Large and small open arrows show convergence direction of the Nazca and
the Caribbean plates, respectively, relative to South America. The brown line is the 3000 m elevation indicating the Andean plateau. The dashed line along the Atlantic coast is the
200 m bathymetry. b. Names of countries and regions mentioned in the text. PeL ¼ Pernambuco Lineament.

2.1.2. Compilation of regional studies In Peru, Tavera and Buforn (2001) determined focal mechanisms
Many regional studies have been published since the compila- for several crustal sub-Andean events by teleseismic waveform
tion of Assumpça ~o (1992) and new intraplate focal mechanisms modeling. In Ecuador and Colombia, Ego et al. (1996) and Legrand
were compiled. In the sub-Andean region above the flat-slab near et al. (2002) analysed focal mechanisms of crustal earthquakes
Sierras Pampeanas, NW Argentina, focal mechanisms determined determined by teleseismic waveform inversion (mainly HRVD so-
by regional waveform modeling indicate predominantly reverse or lutions) as well as from first motions at local network. Inversion of
strike-slip faulting, with P axes generally trending E-W or SE-NW those mechanisms showed E-W compression in most of the
(Alvarado et al., 2005; Alvarado and Beck, 2006; Alvarado and northern Andes (interpreted as due to the Nazca plate convergence)
Ramos, 2011). In the southern Puna plateau a large temporary 74- and a rotation towards NW-SE north of 5oN (interpreted as an effect
station regional network provided 17 focal mechanisms (Mulcahy of the Caribbean plate on the Andean Block). In the Central Andes,
et al., 2014), well constrained by P-wave first motions, ranging Devlin et al. (2012) determined 32 focal mechanisms of crustal
from oblique reverse, strike-slip and oblique normal faulting, earthquakes, using teleseismc P- and SH-wave modeling, and
consistent with roughly E-W oriented compression and N-S compared with InSAR data and global CMT solutions. The preferred
tension. solutions of Devlin et al. (2012) were incorporated in our database.
~o et al. / Journal of South American Earth Sciences 71 (2016) 278e295
M. Assumpça 281

Fig. 1. (continued).

Three small events (Mw ~4.5) in the Guyana shield were well magnitudes up to 3, occurred from 2005 to 2009 (Assumpça ~o et al.,
recorded by the Venezuela network (Palma et al., 2010) and had 2010). Dicelis et al. (2013) determined a composite focal mecha-
well distributed P-wave first motions showing strike-slip mecha- nism for one particular swarm of this sequence, which showed
nism with N-S oriented P axes. Interestingly, this N-S orientation in oblique normal faulting on an E-W oriented, North dipping fault
the Guyana shield seems consistent with the P axes observed plane (Fig. 1a).
further south in the Amazon Basin (Fig. 1). In Northeastern Brazil, several new focal mechanisms have been
In central Brazil, focal mechanisms determined for the northern calculated especially along the Pernambuco Lineament (Fig. 1a,b), a
part of the Parecis basin in the Amazon craton (Barros et al., 2009) major Paleozoic shear zone reactivated under the present neo-
and the Goia s-Tocantins seismic zone, between the Amazon and tectonic stresses (Ferreira et al., 2008; Lopes et al., 2010; Lima-Neto
~o Francisco cratons (Barros et al.,2015; Carvalho et al., 2016),
Sa et al., 2013, 2014), which indicates E-W compressional and N-S
indicate roughly NW-SE oriented P axes, as seen in Fig. 1. Lopes tensional deviatoric stresses. Other areas in Northeast Brazil, around
(2008), Chimpliganond et al. (2010) and Agurto et al. (2015) the Potiguar marginal basin, have been studied by França et al.
showed predominantly reverse faults in the S~ ao Francisco craton (2004), Nascimento et al. (2004), Lima-Neto et al. (2010), Oliveira
with approximately E-W oriented P axes, which suggest a rotation et al. (2010, 2015) and Dantas et al. (2011) showing a predominance
of compressional stresses from E-W in Eastern Brazil to SE-NW of strike-slip and normal faults consistent with earlier interpretation
towards the Amazon craton. of coast-parallel compressional and coast-perpendicular tensional
In the Pantanal Basin, a magnitude 4.8 event in 15-June-2009 deviatoric stresses (e.g., Ferreira et al., 1998).
had an oblique reverse faulting with ENE-WSW oriented P axis
 Basin, a
(Dias et al., 2016a,b). In the northeastern part of the Parana 2.1.3. Additional new mechanisms in Brazil
long sequence of water-well induced earthquakes, with In addition to the data above, compiled from published work,
282 ~o et al. / Journal of South American Earth Sciences 71 (2016) 278e295
M. Assumpça

we present below two examples of new focal mechanisms included the nodal planes of the waveform inversion ranks this mechanism
in this study. as quality C1. This event occurred as a predominantly reverse
faulting mechanism with E-W oriented P axis, consistent with the
2.1.3.1. Miranda (Pantanal basin), 06-Nov-2015. In November 2015 two previous mechanisms in the Pantanal Basin.
an event with magnitude 4 mb occurred in the Pantanal Basin
(West-central Brazil) and was well recorded by some regional  (Central Brazil shield), 12-July-2014.
2.1.3.2. Rio Peti, Southern Para
stations (Fig. 2). The few P-wave polarities were not enough to A sequence of small earthquakes, with maximum magnitude of ~4
constrain the focal mechanism. The Rayleigh and Love waves on 12-July-2014, occurred in the southern part of the Amazon
recorded by two regional stations (AQDB and PP1B at 134 and craton and was recorded by the recently installed stations of the
245 km distance) were then used to constrain the focal mechanism. Brazilian Seismic Network (Fig. 3a). The three closest stations
First, the group velocity dispersion to each station was used to (283e470 km) were used in a regional moment tensor inversion
estimate the velocities of the shallow layers of the upper crust (Fig. 3c) with the ISOLA code. Before doing the waveform inversion,
between epicenter and station. The focal mechanism and depth the shallow layers of the local crustal velocity model were deter-
were then varied to best fit the observed waveforms using the mined with the Rayleigh and Love group-velocity dispersion with
software ISOLA (Sokos and Zahradnik, 2008, 2013) with the the method of Dias (2016) and Dias et al. (2016b) . The strike-slip
frequency-range test developed by Dias (2016) and Dias et al. solution was consistent with the P-wave polarity data at the four
(2016b) The best moment-tensor solution is in general agreement closest stations (Fig. 3b). Relative location of six events of the
with the P-wave polarities. Several aftershocks in the range sequence, using cross-correlation of Lg-wave arrivals (Ciardelli and
3.0e3.5 mb were recorded by regional stations and a HYPODD Assumpça ~o, 2016), suggested the NNW-SSE nodal plane of the
relative location by Ciardelli and Assumnpça ~o (2016) using corre- moment-tensor solution to be the fault. However, the frequency-
lated P- and S-wave arrivals indicated a NS-trending, E dipping range test of the moment tensor inversion showed relatively
plane (dashed line in Fig. 2b). The very good agreement with one of large uncertainty (about 30 ) in the nodal plane orientations and

Fig. 2. Focal mechanism solution of the 3.8 Mw Miranda earthquake of November 6, 2015. a) Map with the epicenter (star) and the stations with first P-wave motion used in the
focal mechanism. b) fault plane solution (solid lines) from the waveform modeling and P polarities from the regional stations. Crosses and circles are compressional and dilatational
P-wave first motions, respectively. The dashed line is the fault plane defined by relocation of the aftershocks. c) Waveform modeling of the AQDB and PP1B stations (134 and 245 km
away). Black and red lines are observed and synthetic traces, respectively.
~o et al. / Journal of South American Earth Sciences 71 (2016) 278e295
M. Assumpça 283

Fig. 3. Focal mechanism of the Rio Peti event of 2014/07/12 with magnitude 3.9 Mw, in the Amazon. a) epicenter (star) and recording stations (black triangles). b) focal mechanism
solution and independent P-wave first motions, which are consistent with the solution. c) waveform fit of the Rayleigh wave (black line is the observed trace, red line the synthetic
seismogram) by regional moment-tensor inversion using the three nearest stations (283e470 km away).

this event was ranked C2 quality (See supplementary material). hydroelectric power plant (Armelin et al., 2000), and overcoring
Altogether, seven new mechanisms in Brazil were determined measurements at ~200 m depth in the Morro da Usina mine
by Dias (2016) and are included in the Supplementary materials. (Magalha ~es, 1999). We also included single breakout data for mid-
plate Brazil reported by Lima et al. (1997).
2.2. Stress patterns from single measurements A predominance of reverse and strike-slip mechanisms is clearly
seen, as typical of most intraplate regions. Normal faulting occurs
Fig. 1a shows the data set of focal mechanisms compiled here, mainly in the high Andean plateau and in some isolated areas in the
with a few additional in-situ measurements. Each colored bar Eastern part of the continent closer to the Atlantic margin. Along
represents one focal mechanism. The orientation of the bar would the Andean foreland belt P axes tend to have an average E-W
be the best estimate of the maximum horizontal stress (SHmax) if orientation and seem to be consistent for hundreds of km. How-
no other data were available. Blue and green bars are the orienta- ever, small-scale changes can be seen.
tion of the P axes of reverse and strike-slip mechanisms. Red bars
are the orientation of the B axes of normal faulting events, which 2.2.1. Andean and sub-Andean belts
are the best estimate for the SHmax orientation in this case. The In the Andean Block of Ecuador and Colombia (North of 0 lati-
earthquake data is complemented by a few in-situ measurements: tude) focal mechanisms are more varied indicating short-scale
hydro-fracturing tests in central Brazil and SE coast, and an over- features possibly caused by relative motion between crustal
coring measurement in the Caraíba copper mine NE Brazil (previ- blocks and slivers (Ego et al., 1996; Alvarado et al., 2014; Nocquet
ously compiled by Assumpça ~o, 1992). Only two new in-situ et al., 2014). Strike-slip mechanisms are more common. South of
measurements were added, both in SE Brazil: shallow hydro- 0 (Ecuador and northern Peru), reverse faulting predominates
fracturing measurement (average ~100 m depth) at the Simplício with a trend of P axes to have an average E-W orientation.
284 ~o et al. / Journal of South American Earth Sciences 71 (2016) 278e295
M. Assumpça

In the Andes, normal faults occur mainly at high elevations, as a predicted by plate-wide forces (e.g., Coblentz and Richardson, 1996;
result of lateral spreading of the Andean plateau (gravitational Naliboff et al., 2009).
collapse). In the Andean plateau of Southern Peru, Bolivia and From the Tocantins province in central Brazil to the Guyana
Northern Argentina, few focal mechanisms are available because of shield in the Amazon craton, a relatively consistent trend of N-S to
the relatively low crustal seismicity in the high Andes. The few NW-SE orientation of the P axes can be observed. Whether this is an
mechanisms (e.g., Deverche re et al., 1989) tend to have normal or evidence of a broad scale stress field, different from the E-W
strike-slip faulting in agreement with the expected extension of the domain in the sub-Andean and southern Brazil, remains to be
plateau due to gravitational collapse (e.g., Turcotte and Schubert, confirmed. The ~ N-S orientation of the P axes in the Amazon basin
1982; Meijer et al., 1997). Most P axes of strike-slip events tend to had been interpreted by Zoback and Richardson (1996) as an effect
align in the E-W orientation, especially in the Central Andes. of local flexural forces from lower crustal loads identified by an
On the other hand, almost all sub-Andean earthquakes are alignment of gravity highs along the axis of the Amazon basin.
characterized by reverse faulting with P axes trending E-W, on Similarly, flexural stresses from a SW-NE oriented lithospheric load,
average. A regional variation of the average “SHmax” orientation is due to a thinned crust, were proposed by Assumpça ~o and Sacek
apparent: E-W compression is observed in Ecuador and northern (2013) for the Tocantins province. The other focal mechanisms in
Peru, changing to SW-NE in Bolivia (between 12oS to 20oS); about the Central Brazil and Guyana shields have not been identified yet
E-W again in Jujuy and Salta provinces (22oS to 25oS); a WNW-ESE with any potential local sources of stress that could rotate the
average orientation near La Rioja (25oS to 30oS), and E-W again resulting stresses towards the NW.
south of 30oS near San Juan. The average orientations of the P axes Along the SE and Northern coast, one can observe a trend of
tend to be perpendicular to the front of the Andean plateau: lateral “SHmax” to be roughly parallel to the coastline. Fig. 5 shows the
gravitational spreading of the Andean plateau causes a local observed data for NE Brazil where the effect of the northern coast
compression in the sub-Andean belt, increasing the regional, plate- seems to influence the stress orientations more strongly than
wide compression (e.g., Assumpça ~o and Araujo, 1993; Heidbach predicted by the regional numerical model of Coblentz and
et al., 2008). Richardson (1996). Other numerical models (Meijer, 1995;
Naliboff et al., 2012) also predict an SHmax orientation changing
more gradually across continental margins than observed. These
2.2.2. Mid-plate region
models take into account the spreading stresses from lateral den-
In midplate South America (mainly Brazil) earthquakes tend to
sity variations across the continent-ocean transition, but not flex-
have reverse or strike-slip faulting. Normal faulting occurs mainly
ural stresses due to sediment load in the passive margins.
in Northeastern Brazil. Fig. 4 shows the distribution of the 76
Evidences of the importance of flexural stresses in mid-plate South
midplate focal mechanisms according to Frohlich's (1992) diagram:
America have been proposed by Ferreira et al. (1998), Assumpç~ ao
pure strike-slip (27) and reverse faults (26) account for 70% of all
(1998b), Watts et al. (2009); Assumpça ~o et al. (2011, 2014), and
data, and pure normal faults (14) for 18%. Less than 10% are clas-
Assumpça ~o and Sacek (2013), and will be discussed below.
sified as oblique faulting (strike-slip with large thrust or normal
component).
3. Stress inversions
Despite the sparse distribution of focal mechanisms, one can
observe a trend of E-W compression in SE Brazil (in the middle and
3.1. Inversion method
southern parts of the Sa~o Francisco craton), extending to the Pan-
tanal and Chaco basins. This regional E-W compression has been
The P and T axes of a focal mechanism may differ substantially
from the principal stresses S1 (maximum compression) and S3
(minimum compression) acting in the crust. The P axis, for
example, can differ up to 90 from S1 (McKenzie, 1969) depending
on the relative value of the intermediate stress (S2) and the
orientation of the fault plane (S1 can lie anywhere in the P-wave
first-motion dilatational quadrant of the nodal plane solution).
Assuming the shear stress in the fault plane has the same direction
of the slip vector, several different focal mechanisms can be used to
narrow the range of possible orientations of the principal stresses,
including the relative magnitude of the intermediate principal
stress S2, here expressed as the shape factor f ¼ (S2-S3)/(S1-S3).
Assuming that all events are subject to the same stress field (i.e., the
orientations of S1, S2, S3, and the shape factor, normally called as
“stress tensor”), several methods can be used such as those of
Angelier (1984), Gephart and Forsyth (1984), and Michael (1984,
1987). The classical method of Angelier (1984) does not distin-
guish between the two nodal planes (i.e., the results are the same
whichever nodal plane is the fault). However, the choice of which
nodal plane is the actual fault does make a difference in the possible
orientation of the principal stresses as seen in Fig. 6. Most methods
find the stress tensor that minimizes the difference between the
directions of the observed slip vector (“rake”) and the calculated
shear stress on the fault plane (e.g., Michael, 1987). This implies that
Fig. 4. Ternary diagram for classification of focal mechanism type (Frohlich, 1992). Blue the only uncertainty in the focal mechanism is the rake and that the
dots ¼ reverse mechanisms, green ¼ strike-slip, red ¼ normal faulting. Two gray dots
indicate that the orientations of the P, T, and B axes cannot be used to estimate SHmax.
fault plane orientation has no error. Gephart and Forsyth (1984), on
Data points refer to mid-plate events only. the other hand, fit the stress tensor assuming uncertainties in
strike, dip and rake of the fault mechanism. A method to solve
~o et al. / Journal of South American Earth Sciences 71 (2016) 278e295
M. Assumpça 285

Fig. 5. Stresses and free-air gravity anomalies in NE Brazil. Colored bars are “SHmax” directions estimated from focal mechanisms (as in Fig. 1a); thick black bars are SHmax inferred
from the average orientation of several breakouts in the Potiguar sedimentary basin (Reis et al., 2013). Thin black solid and open bars are SHmax/Shmin theoretical stresses
modelled by Coblentz and Richardson (1996), CR96 in the legend. Pg ¼ Potiguar basin; TB ¼ Tucano-Jatoba basin, PnB ¼ Parnaíba Basin; SFC ¼ Sa ~o Francisco craton. Gray line is the
Pernambuco Lineament (PeL). Open squares are main cities. Note the high gravity anomaly along the continental shelf between the coastline and the 200 m bathymetry, which may
indicate mass excess and possibly flexural stresses from sediment load. Free-Air anomalies from Sa  (2004).

Rivera and Cisternas (1990) and used for Ecuador by Legrand et al.
(2002), for example. A review of the main methods for stress tensor
determination was given by Maury et al. (2013).
Here we formed groups of earthquakes, as shown in Fig. 7, and
inverted the focal mechanisms for the stress tensor orientation that
best fitted the observed rakes. In the Andes, where more focal
mechanisms are available, the groups spanned smaller areas (a few
hundred km). We tried to avoid mixing events from clearly
different tectonic environment such as Andean plateau and fore-
land basins. In the mid-continent, where there are fewer focal
mechanisms, larger areas (up to 1500 km long) were necessary to
estimate the stress tensor. For each group, a grid search of all
possible orientations of S1 and S3 and the shape factor f ¼ (S2-S3)/
(S1-S3) was carried out using the code of Michael (1987), which
minimizes the misfit between the observed rake and the calculated
Fig. 6. Example of misfit between observed slip (arrow) and the calculated shear stress shear stress direction. During the grid-search, if the fault plane was
resolved in the plane (other end of the thick segment), for the two nodal planes of a not known, the two nodal planes were tested at every orientation
single mechanism. The stress tensor has S1 oriented E-W, S2 oriented NS (shape factor and the plane with the smaller difference between calculated and
f ¼ 0.5) and S3 vertical. The focal mechanism is: nodal plane P1: strike 45 , dip 55 ,
observed rakes was considered. Focal mechanisms were also
slip 120 ; plane P2: strike 180 , dip 45 , slip 55 . The stress tensor is consistent with
the rake ¼ 120 of the nodal plane 1, but has a misfit of 35 if the fault was the nodal weighted according to the quality factor assigned to the fault plane
plane 2. solution (C1, C2 or D).
Our main interest is to determine the regional patterns of the
crustal stresses at length scale of a few hundred km. Because the
simultaneously for the stress tensor and focal mechanisms (when earthquake depths (a few tens of km at most) are much smaller
only polarities are known for a cluster of events) was developed by than the size of the areas (mostly around 2 to 5 ), the average
286 ~o et al. / Journal of South American Earth Sciences 71 (2016) 278e295
M. Assumpça

Fig. 7. Selection of focal mechanisms for the stress tensor estimates. Numbers at the top edge of each area correspond to the group ID in Table 1 and Figs. 8 and 9 (for mid-plate
region) and Figs. S8 and S9 for the Andean region (Supplementary material).

stresses were constrained to be vertical and horizontal (because of respectively. The distributions of the 1000 trials of SHmax orien-
the free surface, one of the principal orientations has to be vertical). tation and shape factor are shown in Fig. 9 for mid-plate events and
The stress inversions for groups 1 to 9 (midplate South America) S9 for sub-Andean events. All results are summarized in Table 1.
are shown in Fig. 8, constrained to give horizontal and vertical The colors in Fig. 9 refer to the stress regime: blue for thrust regime,
stress orientations (groups 9 to 25 in the Andes are shown in Figs S8 green for strike-slip, and red for normal regime, according to the
and S9 of the Supplementary Materials). For example, the five focal most common regime in the bootstrap resampling. Qualities A, B,
mechanisms for group1 (San Francisco craton) resulted in an E-W and C, were assigned according to the number of mechanisms and
oriented S1, a N-S S2, and a vertical S3. The shape factor f ¼ 0.7 the misfit in the calculated rake (or the uncertainty in the SHmax
indicates the deviatoric S2 (Shmin) is 70% of the deviatoric S1 distribution of the bootstrap resamplings), following the WSM
(using the vertical stress, S3 in this case, as reference). In this criteria.
example, the small number of focal mechanisms (five) does not
qualify the stress tensor for quality A or B with the WSM criteria 3.3. Resulting stress patterns
and so it was ranked as C quality in Table 1, which shows the stress-
tensor results for all 25 selected groups. The resulting stress orientations for intraplate South America,
obtained by inverting the focal mechanisms of Fig.1a, are shown in
3.2. Assessing uncertainties Fig. 10, which also includes published results of stress tensors from
local/regional studies in Southern Colombia, Ecuador and northern
We estimated the uncertainties in the SHmax orientation, with a Peru (Ego et al., 1996; Legrand et al., 2002), in Central Peru (Dorbath
resampling bootstrap method. For each of the 25 groups, we et al., 1986; Carey-Gailhardis and Mercier, 1987; and Deverche re
created 1000 sets of focal mechanisms with the same number of et al., 1989); in the Jujuy and Salta provinces of northern
events (i.e., events can be selected more than once in any set) by nchez et al., 2012), and in the Sierra
Argentina (Cahill et al., 1992; Sa
random selection of the original group. Weights of 3, 2 and 1 were Pie de Palo and Sierras Pampeanas near San Juan, NW Argentina
applied to the focal mechanisms with qualities C1, C2, and D, (Siame et al., 2005; Richardson et al., 2012). Near the continental
~o et al. / Journal of South American Earth Sciences 71 (2016) 278e295
M. Assumpça 287

Fig. 8. Stress inversions for the nine groups of focal mechanisms in midplate South America (Groups 1 to 9 in Fig. 7). The principal stress orientations, S1 (gray diamond), S2 (solid
triangle), and S3 (open circle), were constrained to the vertical and horizontal orientations. Arrows indicate the observed slip motion in the fault plane; thick segment on the fault
plane indicates the misfit between observed and calculated rakes.

margin of SE Brazil, four composite focal mechanisms from a dam- the sub-Andean belt as well as along the Atlantic coast can be seen.
induced earthquake sequence (Mendiguren, 1980) were inverted A large-scale NW-SE oriented stress field seems to span most of the
for a stress tensor. Blue bars and yellow arrows are compressional mid-plate region across the Amazon craton.
and tensional deviatoric stresses, respectively. Fig. 10 is com-
plemented with SHmax orientations obtained by averaging several 3.3.1. Andean foreland belt
breakout measurements in a given area, shown as thick black bars The pattern of compressional stresses (compressional SHmax
(Lima et al., 1997; Reis et al., 2013, for northern Brazil, and the and Shmin) in the sub-Andean belt is confirmed as well as the
World Stress Map database for southwestern Argentina). rotation of S1 around the bend of the Andean plateau. In Colombia
The general pattern of predominantly compressional stresses (around 5oN) compression seems to be roughly parallel to the
(compressional SHmax and Shmin) can be seen with S1 oriented relative motion between the Caribbean and South American plates
roughly E-W, on average, especially along the Andean foreland and (Ego et al., 1996). Colombia and Venezuela, however has not been
in Southeastern Brazil. However, regional rotations of SHmax along sampled well enough in our study due to the complexity of the
288 ~o et al. / Journal of South American Earth Sciences 71 (2016) 278e295
M. Assumpça

Table 1
Results of the SHmax orientation and shape factor (f) from the stress-tensor inversions shown in Fig. 10. Group number refers to Fig. 7. Misfit is the average rake residual (o) in
the stress-tensor inversion (Fig. 8 and S8). Average SH and f (and standard deviations) from the boot-strap resamplings are from Fig. 9 and S9. Quality refers to the World Stress
Map criteria depending on the number of focal mechanisms in the group and the rms misfit or the standard deviation of the bootstrap. “Reg” ¼ stress regime: R ¼ thrust/reverse
(S3 Vertical), S ¼ strike-slip (S2 vertical) and N ¼ normal (S1 vertical).

Group Lat 
() Lon 
() No SHmax 
() f misfit 
() WSM Bootstrap results Reg
FMs Qual.
SH() sd f sd

1 15.06 43.53 5 80 0.7 22 C 70 38 0.51 0.20 R


2 20.92 46.24 10 68 0.6 36 B 77 19 0.48 0.34 S
3 8.36 36.07 7 86 0.8 8 C 86 2 0.71 0.15 S
4 4.02 39.51 11 114 0.3 9 B 114 2 0.31 0.12 S
5 5.38 37.62 12 112 0 10 B 115 5 0.06 0.05 R
6 5.67 36.20 13 58 0.3 5 B 58 3 0.37 0.13 S
7 19.71 60.45 10 84 0.3 4 B 84 4 0.30 0.05 R
8 11.57 52.89 8 134 0.5 15 B 134 4 0.51 0.16 S
9 2.87 62.71 6 134 0.3 12 C 140 24 0.34 0.19 R
10 10.59 61.59 14 8 0.7 50 C 18 37 0.78 0.19 N
11 10.16 69.49 13 130 0.5 4 B 128 4 0.54 0.09 R
12 5.43 73.39 22 100 0 20 B 105 12 0.24 0.20 R
13 0.33 77.82 16 82 0.1 17 A 82 4 0.15 0.11 S
14 2.31 78.24 20 100 0 13 A 100 2 0.13 0.14 R
15 6.42 76.81 24 112 0.7 16 A 101 15 0.68 0.14 R
16 10.30 74.95 29 92 0.6 14 B 90 12 0.55 0.19 R
17 11.37 73.88 19 78 0.8 7 A 77 11 0.77 0.05 R
18 14.89 72.60 26 90 0.8 19 A 88 8 0.79 0.11 S
19 16.32 71.59 27 152 0.5 37 A 151 8 0.52 0.17 N
20 17.39 65.43 9 44 0.3 9 B 45 5 0.36 0.13 S
21 23.75 64.62 14 86 0.1 14 B 87 4 0.15 0.08 S
22 26.44 67.22 19 88 0.6 25 A 88 8 0.60 0.19 S
23 27.50 66.83 33 96 0.4 26 A 102 9 0.35 0.15 R
24 31.87 68.01 54 72 0.6 17 A 78 10 0.52 0.11 R
25 35.84 70.34 11 84 0.1 19 B 83 8 0.18 0.15 R

Fig. 9. Distribution of the boostrap results for groups G1 to G9 (midplate South America). Rose diagram ¼ SHMax orientations; histogram ¼ shape factor f. Colors refer to the stress
regime: blue for thrust-faulting regime (vertical principal stress ¼ S3), green for strike-slip (vertical stress ¼ S2), and red for normal-faulting (vertical S1). The mean and standard
deviations are indicated at the top, followed by “(0)” or “(95)” which is the significance of the mean orientation. For each group the number of focal mechanisms is indicated by N at
the top right of the histogram.

various slivers in the North Andean block (e.g., Nocquet et al., 2014) In Ecuador (latitude ~0 ) S1 is roughly E-W, on average, con-
whose interaction and relative motion require a more detailed firming previous estimate by Ego et al. (1996) and Legrand et al.
analysis. (2002). However, some variation of the compressional orientation
~o et al. / Journal of South American Earth Sciences 71 (2016) 278e295
M. Assumpça 289

Fig. 10. Stress pattern in intraplate South America. Blue bars and yellow arrows (with a small circle in the middle) are stress orientations from inversion of focal mechanisms, as
selected in Fig. 7. Open bars are in-situ measurements and black solid bars are statistically significant average SHmax orientation from several breakout measurements (Lima et al.,
1997; Reis et al., 2013). The gray bar in the Parana basin is the average orientation of the six breakouts from Fig. 1a: although not statistically significant at the 90% level it is
nevertheless similar to the nearby SHmax orientations.

seems to be suggested. In the southern part of the North Andean the Andean front as indicated by Assumpç~ ao and Araujo (1993).
sliver (near 0 latitude), SHmax tends to be parallel to the Nazca South of about 30oS, where the Andean plateau gets narrow
plate convergence in agreement with crustal deformation again, the maximum compression tends to get parallel to the
measured by GNSS (Global Navigation Satellite System) data by Nazca-South America convergence, possibly because of a smaller
Nocquet et al. (2014). The stress tensors include both our estimates influence from spreading stresses of the narrow Andean plateau.
as well as those of Ego et al. (1996) and Legrand et al. (2002). In the However, both seismicity and stresses result from a complex
sub-Andean regions of southern Ecuador and northern Peru (~2 to interaction not only of plate boundary forces but also of local forces
3oS), the SHmax orientation tends to be WNW-ESE, different from related to the flat subduction in the Sierras Pampeanas region (e.g.,
the Nazca plate convergence but in agreement with measured Alvarado et al., 2005; Alvarado and Ramos, 2011).
crustal deformation in the so called “Inca Sliver” (Nocquet et al.,
2014). Clearly, the relative motion between different lithospheric 3.3.2. Andean plateau
blocks and slivers cause a short-scale variation of the stress field. Present extension in the high Andean plateau, due to gravita-
In the Bolivian sub-Andes (~18oS) S1 is oriented NE-SW, tional collapse, is well known both from focal mechanism studies as
perpendicular to the local front of the Andean plateau. Near Jujuy, well as from recent geological faulting (e.g., Mercier et al. (1992) for
Argentina (~24oS), S1 changes to roughly E-W and in La Rioja observations and Meijer et al. (1997) for numerical modeling). This
(~28oS) S1 rotates to ESE-WNW, probably as an effect of the can be seen in Fig. 10 by the N-S tensional stresses in southern Peru
spreading stresses of the Puna plateau. The gravitational spreading and in the southern Puna plateau. The regional E-W compression
of the plateau contributes to making S1 roughly perpendicular to from the Nazca plate convergence is thought to partially
290 ~o et al. / Journal of South American Earth Sciences 71 (2016) 278e295
M. Assumpça

compensate lateral E-W spreading of the plateau making N-S insufficient accuracy of the GNSS data series to determine very low
extension more pronounced. strain rates in midplate South America. Similar difficulties were
found by Marotta et al. (2015) in a detailed study of strain rates in
3.3.3. MidPlate region Northeast Brazil, with rather scattered results and no clear relation
Near the Chaco and Pantanal basins (~20oS, Figs. 1 and 9) and in with the stresses derived from the focal mechanisms.
SE Brazil (central and southern part of the Sa ~o Francisco craton)
compressional stresses trend roughly E-W, on average. S1 seems to 4. Discussion
rotate towards SE-NW across the Amazon craton. This trend is
based on very few focal mechanisms but is consistent with one The observed patterns of crustal stresses (Fig. 10) show large-
hydrofrac measurement in central Brazil and breakout data in the scale features (such as the roughly E-W compressional stresses
Amazon basin. A possible explanation for the orientation of S1 in along most of the sub-Andean region) and local rotations, such as in
the Amazon craton is the relative convergence of the North-Andean NE Brazil. This highlights the importance of local stresses that may
block and the Caribbean plate with the South-American plate. This have the same, or even larger, magnitude as the regional stresses.
source of stress was modelled by Meijer (1995) who showed that its Numerical models of the regional, plate-wide lithospheric stresses
effect could reach large areas of the Amazon. However, effects of have been calculated using different methods. We now compare
local sources of stress, such as flexural forces from lithospheric some of these theoretical estimates to help separate regional from
loads (Zoback and Richardson (1996) for the Amazon and local sources of stress.
Assumpça ~o and Sacek (2013) for central Brazil) cannot be ruled out
as perturbations of a more uniform mid-plate stress field, as dis- 4.1. Numerical models
cussed before. Other alternative explanations related to astheno-
spheric drag will be discussed below. An early numerical model of the South American plate was
In NE Brazil, a clear trend of S1 parallel to the coastline can be given by Coblentz and Richardson (1996) where a purely elastic,
observed. Also, Shmin in NE Brazil tends to be tensional perpen- 100 km thick lithospheric shell was subject to “plate boundary”
dicular to the northern coast. A detailed look at the stress pattern in forces corresponding to ridge-push (mainly from mid-Atlantic
NE Brazil (as discussed in Fig. 5) shows that S1 varies rapidly along ridge), spreading stresses from the continent/ocean transition as
the coast following closely the coastline. The breakout data shown well as from the Andean plateau, plate collision forces (especially
along the Atlantic margin (10oS to 15oS in Fig. 10) are averages of from the Nazca subduction and the Caribbean transpressional
many measurements (Lima et al., 1997); although there was a large motion) and an asthenospheric drag pushing the South American
scatter in the breakout data set, all averages are statistically sig- plate towards the NW (Fig. 12a). The model reproduced the mainly
nificant. The SHmax parallel to the coast can be interpreted as due E-W compression in mid-plate South America, with a slight trend of
to extension perpendicular to the coastline overcoming the E-W rotation towards SE-NW in the Amazon, and the tensional stresses
regional compression from plate-wide forces. This means that local in the high Andean plateau. However, differential stresses along the
sources of stress in NE Brazil, related to the continental margin, are Andean foreland had very low-magnitudes (i.e., Shmin y SHmax)
very important and are probably larger than the regional, long- and the E-W tension in the Andean plateau was not consistent with
wavelength component from plate-wide forces. This suggests a the observations of N-S tension (and E-W SHmax).
significant contribution of local flexural stresses due to sediment In that type of calculation, the spreading stresses (“ridge-push”,
load in the continental shelf as suggested by Assumpça ~o (1998a) continent/ocean transition and Andean plateau) could be estimated
and Assumpça ~o et al. (2011). directly from the topography or gravity field (with some assump-
tions regarding isostasy). On the other hand, contact forces such as
3.4. Midplate stress patterns and GPS strain rate the Nazca plate “collision” and asthenospheric drag were parame-
ters to be adjusted to best fit the observed stresses. In the model of
The SIRGAS (Geocentric Reference System for the Americas) Coblentz and Richardson (1996) the magnitude of all contact forces
network of GNSS stations has been used to estimate intraplate were chosen to best fit the available stress data of the World Stress
strain rates. In the sub-Andean region, large shortening strain rates Map Project at the time (Assumpç~ ao, 1992). An even better fit to the
(on the order of 1 x 108/y) have been observed consistent with NW-SE compressional orientation in the Amazon could be obtained
the predominantly E-W oriented compressional stresses (e.g. by increasing the magnitude of the collision with the Caribbean
Marotta et al., 2013; Nocquet et al., 2014). However, measurements plate (Meijer, 1995). While such type of modeling was helpful in
of the very low strain rates in the central part of the continent showing what kind of forces are predominant, little can be learned
(stable, midplate region) are probably below the GNSS accuracy about upper mantle convection, for example.
given the short time span between measurements. Global models of intraplate stresses have now been used where
We tried to estimate regional strain rates using the 2009 model there is no need to parameterize contact forces as they are taken
of the absolute velocity field for South America, called VEMOS2009, into account automatically through the asthenospheric drag arising
which is a smoothly varying velocity field interpolated on the from convection models (e.g. Bird, 1998; Lithgow-Bertelloni and
SIRGAS data (Drewes and Heidbach, 2012). The overall strain rate Guynn, 2004; Ghosh et al., 2008, 2009; Naliboff et al., 2009, 2012).
across the Amazon region, between Central Brazil and the Guyana While there are less free parameters to improve the fit to the
shield, is a shortening rate of 0.03  108/y, which would be observed stress data, more can be learned about the reliability of
consistent with the predominantly compressional stresses upper- and lower-mantle density models as well as global crustal
observed from focal mechanisms (Figs. 1a and 10). On the other models. Numerical models of Lithgow-Berteloni and Guynn (2004)
hand, Marotta et al. (2013) analysed the original SIRGAS velocity for global stresses showed that the best fit to the observed stresses
data in South America, using triangular sub-networks, and found in South America were achieved for a global convection model
predominantly dilatational strains in most of the stable mid-plate derived from slab history and crustal structure from CRUST2.0
region (of the order of þ0.1  108/y), even though statistically (model SLB þ TD0 of Lithgow-Berteloni and Guynn, 2004). How-
not significant in most of the mid-continent region. Predominance ever, this model predicted strike-slip stresses in mid-plate South
of dilatational strain is not consistent with the stresses observed America, not compressional stresses as we observe in Fig. 10. When
from focal mechanisms. This discrepancy is probably due to crustal densities are based on isostasy (their model TD5) a rotation
~o et al. / Journal of South American Earth Sciences 71 (2016) 278e295
M. Assumpça 291

of SHmax towards NW-SE in the Amazon was obtained. However, subducted slab which tends to sink and cause counter flow beneath
in this model the southern part of South America was poorly fit the mid-plate region of the South American plate). In the plate-
with N-S trending SHmax. driven model, upper mantle flow is constrained by the absolute
Here we improved the global model of Naliboff et al. (2009, NUVEL-1A surface plate motions (DeMets et al., 1994). The final
2012) by using CRUST1.0 to derive more accurate spreading stresses are shown in Fig. 12c. In this numerical test, stresses from
stresses from the gravitational potential energy (GPE) shown in basal tractions dominate the GPE stresses and magnitudes can
Fig. 11. These spreading stresses are mostly less than 10 MPa in mid- reach 60e65 MPa.
plate South America but can reach values as high as 25e30 MPa in
the Andean plateau and sub-Andean regions. 4.2. Comparison with observations
Stresses resulting from lithosphere basal tractions are then
added to get the total crustal stresses. Basal tractions (or When comparing the calculated stresses with the observations,
“asthenospheric drag”) are calculated with a combination of two one must take into account that small-scale features (such as the
flow fields: a plate-driven and a density-driven mantle convection complex structure of the North Andean Block) are not expected to
model (Naliboff et al., 2009). In the density-driven convection, be modelled in detail due to the lower resolution of the global
upper mantle seismic velocity anomalies (from the tomography crustal model and numerical model. For consistency, the numerical
model of Ritsema et al., 2004) are converted to density anomalies, model resolution (1ox1o) matches the resolution of Crust 1.0. The
which drive mantle flow (such as the high-velocity Nazca calculated stresses (Fig. 12c) compare reasonably well with the

Fig. 11. Principal stresses due to the gravitational potential energy (GPE), i.e., “spreading stresses” from lateral density variations, based on the CRUST1.0 model. Solid and white
lines are compressional and tensional stresses, respectively. Blue areas, generally related to high topography, indicate an extensional regime; green areas a compressional regime.
292 ~o et al. / Journal of South American Earth Sciences 71 (2016) 278e295
M. Assumpça

Fig. 12. Theoretical and observed intraplate stresses in South America. a) Stresses from regional plate-wide forces acting on a 100 km thick elastic plate (model 3 of Coblentz and
Richardson, 1996; their Fig. 6). b) Observed stresses as in Fig. 10 c) Total stresses arising from superposition of GPE stresses (Fig. 11) and tractions at the base of the lithosphere
(asthenospheric drag) arising from a combination of density-driven and plate-motion mantle convection within a global circulation model; background colors as in Fig. 11.

observed stresses (Fig. 12b) in the central and southern Andes. In 4.3. Importance of local sources of stress
the Andean plateau, the ~E-W compression and N-S tension are
well correlated. In the sub-Andean region of Ecuador and Peru ~ E- At present, global models of lithospheric stresses do not have
W oriented compressional stresses are also consistent with the sufficient resolution or sufficient regional crustal density models to
observations. Near the Bolivian orocline, the observed stresses map small-scale features of the stress field. For this reason, flexural
indicate ENE-WSW compression and a small NNW-SSE tension, stresses and other sources of local perturbations (such as major
which is also seen in the numerical model. South of ~25oS purely faults that act as weakness zones) are not commonly taken into
compressional stresses are seen again in the sub-Andean belt of account. The numerical model shown in Fig. 12c does not include
Argentina. local flexural effects. However, stresses from local sources can be as
In SE Brazil (Paran a basin and Sa ~o Francisco cratons) the important as the regional stresses (e.g., Heidbach et al., 2010;
observed ENE-WSW SHmax is reasonably well reproduced in the Assumpça ~o and Sacek, 2013).
numerical model, although the stress regime is not an exact fit. In In the mid-continent stable platform Assumpça ~o et al. (2014)
the equatorial margin (from NE Brazil to the Amazon fan) the coast showed that intraplate seismicity correlates with high values of
parallel SHmax is also well matched. free-air gravity anomalies, and proposed that flexural stresses were
The main discrepancy between the observed stresses and the an important contribution to crustal stresses. In the mid-continent,
numerical model is seen in the mid-plate region. The numerical crustal loads (evidenced by high gravity anomalies) would produce
model shows a much more uniform ENE-WSW compression all compressional stresses in the upper crust right above the load in a
along the stable continental area, whereas the observed data zone about 100e200 km wide (e.g. Assumpça ~o and Sacek, 2013). At
strongly indicates a change from ENE-WSW compression in the the continental margins, such as in NE Brazil (Fig. 5), the sediment
south to SE-NW compression all across the Central Brazil and load of the continental shelf would cause compressional stresses
Guyana shields. GPE stresses (Fig. 11) do not indicate any change in offshore (beneath the load; e.g., Assumpça ~o, 1998a) as well as
the predominant orientation of SHmax similar to the observations. tensional stresses onshore, towards the peripheral bulge. Effective
Because stresses from basal traction dominate the GPE stresses, elastic thicknesses in the coastal area of NE Brazil vary from 10 to
convection models with improved upper mantle density structures 25 km (Pe rez-Gussinie et al., 2007) which would produce the
may provide a better fit to the change in the predominant SHmax largest tensional stresses about 50e100 km from the sediment
orientation. load, consistent with the focal mechanisms shown in Fig. 5. This
It is important to note, however, that the thin-shell approxi- could help explain the observed tensional stresses perpendicular to
mation used to calculated stresses assumes full coupling between the coastline. The few focal mechanisms in Fig. 5 show that the
the entire lithosphere and basal tractions, which will not occur if reverse faulting events tend to be farther away from the coast and
rheological layering fully or partially decouples portions of the the events closer to the continental shelf tend to be strike-slip or
lithosphere (Naliboff et al., 2009). Similarly, assumptions in calcu- normal faulting, which is consistent with the proposed tensional
lating GPE regarding the lithosphere's density distribution and local stresses from flexural effects.
thickness, thin-shell/sheet approximation (i.e. vertical averaging It is interesting to note that a clear difference in the depth dis-
and no depth-dependence) and lack of flexural stresses also tribution between sub-Andean events and mid-plate events can be
introduce unrealistic deviations from observed stress patterns observed (Fig. 13). Alvarado and Ramos (2011) and Devlin et al.
(Naliboff et al., 2012). However, it is not the purpose of this paper to (2012) show that crustal events in the Andean foreland tend to
explore all the possible variations in the numerical models, but only occur in different crustal levels from about 5 to about 35 km depth,
to indicate the potential use of the South American stress data in with lower crust events more common above the flat-slab regions
future studies of global stresses and upper mantle dynamics. of Peru and Argentina, presumably due to a thicker thermal
~o et al. / Journal of South American Earth Sciences 71 (2016) 278e295
M. Assumpça 293

numerical models of intraplate stresses (e.g., Lithgow-Berteloni and


Guynn, 2004). However, local flexural effects can cause large de-
viations from the “regional” stress orientations. For this reason as
well as the fundamental assumptions of thin-sheet/shell models,
comparison of the observed stresses with results of numerical
models must be made taking into account significant potential
“distortions” of the regional field. Clearly, more focal mechanisms
in mid-plate South America are necessary to separate regional
component from more localized sources of stress.
Numerical models of global lithospheric stresses (e.g., Lithgow-
Berteloni and Guynn, 2004; Ghosh et al., 2008, 2009; Naliboff et al.,
2009, 2012) often correctly predict large-scale patterns of the stress
field, such as the general E-W orientation of SHmax in South
America (e.g., Fig. 12c), but have not been able to accurately
reproduce many observed smaller scale (<1000 km) variations of
the observed stress field. This is primarily due to the use of global
crustal models that frequently deviate from regional models, un-
certainty regarding the mantle lithosphere and asthenosphere
density structure, numerical resolution, lack of flexural stresses and
depth-averaging required by the GPE and thin sheet and shell ap-
Fig. 13. Depth distribution of intraplate crustal earthquakes: a) sub-Andean belt, data
proximations. Improved crustal and upper mantle density models
from Devlin et al. (2012); b) mid-continental region (Brazil and Paraguay). In both
cases only reliable focal depths were used, such as those from teleseismic body-wave
are clearly necessary for the South American plate, as well as 3-D
modeling, or local network hypocentral determination. Depths inferred only from lithospheric models that include vertical and lateral variations in
surface-wave modeling (such as GCMT or regional waveform inversion) were not used. rheology. The observed pattern of stress field in South America
(Figs. 1a and 10) should be useful in constraining the more accurate,
higher resolution new generation of numerical models of litho-
lithosphere. In contrast, earthquakes in the mid-continent region spheric stresses.
tend to be much shallower and concentrate in the upper crust.
Events with magnitude 5.0 mb at about 1 km depth are not un-
Acknowledgments
common (e.g., Chimpliganond et al., 2010; Barros et al., 2015). The
very shallow depths in the mid-continental area make the seis-
Work carried out with CNPq grants 30.9724/2009-0 and
micity more susceptible to flexural stresses which are higher in the
30.6547/2013-9, INCT/”Tectonic Studies” project 573713/2008-1,
brittle upper crust, close to the surface.
FAPESP scholarship 2011/20165-9 and 2014/26015-7, and Petro-
bras/Geotectonic BRASIS project. We thank Oliver Heidbach and an
5. Conclusions anonymous reviewer for detailed comments and suggestions on an
early version of the paper.
The revised and updated compilation of focal mechanisms in
intraplate South America shows that horizontal compressional Appendix A. Supplementary data
stresses predominate, not only in the Andean foreland belt, but also
in mid-plate areas of the continent. Far from the coast, the SHmax Supplementary data related to this article can be found at http://
orientation seems to be relatively uniform over continental-scale dx.doi.org/10.1016/j.jsames.2016.07.005.
regions, and two main domains could be proposed. a) Average E-
W orientation from the sub-Andean belt to SE Brazil, with small
References
local rotations observed along the Sub-Andes of Bolivia and
Argentina, caused mainly by gravitational effects of the Andean Agurto, H., Assumpça ~o, M., Ciardelli, C., Albuquerque, D.F., Barros, L.V., França, G.S.L.,
plateau, as well as near the North Andean block in northern Peru, 2015. The 2012-2013 Montes Claros earthquake series in the Sa ~o Francisco
Ecuador and Colombia, due to relative motion between small Craton, Brazil: new evidence for non-uniform intraplate stresses in mid-plate
South America. Geophys. J. Int. 200, 216e226. http://dx.doi.org/10.1093/gji/
crustal slivers, and b) Average SE-NW orientation across the whole ggu333.
Amazon craton (from the Central Brazil to the Guyana shields). Alvarado, A., Audin, L., Nocquet, J.M., Lagreulet, S., Segovia, M., Font, Y.,
While the first, ~E-W domain has been relatively well reproduced Lamarque, G., Yepes, H., Mothes, P., Rolandone, F., Jarrín, P., Quidelleur, X., 2014.
Active tectonics in Quito, Ecuador, assessed by geomorphological studies, GPS
in numerical models employing basal tractions from subduction- data, and crustal seismicity. Tectonics 33. http://dx.doi.org/10.1002/
dominated mantle convection (e.g. Fig. 12c), the Amazonian 2012TC003224.
domain has not been well explained by global models. Perhaps the Alvarado, P., Beck, S., Zandt, G., Araujo, M., Triep, E., 2005. Crustal deformation in the
south-central Andes backarc terranes as viewed from regional broad-band
effect of the thick lithosphere in the Amazon craton has not been seismic waveform modelling. Geophys. J. Int. 163, 580e598.
properly taken into account in current mantle convection models. Alvarado, P., Beck, S., 2006. Source characterization of the san Juan (Argentina)
Along the Atlantic coast, effects of the continental margin crustal earthquakes of 15 January 1944 (Mw 7.0) and 11 June 1952 (Mw 6.8).
Earth Planet. Sci. Lett. 243, 615e631.
(tensional stresses perpendicular to the coastline) strongly rotate Alvarado, P., Ramos, V.A., 2011. Earthquake deformation in the northwestern Sierras
the resulting stress field causing a trend of SHmax to be parallel to Pampeanas of Argentina based on seismic waveform modelling. J. Geodyn. 51,
the coastline. Besides gravitational effects across the continent- 205e218.
Angelier, J., 1984. Tectonic analysis of fault slip data sets. J. Geophys. Res. 89,
ocean transition, flexural stresses from sediment load in the con-
5835e5848.
tinental shelf are suggested to be equally important, such as in SE Armelin, J.L., Porfirio, N.T., Piovezani, J.D., 2000. Determinaça ~o do estado de tenso ~ es
Brazil (Assumpça ~o, 1998a; Assumpç~ ao et al., 2011) and in the in situ no maciço rochoso do AHE Simplício. In: 2o. Congresso Brasileiro de
Túneis e Estruturas Subterra ^neas, Proceedings, CD-ROM.
Amazon fan (Watts et al., 2009).
Assumpça ~o, M., 1992. The regional intraplate stress field in South America.
Flexural stresses from sediment load are usually considered a J. Geophys. Res. 97, 11.889e11.903.
localized effect and are not often taken into account in global Assumpça ~o, M., Araujo, M., 1993. Effect of the Altiplano-Puna plateau, South
294 ~o et al. / Journal of South American Earth Sciences 71 (2016) 278e295
M. Assumpça

America, on the regional intraplate stress. Tectonophysics 221, 475e496. dx.doi.org/10.1029/GL013i010p01023.


Assumpç~ ao, M., 1998a. Seismicity and stresses in the Brazilian passive margin. Bull. Drewes, H., Heidbach, O., 2012. The 2009 horizontal velocity field for South America
Seism. Soc. Am. 88 (1), 160e169. and the Caribbean. In: Kenyon, S., Pacino, M.C., Marti, U. (Eds.), Geodesy for
Assumpç~ ao, M., 1998b. Focal mechanisms of small earthquakes in SE Brazilian Planet Earth, vol. 136. IAG Symposia, pp. 657e664.
shield: a test of stress models of the South American plate. Geophys. J. Int. 133, brier, M., Lavenu, A., Yepes, H., Egues, A., 1996. Quaternary state of stress in
Ego, F., Se
490e498. the Northern Andes and the restraining bend model for the Ecuadorian Andes.
Assumpç~ ao, M., Yamabe, T.H., Barbosa, J.R., Hamza, V., Lopes, A.E.V., Balancin, L., Tectonophysics 259 (1e3), 101e116.
Bianchi, M.B., 2010. Seismic activity triggered by water wells in the Paran a Ferreira, J.M., Oliveira, R.T., Takeya, M.K., Assumpça ~o, M., 1998. Superposition of
Basin, Brazil. Water Resour. Res. 46, W07527. http://dx.doi.org/10.1029/ local and regional stresses in NE Brazil: evidence from focal mechanisms
2009WR008048. around the Potiguar marginal basin. Geophys. J. Int. 134, 341e355.
Assumpç~ ao, M., Dourado, J.C., Ribotta, L.C., Mohriak, W.U., Dias, F.L., Barbosa, J.R., Ferreira, J.M., Bezerra, F.H.R., Sousa, M.O.L., Nascimento, A.F., Sa , J.M., França, G.S.,
2011. The S~ ao Vicente earthquake of april 2008 and seismicity in the conti- 2008. The role of Precambrian mylonitic belts and present-day stress field in
nental shelf off SE Brazil: further evidence for flexural stresses. Geophys. J. Int. the coseismic reactivation of the Pernambuco lineament, Brazil. Tectonophysics
http://dx.doi.org/10.1111/j.1365-246X.2011.05198.x. 456, 111e126.
Assumpç~ ao, M., Sacek, V., 2013. Intra-plate seismicity and flexural stresses in França, G.S., Ferreira, J.M., Takeya, M.K., 2004. Seismic activity in Senador S a-CE,
Central Brazil. Geophys. Res. Lett. 40, 487e491. http://dx.doi.org/10.1002/ Brazil, 1997e1998. Braz. J. Geophys. 22, 115e125.
grl.50142, 2013. Frohlich, C., 1992. Triangle diagrams: ternary graphs to display similarity and di-
Assumpç~ ao, M., Ferreira, J., Bezerra, F.H., França, G.S., Barbosa, J.R., Menezes, E., versity of earthquake focal mechanisms. Phys. Earth Planet. Inter. 75, 193e198.
Ribotta, L.C., Pirchiner, M., Nascimento, A., Dourado, J.C., 2014. Intraplate seis- Gephart, J.W., Forsyth, D.W., 1984. An improved method for determining the
micity in Brazil. In: Talwani, P. (Ed.), Intraplate Earthquakes. Cambridge U.P., regional stress tensor using earthquake focal mechanism data: application to
ISBN 978-1-107-04038-0, 398 pp., (chapter 3), p 50e71. the San Fernando earthquake sequence. J. Geophys. Res. 89, 9305e9320.
Barros, L.V., Assumpç~ ao, M., Quinteros, R., Caixeta, D., 2009. The intraplate Porto dos Ghosh, A., Holt, W.E., Wen, L., Haines, A.J., Flesch, L.M., 2008. Joint modelling of
Gaúchos seismic zone in the Amazon craton d Brazil. Tectonophysics 469 lithosphere and mantle dynamics elucidating lithosphere-mantle coupling.
(1e4), 37e47. http://dx.doi.org/10.1016/j.tecto.2009.01.006. Geophys. Res. Lett. 35, L16309.
Barros, L.V., Assumpça ~o, M., Chimpliganond, C., Carvalho, J.M., Von Huelsen, M.G., Ghosh, A., Holt, W.E., Flesch, L.M., 2009. Contribution of gravitational potential
Caixeta, D., França, G.S., Albuquerque, D.F., Ferreira, V.M., Fontenele, D.P., 2015. energy differences to the global stress field. Geophys. J. Int. 179, 787e812.
The Mara Rosa 2010 GT-5 earthquake and its possible relationship with the Heidbach, O., Reinecker, J., Tingay, M., Müller, B., Sperner, B., Fuchs, K., Wenzel, F.,
continental-scale Transbrasiliano Lineament. J. South Am. Earth Sci. 60, 1e9. 2007. Plate boundary forces are not enough: second- and third-order stress
http://dx.doi.org/10.1016/j.jsames.2015.02.002. patterns highlighted in the World Stress Map database. Tectonics 26, TC6014.
Bird, P., 1998. Testing hypotheses on plate-driving mechanisms with global litho- http://dx.doi.org/10.1029/2007TC002133.
sphere models including topography, thermal structure and faults. J. Geophys. Heidbach, O., Iaffaldano, G., Bunge, H.P., 2008. Topography growth drives stress
Res. 103, 10,115e10,129. rotations in the central Andes: observations and models. Geophys. Res. Lett. 35,
Cahill, T., Isacks, B., Whitman, D., Chatelain, J.L., Pe rez, A., Jer Ming, C., 1992. Seis- L08301. http://dx.doi.org/10.1029/2007GL032782.
micity and tectonics in Jujuy province, northwestern Argentina. Tectonics 11, Heidbach, O., Rajabi, M., Ziegler, M., Reiter, K., the WSM Team, 2016. The World
944e959. Stress Map Database Release 2016-Global Crustal Stress Pattern Vs. Absolute
Carey-Gailhardis, E., Mercier, J.L., 1987. A numerical method for determining the Plate Motion. EGU Abstract 4861.
state of stress using focal mechanisms of earthquake populations: application Heidbach, O., Tingay, M., Barth, A., Reinecker, J., Kurfes, D., Müller, B., 2010. Global
to Tibetan teleseisms and microseismicity of Southern Peru. Earth Planet. Sci. crustal stress pattern based on the World Stress Map database release 2008.
Lett. 82, 165e179. Tectonophysics. 482, 3e15.
Carvalho, J., Barros, L.V., Zahradnik, J., 2016. Focal mechanisms and moment mag- Legrand, D., Calahorrano, A., Guillier, B., Rivera, L., Ruiz, M., Villago mez, D., Yepes, H.,
nitudes of micro-earthquakes in central Brazil by waveform inversion with 2002. Stress tensor analysis of the 1998e1999 tectonic swarm of northern Quito
quality assessment and inference of the local stress field. J. South Am. Earth Sci. related to the volcanic swarm of Guagua Pichincha volcano, Ecuador. Tecto-
http://dx.doi.org/10.1016/j.jsames.2015.07.020 (This Special Issue). nophysics 344 (1e2), 15e36.
Ciardelli, C., Assumpça ~o, M., 2016. Rupture-length x Magnitude for Intraplate Lima, C., Nascimento, E., Assumpç~ ao, M., 1997. Stress orientations in Brazilian
Earthquakes in Brazil Using Cross-correlation of Regional Lg Waves. Abstract, II sedimentary basins from breakout analysis - implications for force models in
Regional Assembly of the IASPEI's Latin American and Caribbean Seismological the South American plate. Geophys. J. Int. 130 (1), 112e124.
Commission, Costa Rica, June 2016. Lima Neto, H.C., Ferreira, J.M., Nascimento, A.F., Bezerra, F.H.R., Spineli, R.M.P.,
Chimpliganond, C., Assumpça ~o, M., von Huelsen, M., França, G.S., 2010. The intra- Costa, N.P., Menezes, E.A.S., 2010. Estudo das re plicas do sismo de magnitude
cratonic caraíbas-itacarambi earthquake of December 09, 2007 (4.9 mb), Minas 4.3 em Taipu - RN ocorrido em 2010. IV Simp. Bras. Geofísica. Brasilia-DF, 2010.
Gerais state, Brazil. Tectonophysics 480, 48e56. Extended abstract.
Coblentz, D.D., Richardson, R.M., 1996. Analysis of the South America intraplate Lima Neto, H.C., Ferreira, J.M., Bezerra, F.H.R., Assumpça ~o, M., do Nascimento, A.F.,
stress field. J. Geophys. Res. 101, 8,643e8,657. Sousa, M.O.L., Menezes, E.A.S., 2013. Upper crustal earthquake swarms in S~ ao
Coffin, M.F., Gahagan, L.M., Lawver, L.A., 1998. Present-day Plate Boundary Digital Caetano: reactivation of the Pernambuco shear zone and trending branches in
Data Compilation. University of Texas Institute for Geophysics Technical Report intraplate Brazil. Tectonophys. 608, 804e811. http://dx.doi.org/10.1016/
No. 174, pp. 5. Data available on. http://www-udc.ig.utexas.edu/external/plates/ j.tecto.2013.08.001.
data.htm (accessed 14.06.16, last update in the data in July 2012). Lima Neto, H.C., Ferreira, J.M., Bezerra, F.H.R., Assumpça ~o, M., do Nascimento, A.F.,
Dantas, R.R.S., Lima Neto, H.C., Ferreira, J.M., Nascimento, A.F., Assumpç~ ao, M., Sousa, M.O.L., Menezes, E.A.S., 2014. Earthquake sequences in the southern
Bezerra, F.H.R., Menezes, E.A.S., Spinelli, R.M.R.P., Costa, N.P., 2011. Atividade block of the Pernambuco Lineament, NE Brazil: stress field and seismotectonic
sísmica em Pedra Preta-RN em 2010. In: Congress of the Braz. Geophysical Soc., implications. Tectonophys. 633, 211e220. http://dx.doi.org/10.1016/
Rio de Janeiro, 2011 (Extended abstract). j.tecto.2014.07.010.
DeMets, C., Gordon, R., Argus, D., Stein, S., 1994. Effect of recent revisions to the Lithgow-Bertelloni, C., Guynn, J.H., 2004. Origin of the lithospheric stress field.
geomagnetic reversal time scale on estimates of current plate motions. Geo- J. Geophys. Res. 109, B01408. http://dx.doi.org/10.1029/2003JB002467.
phys. Res. Lett. 21, 2191e2194. http://dx.doi.org/10.1029/94GL02118. Lopes, A.E.V., 2008. Estudo das tenso ~ es intraplaca no Brasil. PhD thesis. IAG, Uni-
Deverche re, J., Dorbath, C., Dorbath, L., 1989. Extension related to a high topog- versity of Sa ~o Paulo, Brazil, 177 pp.
raphy: results from a microearthquake survey in the Andes of Peru and tec- Lopes, A.E.V., Assumpç~ ao, M., do Nascimento, A.F., Ferreira, J.M., Menezes, E.A.S.,
tonics implications. Geophys. J. Int. 98, 281e292. Barbosa, J.R., 2010. Intraplate earthquake swarm in Belo Jardim, NE Brazil:
Devlin, S., Isacks, B.L., Pritchard, M.E., Barnhart, W.D., Lohman, R.B., 2012. Depths reactivation of a major Neoproterozoic shear zone (Pernambuco lineament).
and focal mechanisms of crustal earthquakes in the central Andes determined Geophys. J. Int. 180, 1302e1312. http://dx.doi.org/10.1111/j.1365-
from teleseismic waveform analysis and InSAR. Tectonics 31, TC2002. http:// 246X.2009.04485.x.
dx.doi.org/10.1029/2011TC002914. Magalha ~es, F.S., 1999. Tenso~ es regionais e locais: Casos no territo  rio brasileiro e
Dias, F.L., 2016. Focal Mechanisms and the Intraplate Stress Pattern in Brazil. Ph.D. padra ~o geral. PhD thesis. Escola de Engenharia de Sa ~o Carlos, USP, 225 pp.
thesis (in Portuguese). University of Sa ~o Paulo, Dept. of Geophysics, 158 pp. Marotta, G.S., França, G.S., Monico, J.F.G., Fuck, R.A., Araújo Filho, J.O., 2013. Strain
Dias, F.L., Assumpça ~o, M., Facincani, E.M., França, G.S., Assine, M.L., Paranhos rate of the South American lithospheric plate by SIRGAS-CON geodetic obser-
Filho, A.C., Gamarra, R.M., 2016a. The 2009 earthquake, magnitude 4.8 mb, in vations. J. South Am. Earth Sci. 47, 136e141. http://dx.doi.org/10.1016/
the Pantanal Wetlands, western Brazil. Ann. Braz. Acad. Sci. (in press). j.jsames.2013.07.004.
Dias, F., Zahradnik, J., Assumpça ~o, M., 2016b. Path-specific, dispersion-based ve- Marotta, G.S., França, G.S., Monico, J.F.G., Bezerra, F.H.R., Fuck, R.A., 2015. Strain rates
locity models and moment tensors of moderate events recorded at few distant estimated by geodetic observations in the Borborema Province. Braz. J. South
stations: examples from Brazil and Greece. J. South Am. Earth Sci. This Special Am. Earth Sci. 58, 1e8.
Issue. Maury, J., Cornet, F.H., Dorbath, L., 2013. A review of methods for determining stress
Dicelis, G., Assumpça ~o, M., Prado, R., Agurto, H., Barbosa, J.R., 2013. Relocated fields from earthquake focal mechanisms: application to the Sierentz 1980
earthquakes in Bebedouro, Parana  basin, Brazil: confirmed induction by water seismic crisis (Upper Rghine graben). Bull. Soc. Geol. Fr. 184 (4e5), 319e334.
wells. In: American Geophysical Union, Fall Meeting, Abstract. Mazzotti, S., 2007. Geodynamic models for earthquake studies in intraplate North
Dorbath, C., Dorbath, L., Cisternas, A., Deverchere, J., Diament, M., Ocola, L., America. In: Stein, S., Mazzotti, S. (Eds.), Continental Intraplate Earthquakes:
Morales, M., 1986. On crustal seismicity of the Amazonian foothills of the Science, Hazard, and Policy Issues, vol. 425. Geological Society of America,
central Peruvian Andes. Geophys. Res. Lett. 13 (10), 1023e1026. http:// pp. 17e33.
~o et al. / Journal of South American Earth Sciences 71 (2016) 278e295
M. Assumpça 295

McKenzie, D.P., 1969. The relationship between fault plane solutions for earth- Geophys. Geosystems 8 (5), Q05009. http://dx.doi.org/10.1020/2006GC001511.
quakes and the directions of the principal stresses. Bull. Seism. Soc. Am. 59, Reiter, K., Heidbach, O., Schmitt, D., Haug, K., Ziegler, M., Moeck, I., 2014. A revised
591e601. crustal stress orientation database for Canada. Tectonophys. 636, 111e124.
Meijer, P.T., 1995. Dynamics of Active Continental Margins: the Andes and the Reis, A.F.C., Bezerra, F.H.R., Ferreira, J.M., Nascimento, A.F., Lima, C.C., 2013. Stress
Aegean Regions. PhD Thesis. Utrecht University, The Netherlands. magnitude and orientation in the Potiguar Basin, Brazil: implications on
Meijer, P.T., Govers, R., Wortel, M.J.R., 1997. Forces controlling the present-day state faulting style and reactivation. J. Geophys. Res. 118, 5550e5563. http://
of stress in the Andes. Earth Planet. Sci. Let. 148, 157e170. dx.doi.org/10.1002/2012JB009953.
Mendiguren, J.A., 1980. A procedure to resolve areas of different source mechanism Richardson, T., Gilbert, H., Anderson, M., Ridgway, K.D., 2012. Seismicity within the
when using the method of composite nodal plane solution. Bull. Seism. Soc. Am. actively deforming eastern Sierras Pampeanas, Argentina. Geophys. J. Int. 188,
70 (4), 985e998. 408e420. http://dx.doi.org/10.1111/j.1365-246X.2011.05283.x.
Mendiguren, J.A., Richter, F.M., 1978. On the origin of compressional intraplate Ritsema, J., van Heijst, H., Woodhouse, J., 2004. Global transition zone tomography.
stresses in South America. Phys. Earth Planet. Inter. 16, 318e326. J. Geophys. Res. 109, B02302. http://dx.doi.org/10.1029/2003JB002610.
Mercier, J.L., Sebrier, M., Lavenu, A., Cabrera, J., Bellier, O., Dumont, J.F., Machare , J., Rivera, L., Cisternas, A., 1990. Stress tensor and fault plane solutions for a population
1992. Changes in the tectonic regime above a subduction zone of Andean type: of earthquakes. Bull. Seism. Soc. Am. 80, 600e614.
the Andes of Peru and Bolivia during Pliocene-Pleistocene. J. Geophys. Res. 97, S
a, N.C., 2004. O campo de gravidade, o geo ide e a estrutura crustal na America do
11945e11982. Sul: Novas estrate gias de representaç~ ao. Tese de Livre-Doce ^ncia. Universidade
Michael, A.J., 1984. Determination of stress from slip data: faults and folds. de S~ ao Paulo, 121pp.
J. Geophys. Res. 89, 11517e11526. S
anchez, G., Araujo, M., Alvarado, P., 2012. Last main earthquakes in the northwest
Michael, A.J., 1987. The use of focal mechanisms to determine stress: a control study. of Argentina. Seismology Simposium. Peruvian Geol. Congr. extended abstr.,
J. Geophys. Res. 92, 357e368. 4pp.
Mulcahy, P., Chen, C., Kay, S.M., Brown, L.D., Isacks, B.L., Sandvol, E., Heit, B., Yuan, X., brier, M., Araujo, M., 2005. Deformation partitioning in flat
Siame, L.L., Bellier, O., Se
Coira, B.L., 2014. Central Andean mantle and crustal seismicity beneath the subduction setting: case of the Andean foreland of western Argentina
o o
Southern Puna plateau and the northern margin of the Chilean-Pampean flat (28 Se33 S). Tectonics 24, TC5003. http://dx.doi.org/10.1029/2005TC001787.
slab. Tectonics 33. http://dx.doi.org/10.1002/2013TC003393. Sokos, E.N., Zahradnik, J., 2008. ISOLA a Fortran code and a Matlab GUI to perform
Naliboff, J.B., Conrad, C.P., Lithgow-Bertelonni, C., 2009. Modification of the litho- multiple-point source inversion of seismic data. Comput. Geosci. 34 (8),
spheric stress field by lateral variations in plate-mantle coupling. Geophys. Res. 967e977.
Lett. 36, L22307. http://dx.doi.org/10.1029/2009GL040484. Sokos, E.N., Zahradnik, J., 2013. Evaluating centroid-moment-tensor uncertainty in
Naliboff, J.B., Lithgow-Bertelonni, C., Ruff, L.J., Koker, N., 2012. The effects of litho- the new version of ISOLA software. Seismol. Res. Lett. 84 (4), 656e665.
spheric thickness and density structure on Earths stress field. Geophys. J. Int. Talwani, P., 2014. Unified model for intraplate earthquakes. In: Talwani, P. (Ed.),
188, 1e17. http://dx.doi.org/10.1111/j.1365-246X.2011.05248.x. Intraplate Earthquakes. Cambridge U.P, ISBN 978-1-107-04038-0, pp. p275e302
Nascimento, A.F., Cowie, P.A., Lunn, R.J., Pearce, R.G., 2004. Spatio-temporal evolu- (chapter 11).
tion of induced seismicity at Açu reservoir, NE Brazil. Geophys. J. Int. 158, Tavera, H., Buforn, E., 2001. Source mechanism of earthquakes in Peru. J. Seismol. 5,
1041e1052. http://dx.doi.org/10.1111/j.1365-246X.2004.02351.x. 519e539.
Nocquet, J.-M., Villegas-Lanza, J.C., Chlieh, M., Mothes, P.A., Rolandone, F., Jarrin, P., Tingay, M.R.P., Müller, B., Reinecker, J., Heidbach, O., 2006. In: State and Origin of the
Cisneros, D., Alvarado, A., Audin, L., Bondoux, F., Martin, X., Font, Y., Re gnier, M., Present-day Stress Field in Sedimentary Basins: New Results from the World
Vallee, M., Tran, T., Beauval, C., Maguin~ a Mendoza, J.M., Martinez, W., Tavera, H., Stress Map Project. Paper Presented at the 41st U.S. Symposium on Rock Me-
Yepes, H., 2014. Motion of continental slivers and creeping subduction in the chanics (USRMS), 06-1049, Golden, CO.
northern Andes. Nat. Geosci. http://dx.doi.org/10.1038/NGEO2099. March 2014. Turcotte, D.L., Schubert, G., 1982. Geodynamics Applications of Continuum Physics
Oliveira, P.H.S., Ferreira, J.M., Nascimento, A.F., Bezerra, F.H.R., Soares, J.E., Fuck, R., to Geological Problems. John Wiley & Sons, Inc., ISBN 0-471-06018-6, p. 450
2010. Estudo da Sismicidade na Regia ~o de Sobral e CE, NE do Brasil, em 2008. IV Watts, A.B., Rodger, M., Peirce, C., Greenroyd, C.J., Hobbs, R.W., 2009. Seismic
Simp. Bras. Geofísica. Brasilia-DF, 2010. Extended abstract. structure, gravity anomalies, and flexure of the Amazon continental margin, ́ NE
Oliveira, P.H.S., Ferreira, J.M., Bezerra, F.H.R., Assumpç~ ao, M., Nascimento, A.F., Brazil. J. Geophys. Res. 114, B07103. http://dx.doi.org/10.1029/2008JB006259.
Sousa, M.O.L., Menezes, E.A.S., 2015. Influence of the continental margin on the Zevallos, 2001. Esforços Na Placa Continental Sul-americana. M.Sc. Dissertation. IAG,
stress field and seismicity in the intraplate Acaraú seismic zone, NE Brazil. University of Sa ~o Paulo, p. 69.
Geophys. J. Int. 202 (3), 1453e1462. http://dx.doi.org/10.1093/gji/ggv211. Zoback, M.L., 1992a. First- and second-order patterns of stress in the lithosphere:
Palma, M., Audemard, F., Romero, G., 2010. New focal mechanism solutions for the world stress map project. J. Geophys. Res. 97 (B8), 11703e11728.
Venezuela and neighbouring areas 2005-2008: importance of the National Zoback, M.L., 1992b. Stress field constraints on intraplate seismicity in eastern North
Seismological Network's density and distribution. Rev. Te c. Ing. Univ. Zulia 33 America. J. Geophys. Res. 97 (B8), 11761e11782.
(2), 108e121. Zoback, M.L., Richardson, R.M., 1996. Stress perturbation associated with the
rez-Gussinie
Pe , M., Lowry, A.R., Watts, A.B., 2007. Effective elastic thickness of South Amazonas and other ancient continental rifts. J. Geophys. Res. 101, 5459e5475.
America and its implications for intracontinental deformation. Geochem.

You might also like