Sie sind auf Seite 1von 15

See

discussions, stats, and author profiles for this publication at: http://www.researchgate.net/publication/266796712

Variations in elastic thickness and flexure of


the Maracaibo block
ARTICLE in JOURNAL OF SOUTH AMERICAN EARTH SCIENCES SEPTEMBER 2014
Impact Factor: 1.37 DOI: 10.1016/j.jsames.2014.09.014

READS

45

2 AUTHORS:
Mariano S. Arnaiz-Rodrguez

Franck A. Audemard

Central University of Venezuela

Fundacin Venezolana de Investigaciones

9 PUBLICATIONS 13 CITATIONS

182 PUBLICATIONS 1,049 CITATIONS

SEE PROFILE

SEE PROFILE

Available from: Mariano S. Arnaiz-Rodrguez


Retrieved on: 29 September 2015

Journal of South American Earth Sciences 56 (2014) 251e264

Contents lists available at ScienceDirect

Journal of South American Earth Sciences


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

Variations in elastic thickness and exure of the Maracaibo block


Mariano S. Arnaiz-Rodrguez a, *, Franck Audemard b, c
a

Departamento de Geofsica, Escuela de Geologa, Minas y Geofsica, Facultad de Ingeniera, Universidad Central de Venezuela, Venezuela
n de Investigaciones Simolo
gicas (FUNVISIS), Caracas, Venezuela
Departamento de Ciencas de la Tierra, Fundacio
c
Departamento de Geologa, Escuela de Geologa, Minas y Geofsica, Facultad de Ingeniera, Universidad Central de Venezuela, Venezuela
b

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 14 May 2014
Accepted 8 September 2014
Available online 30 September 2014

We estimate the lateral variations of the elastic thickness of the Maracaibo block with a 3D numerical
approach by using centered nite differences. The calculation is based on solving the fourth-order partial
differential equation that governs the bending of a thin plate xed on its boundaries (zero displacement)
with variable thickness (or elastic thickness for this particular case). An initial plate-load model is built
and is iteratively modied to t the general basement conguration and gravity data. The nal result is
an elastic thickness map that covers the Maracaibo block and the surrounding sections of the South
American plate. It shows that the elastic thickness ranges from 30 km to 18 km with a mean value of
23.6 km and a mode of 26 km. The largest elastic thickness values are associated with the location of the
rida AndesSanta Marta Mountains and the Barinas Apure Basin, while the smallest ones with the Me
Maracaibo Basin exural system. The current basement conguration within the Maracaibo basin,
formed as a result of its geodynamic evolution, has affected the mechanical properties of the Maracaibo
rida Andes position. The load of the Perija
 Range is compensated by a complex
block near the current Me
stress tensor, and that of the Santa Marta Mountains does not have an isostatic root as it is held by a
relatively strong lithosphere.
2014 Elsevier Ltd. All rights reserved.

Keywords:
Maracaibo block
Lithosphere exure
Finite differences
Gravity

1. Introduction
A sedimentary basin is a depressed region in the Earth surface
that has been lled by sediments (Turcotte and Schubert, 2002). A
exural basin or foreland basin is a sedimentary one formed in
response to subsidence driven by vertical stress over the elastic
lithosphere (e.g. DeCelles and Giles, 1996; Watts, 2001). These basins are characterized by: (1) a thrust front of an adjacent orogen
(or load), which is responsible for the vertical stress that bends the
plate; (2) a sediment ll with a wedge shape in transverse section;
(3) the depocenter located contiguous to the thrust belt that generates the depression (e.g. Jordan, 1995); and (4) a exural bulge, or
forebulge, that marks the end of the basin and separates it from the
undeformed craton or plate (e.g. Karner and Watts, 1983). Typical
foreland basins are divided in four discrete sections: (a) the wedgetop depozone that buries the active thrust front; (b) the foredeep
depozone formed by the subsidence driven by the load of the thrust
belt; (c) the forebulge, a region of exural uplift which is the result

* Corresponding author.
E-mail addresses: marianoarnaiz@gmail.com,
(M.S. Arnaiz-Rodrguez).
http://dx.doi.org/10.1016/j.jsames.2014.09.014
0895-9811/ 2014 Elsevier Ltd. All rights reserved.

marianoarnaiz@hotmail.com

of a damped sinusoidal deformation; and (d) the backbulge depozone, a broad region of shallow secondary exural subsidence
(DeCelles, 2012).
In Venezuela there are three foreland basins: the Eastern
Venezuela basin, the Barinas-Apure basin and the Maracaibo basin.
The two last basins are part of a double exural system driven by
rida Andes load. The Maracaibo basin (Fig. 1) is located
the Me
within an independent piece of crust known as the Maracaibo
block, while the Barinas-Apure Basin is situated in the South
American plate. These basins have been largely studied because of
their natural resources. Nonetheless little is known about the
behavior of the lithosphere in this region.
Audemard and Audemard (2002) pointed out that both, the
Maracaibo basin and the Barinas-Apure basin, have different mechanical behavior, as the depocenter in the Maracaibo basin
dwhich have a major inuence over the study aread is at least
2 km deeper that in the Barinas-Apure basin. Chacn et al. (2005)
calculated the elastic thickness for the Barinas-Apure basin in
25 km, whereas Medina (2009) afrmed that the effective elastic
thickness variations within it ranged from 30 km near the craton
rida Andes. Arnaiz(Guayana Shield) to 10 km near the Me
Rodrguez et al. (2011) considered that the elastic thickness for
the Barinas-Apure basin was around 24 2 km, and estimated that,

252

M.S. Arnaiz-Rodrguez, F. Audemard / Journal of South American Earth Sciences 56 (2014) 251e264

Fig. 1. (a) Shaded relief topography of the study area showing the major tectonic features in northwestern Venezuela. The red box represents the modeled study area. Abbreviations
rida Andes; PR, Perija
 Range; SMM, Santa Marta Mountain Range; NCA, Northern Colombia Andes; MBa, Maracaibo Basin; BABa, Barinas-Apure Basin, BF, Bocono

stand for: MA, Me
n fault. Quaternary faults from Audemard et al. (2000) (b) Structural map of the Maracaibo block. Major structures are
fault; IF, Icotea fault; SMF, Santa Marta fault; O-AF, Oca-Anco
the same as in Fig. 1a; thin gray lines represent minor faults (French and Schenk, 2004). Red dashed lines represent sediment thickness to the top of the basement in km (Di Croce,
 n et al., 2007). Blue shapes denote half-graben and basement troughs while orange shapes denote basement uplifts (Erlich
1995; Parnaud et al., 1995; Laske and Masters, 1997; Cero
et al., 1999). (For interpretation of the references to colour in this gure legend, the reader is referred to the web version of this article).

for the Maracaibo basin, it was around 16 2 km. Further, ArnaizRodrguez et al. (2011) pointed out that the formation of the Maracaibo basin and the exural system in the Maracaibo block are not
rida Andes, but also by the Perij
only controlled by the Me
a Range
and the Santa Marta Mountains, recommending that a 3D approach
to study the exure of this microplate was necessary.
Seismic anisotropy analysis suggested that SKKS split orientations at ~N45 E was most likely to be caused by lithospheric
deformation parallel to the Bocono fault (Masy et al. (2011). Curie
Point Depth analysis reported that Maracaibo block was a thermally
stable continental basin, and that a Curie Point Depth anomaly was
rida Andes, or to the graben
due to the exure produced by the Me
systems located within the Barinas-Apure basin (Arnaiz-Rodriguez
and Orihuela, 2013). The studies previously referred assert that the
Maracaibo block and the South American lithospheres behave
differently.
The present research is part of an ongoing multidisciplinary
rida Andes, its adjacent
effort to understand the dynamics of the Me
basins, structures and terranes using different approaches: GIAME
rida-Integral Geoproject (Geociencia Integral de los Andes de Me
rida Andes; Schmitz et al., 2013). It focuses upon
science of the Me
the Maracaibo block with the purpose of estimating the lateral
variations of elastic thickness of the block and its adjacent regions
by using a 3D numerical method.
2. Tectonic setting
The Maracaibo block (Fig. 1a), an independent piece of continental crust localized in northwestern Venezuela, is limited by
three fault systems (e.g., Mann and Burke, 1984; Taboada et al.,
 and Oca-Anco
 n fault sys2000; Audemard et al., 2005): Bocono
tems, both with dextral strike slip; and Santa Marta-Bucaramanga
with sinistral strike slip. Its formation and expulsion (in NNE direction relative to South America) are related to the compression

generated by the subduction of the Carnegie Ridge and the collision


 Arc against northern South America (e.g.
of the Panama
Pennington, 1981; Audemard, 1993; Kellogg and Vega, 1995). The
Maracaibo block and the Bonaire block have been overriding the
Caribbean plate, creating an ESE-dipping, amagmatic at oceanic
subduction (e.g., Kellogg and Bonini, 1982; Freymueller et al., 1993;
Van der Hilst and Mann, 1994; Kellogg and Vega, 1995; Kaniuth
et al., 1999; Taboada et al., 2000; Audemard and Audemard,
2002; Mann et al., 2006; Bezada et al., 2010). It is worth noting
that the northern Andes block, of which the Maracaibo block is a
piece, is bound to the N and NW by a complex deformation belt,
where the Caribbean and South American plates meet (Taboada
et al., 2000). Given the geometry of the northwestern corner of
South America, the Caribbean plate has been subducting in two
stages: an older one, found in the NW and W, began in the EoceneOligocene (~50 Ma ago; e.g., Kellogg and Bonini, 1982; Kellogg,
1984; Pindell and Kennan, 2009); a younger one, found in the N,
began in the Pliocene (~5 Ma ago) from the Southern Caribbean
Deformation Belt under the Bonaire block and the Maracaibo block
(Audemard, 1991; Taboada et al., 2000; Audermard and Audemard,
2002; Duerto et al., 2006; Bezada et al., 2010). Several authors have
described the older stage inuence in the Maracaibo Block geodynamic evolution (e.g., Kellogg and Bonini, 1982). However, the
stage that we refer in the text is the younger one, as is the closest to
the Maracaibo Block.
Within the Maracaibo block there are three important mountain
rida Andes, the Perija
 Range and the Santa Marta
chains: the Me
rida Andes is an over 400 km long and 40 km
Mountains. The Me
wide mountain range with a maximum elevation of 5 km, which
has no direct genetic relationship with the rest of the Andean
Range. Colletta et al. (1997) described the internal structure of the
rida Andes as a compressional positive ower structure that has
Me
been assumed to be either symmetrical (e.g. Gonz
alez de Juana,
1952) or asymmetrical (e.g. Audemard, 1991; Audemard y

M.S. Arnaiz-Rodrguez, F. Audemard / Journal of South American Earth Sciences 56 (2014) 251e264

rida Andes has also been compared to the


Audemard, 2002). The Me
Laramide-Rocky Mountains; i.e. a compressive basement block
uplift overthrusting toward the adjacent basins along blind thrust
faults (Kellogg and Bonini, 1982; De Toni and Kellogg, 1993;
 Range sits between Venezuela and
Audemard, 1991). The Perija
Colombia with a maximum elevation of 3.6 km. This mountain belt
is characterized as an ESE dipping monocline, resulting from the
reactivation of Jurassic faults during the Cenozoic (Garrity et al.,
2004; Duerto et al., 2006). The Santa Marta Mountains is a triangular shaped mountain ranged located in northern Colombia,
covering an approximate area of 3830 km2 with maximum altitude
of 5.7 km. It is usually described as an isolated uplifted massif block
of Precambrian to Mesozoic rocks that were uplifted in three pulses
from the late Maastrichtian to the Late Miocene (Cardona et al.,
2008; Ceron-Abril, 2008). These three mountains play a major
role in the exure of the lithosphere of the South American Plate
and the Maracaibo Block. Arnaiz-Rodrguez et al. (2011) proposed
rida Andes load did not uniquely control the exure of
that the Me
 Range,
the lithosphere in the Maracaibo block, and that the Perija
the Santa Marta Mountains and the Caribbean at slab impact on
the dynamic equilibrium in the region, as well as in the basement
morphology found today.
rida Andes, lays the
Adjacent to the northern foothills of the Me
Maracaibo basin, a foreland basin resulting from the loading of the
rida Andes and Perij
Me
a Range (e.g., Audemard and Audemard,
2002; Audemard, 2003). The Maracaibo basin is a small basin
with a deep asymmetric depocentre (Fig. 1b) and the apparent
absence of exural bulge (Mann et al., 2006). The great depth of this
basin (at least 9 km), in comparison with the Barinas-Apure basin
(at least 4.5 km) shows either that the Maracaibo block has a
different elastic thickness than the rest of the South American plate
(Audemard and Audemard, 2002; Arnaiz-Rodrguez et al., 2011) or
rida Andes was asymmetric
that the basement thrusting in the Me
(De Toni and Kellogg, 1993).

253

come to similar results in terms of the gravity and the isostasy in


the region (e.g., Folinsbeei, 1972; Kellogg and Bonini, 1982; Escobar
and Rodrguez, 1995; Chacn et al., 2005; Arnaiz-Rodrguez et al.,
2011). In this section, we present gravimetric maps of the Maracaibo block with a brief discussion on the signicant anomalies to
illustrate its isostatic state. The free air anomaly (Sandwell and
Smith, 2009) and the total Bouguer anomaly map (Arnaiz n, 2012) of the region are shown in Fig. 2.
Rodrguez and Garzo
The free air anomaly of the Maracaibo block ranges from
593 mGal to 149 mGal, with mean values of 16.3 mGal (Fig. 2a).
Positive values are associated with the topography of the mountain
ranges in the area, while negative values are associated with the
adjacent foreland basins. Differences in the negative free air
anomaly values at the northern and southern foothills of the
rida Andes show the discrepancy between the depocenter
Me
depths of the Maracaibo basin (9 km) and the Barinas-Apure basin
(4.5 km). This difference is associated with the lateral variations of
mechanical properties between the Maracaibo block and South
America and to the asymmetric distribution of the masses (loads)
rida Andes structure (Audemard and Audemard,
within the Me
2002; Arnaiz-Rodrguez et al., 2011). Positive free air anomaly
values show complex distribution of loads in the area. Four
rida Andes, Perij
mountain ranges load the lithosphere: the Me
a
Range, Santa Marta Mountains and the Northern Colombian Andes.
rida Andes seems to be the most signicant of these, as the
The Me
depocenter for the Maracaibo basin is immediately adjacent to
them; however, the Perij
a Range and the Northern Colombian
Andes clearly bound the basin on its western side. The deepest
section of the Barinas-Apure basin is linked to the joint contriburida Andes
tion of the Northern Colombian Andes and the Me
(Arnaiz-Rodrguez et al., 2011).
The Bouguer anomalies range from 265 mGal to 145 mGal,
with a mean of 45 mGal (Fig. 2b). The highest values are associated with the Santa Marta Mountains and Perij
a Range, indicating
that they lack local isostatic compensation (e.g. Kellogg and Bonini,
1982). Other positive anomalies are likely related to upper crust
density contrast, basement uplift (Fig. 1b), or shallow basement in
the Barinas-Apure basin and the Maracaibo basin. Watts (2001)
proposes that positive values of Bouguer anomaly often indicate

3. Gravimetric and isostatic setting


rida
Gravimetric studies of the Maracaibo basin and the Me
Andes have been carried out since the 70s, and most of them have
11
0

50

11

10

5
450 5

350
25
2000
15
0

50

50
0
5
0

30

50

10

50

100

10

50

20

0
5

40

200

10
0

50

8
0

50

10 150
0

200

50

25

00

50

50

0 00
35 3

1
0

20
0
150

100

50

100

250

50

250

50

150

1
00

00
1

10500

50

200

15
10

(a)

7
74

(b)

73

72

71

70

7
74

73

72

71

70

Gravity Anomaly
(mGal)
10050 0 50 100 150 200 250

Fig. 2. (a) Free air anomaly map of the studied region (Sandwell and Smith, 2009). (b) Complete Bouguer anomaly map of the studied region, reduced with 2.67 g/cm3 Bouguer
n, 2012). Quaternary faults from Audemard et al. (2000). Both contour maps are colored in the same color scale and contours are every
density (BA; Arnaiz-Rodrguez and Garzo
50 mGal. Gravimetric positive anomalies in the Santa Marta and Perija mountains indicate absence of isostatic compensation, while displacement to the northwest of the
rica Andes isostatic root reveals a complex regional compensation system. (For interpretation of the references to colour in this
gravimetric low that could be associated with the Me
gure legend, the reader is referred to the web version of this article).

254

M.S. Arnaiz-Rodrguez, F. Audemard / Journal of South American Earth Sciences 56 (2014) 251e264

Bandpass filter (wavelenght = 0.0105883 radians/km)

15

45 5 km (Moho depth)
10

log (Power)

17.4 2 km (Upper Crust - Lower Crust)


5
9.5 1 km (Basement/ Maracaibo Basins depocenter)
0
0.1

0.05

0.25

0.2

0.15

-5

-10

Wavenumber(Radians/km)
Fig. 3. Radially averaged power spectrum of the complete Bouguer anomaly -BA-showing the source depths estimated from the slopes of the curve. The largest wavelength
component is most likely associated with the Moho discontinuity, the mid wavelength component with the lower crusteupper crust boundary and, the shortest one, with the
basement.

rida Andes
Rodrguez et al., 2011), and not to the position of the Me
isostatic root.
Fig. 3 exhibits the radially averaged power spectrum of the
Bouguer anomaly. From the slopes of this spectrum we estimate the
depth to three mayor sources of gravimetric anomalies: the Moho,
the upper crust-lower crust boundary and the basement, using a
horizontal prism model (e.g. Spector and Grant, 1970). The longest
wavelength is associated with interfaces between 50 and 40 km,
which limit the gravimetric interpretation to crustal depth. Filtering
all but the longest wavelength with a band-pass algorithm (Fig. 4),
we produce a regional map (Fig. 4a) and a residual map (Fig. 4b).

buried loads (density contrasts in the subsurface), and thus, those


rida Andes and the Northern Colombian Andes, can
within the Me
be considered as indicative of this sort of loads. Regarding the low
Bouguer anomalies, these are associated with basin regions:
notable negative values are located over the depocenters for the
Maracaibo basin and Magdalena Basin. One of the most important
characteristics of the Bouguer anomaly map is that the gravimetric
low that characterizes an isostatically compensated mountain is
rida Andes. The
displaced over the northern foothills of the Me
negative gravimetric anomalies are most likely due to the summation of deep (Moho) and mid-depth (basement) effects (Arnaiz-

11

11

10

50

50

50

10

5
0

1
00

1
0

50

50

50

(a)

7
74

(b)

73

72

71

7
74

70

73

72

71

70

Gravity Anomaly
(mGal)

100

50

50

100

Fig. 4. (a) Regional map from ltering the longest wavelength of the complete Bouguer anomaly map showing gravimetric anomalies due to Moho and basement variations. (b)
Residual map from ltering the longest wavelengths of the complete Bouguer anomaly map showing the gravimetric anomalies due to the density contrast in the upper crust and to
some structures shown in Fig. 1b (half-graben and basement troughs). Quaternary faults from Audemard et al. (2000).

M.S. Arnaiz-Rodrguez, F. Audemard / Journal of South American Earth Sciences 56 (2014) 251e264

The regional map (Fig. 4a) presents the gravimetric contribution


of the deeper structures, particularly the Moho. The Moho depth in
the area varies from 25 km in the Santa Marta Mountains to 45 km
rida Andes northern foothills (Ceron et al., 2007). Niu et al.
in the Me
(2007) suggested that the Moho in the region has a maximum value
of 49 km, while the mean Moho value is around 42 km. Positive
values in the regional map reect the location of thinner crust as
well as the Santa Marta Mountains, which is congruent with the
values proposed by Ceron et al. (2007). Negative values
under 50 mGals most likely correspond to the extent of the
regional exure due to the load distribution. Those zones with
anomalies below 100 mGals may be related to places where the
crust and/or the sedimentary section are thicker.
The residual map (Fig. 4b) displays the gravimetric signature of
all the shallower structures such as basement uplifts, troughs and
faults. Some positive anomalies (exceptionally within the center of
rida Andes) are associated with some of the basement uplifts
the Me
in Fig. 1b, although those within the Barinas-Apure Basin do not
seem to have a clear gravimetric response. Other signicant residual positive anomalies are associated with the Perij
a Range and
the Santa Marta Mountains. This could be related to Paleozoic highdensity rocks and deformed basement present in those mountain
ranges. We do not consider here that some of these positive
anomalies represent small-buried loads, but it cannot be ruled out.
Some negative values can be linked to the locations of the basement
troughs (Fig. 1b), though not as clear as the anomalies produced by
uplifts. Other residual positive and negative values are related to
smaller structures and density contrasts that go beyond our aims.
4. Methodology

In exural studies, the lithosphere is usually represented as a 2D elastic beam that lays over a viscous medium (Watts, 2001). This
beam is then deformed by vertical stresses linked to the existence
of a vertical column of mass laying over it (thrust belts, ice caps,
sedimentary layers, etc). Assuming the absence of horizontal stress,
two models have been largely applied: the innite plate model (e.g.
Watts et al., 1985) and the broken plate model (e.g. Karner and
Watts, 1983a,b). The rst one is applied in cases where the load is
located relatively far away from a plate margin, while the second
one is applied when the loads are set near the limit of the plate. In
the rst scenario, the deformation is computed by solving the
fourth-order differential equation (Equation (1))

d4 w



pq 0

dx4
p rm wg

(1)

q rin wg
where:
w is the deection of the beam
p Winkler foundation term
q Sedimentary load term
rm Mantle density
rin Sediments density
g Gravity aceleration
D Flexural rigidity
D, in Equation (2), depends on the efcient elastic thickness (Te,
how much of the lithosphere behaves elastically), the Young
modulus (E) and the Poisson radius (y) of the beam (Watts, 2001).

ETe3


12 1  y2

(2)

These equations have been widely used to study, within a simple


approach, the behavior of the lithosphere, assuming that D and Te
are constants. When more complex situations are presented, and it
is not possible to assume Te as a constant, a numerical approach can
be used to compute the deection of a beam with variable mechanical parameters (e.g., Bodine, 1981). The problem becomes
much more intricate when the exure of the lithosphere cannot be
simplied into a 2-D elastic beam.
In a 3-D scenario, an elastic plate is used to represent the lithosphere (rather than a beam), and the deformation is computed by
solving the fourth-order partial differential equation with variable
coefcients that governs the bending of a thin plate xed on its
boundaries and variable thickness (Equation (3); Eq (3.83) in
Ventsel and Krauthammer, 2001).



vD v  2 
vD v  2 
V w 2
V w V2 D V2 w
DV2 V2 w 2
vx vx
vy vy
!
!
v2 D v2 w
v2 D v2 w v2 D v2 w


2
P
 1y
vxvy vxvy vy2 vx2
vx2 vy2

(3)

Where w represents the bending of a plate, whose thickness varies


gradually (there is no abrupt variation in thickness). P represents
the system of transverse loads applied to the plate. D is described by
Equation (4)

4.1. Mechanical background

255

ETex; y3


12 1  y2

(4)

To perform the mechanical modeling it is necessary to solve


Equation (3). At this point the boundary conditions imposed to the
equation are: (1) the boundaries of the plate are xed
(displacement 0) and are far away (at least 100 km) from the
loads, and (2) the thickness of the plate (that represents the elastic
thickness of the lithosphere) is variable but cannot vary abruptly.
Two more boundary conditions must be imposed, related to the
geologic situation at hand: (3) the plate sits over a Winkler foundation that represents the mantle, and (4) the depression after the
exure is ll with sediments. Cardozo (2009) developed a code to
solve Equation (3) by using centered nite differences, and
considering the parameters and conditions previously specied. To
compute w(x,y), the distribution of loads P(x,y) and the variations of
the elastic thickness of the plate Te(x,y) are needed. The parameters
of the mantle and inll material (rm, rin) are also required.
4.2. The Maracaibo block scenario and modeling approach
Arnaiz-Rodriguez et al. (2011) pointed out from a series of
simple 2D models that: (a) the Maracaibo block and the South
American plate cannot behave as a plate with constant elastic
thickness, (b) the Maracaibo block exure depends on the loads
distribution, and (c) elastic thickness variations must exist in the
region to explain the current basement morphology. They
concluded that is was necessary to apply 3D modeling to estimate
the elastic thickness lateral variations in the region. Considering the
mountain belt distribution (Fig. 1a), and given the fact that it is
difcult to establish the physical limit between both plates, we
chose to model the Maracaibo Block-South America interaction
region with a single continuous plate with variable elastic thickness
and xed boundaries.
Thus, we build an initial model taking into consideration the
rida Andes, Perij
area in Fig. 1a and the main loads within it (Me
a

256

M.S. Arnaiz-Rodrguez, F. Audemard / Journal of South American Earth Sciences 56 (2014) 251e264

Range, Santa Marta Mountains and Northern Colombian Andes).


These loads were initially represented from the down-sampled
topographic grid extracted from the V15 Global Topography
(Sandwell and Smith, 2009). Mechanical parameters needed for the
model (r, m E) as well as initial values of elastic thickness were
extracted from previous research; they can be seen in Fig. 5 (Chacn
et al., 2005; Medina, 2009; Arnaiz-Rodrguez et al., 2011). Once the
initial model was built, the exure was computed using Cardozo's
(2009) code. Loads and elastic thickness were iteratively modied
in small steps so that the plate would t the general basement
conguration; and the residual topography, would t the downsampled topography. A owchart illustrating this process is presented in Fig. 5.
Fig. 6 presents some steps of the modeling process: the rst is
the exure of the initial model; the second is a middle step; the
third is the nal model. Finally, we computed the gravity anomaly
of the model to compare it to the regional component of the
observed total Bouguer anomaly using the Oasis Montaj 3D GM-SYS
module (Geosoft, 2007). The residual topography and gravimetric
anomaly of the nal model are presented in Fig. 7; the resulting
elastic thickness map, in Fig. 8.
5. Results
The elastic thickness within the Maracaibo block (Fig. 8) ranges
from 30 km to 18 km, with a mean value of 23.73 km and a mode of
26 km (Fig. 9). The orientation of the elastic thickness contours is
rida Andes and the Bocono
 fault
roughly N45E, similar to the Me
System. The largest elastic thickness values (higher than 26 km) are
associated with the location of the Santa Marta Mountains, and
with the deformed Guayana Shield to the southeast (Barinas Apure
Basin). The smallest values (less than 20 km) are associated with
rida Andes-Maracaibo basin exural system. Elastic thickthe Me
rida
ness minimum values appear in the northern ank of the Me
 fault, which is congruent with the graviAndes and the Bocono
metric data of the area, where the Bouguer anomaly minimum
(that characterizes a locally compensated mountain) is displaced to
the north (Fig. 2b).
The residual topography, i.e. the height of the topographic load
after the exure, ts the real downsampled topography with relatively low deviations. The largest difference is of 172 m, which
represents 3.7% of the real topography (Fig. 7a). Fig. 6(III) presents

the exure of the mechanical model in meters. The exure of the


modeled plate is similar to the real basement conguration within
the basins. Accordingly, the model ts the morphologic data; the
error of the estimation is difcult to judge, since the modeling and
data-tting is done manually. The modeling process showed that
rida Andes, were
values in the center of the model, near the Me
more sensitive to the variations of elastic thickness and load size (as
would be expected) suggesting that the error in this area should be
small (1.0 km) due to the relatively good t of the plate conguration to the basin's basement, and the insignicant difference
between the residual topography and the real topography. Larger
errors (2.5 km) can be expected near the edges of the region
considered for the modeling (Fig. 1).
6. Discussion
The results of the mechanical model prove that the basement
conguration within the Maracaibo basin is controlled by two
rida Andes and the Perij
important load systems: the Me
a Rangerida Andes is clearly the
Northern Colombian Andes. The Me
largest load in the system as the orientation of the elastic thickness
contours is similar to the one of this mountain range. This afrmation is supported by gravimetric data, as previously discussed.
After having compared the elastic thickness gradients on both sides
rida Andes foothills, we propose that elastic thickness
of the Me
variations in the Barinas-Apure basin (from 27 to 24 km) are most
rida Andes load over the
likely due to the exure caused by the Me
relatively stable South American lithosphere (Arnaiz-Rodrguez
et al., 2011); while elastic thickness variations within the Maracaibo basin (from 24 to 18 km) are due to lithospheric weakening
caused by different processes from the Jurassic extension (and
rida
graben formation) to the present compression (uplift of the Me
Andes and convergence between the Maracaibo block and South
America; Audemard and Audemard, 2002).
rida Andes isostatic state, it is evident from
Regarding the Me
gravimetric data that local isostasy is not the compensation
mechanism that supports this range (Kellogg and Bonini, 1982;
Escobar and Rodrguez, 1995). Flexural evidence, regional gravimetric anomalies (Fig. 4a) and elastic thickness gradients adjacent
to the mountain (Fig. 8) suggest a regional compensation mechanism (e.g., Chacn et al., 2005; Arnaiz-Rodrguez et al., 2011). In a
regional isostasy scenario, we propose that the Maracaibo block

Compute gravity
of the model

Initial model

Does the mechanical model


fit geophysical data?

YES!
Initial Loads:
downs-sampled topographic grid
(Sandwell and Smith, 2009)

Mrida Andes
Perij Range
Santa Marta Mountains
Northern Colombian Andes

Parameters:
Mantle Density: 3.3 g/cm3
Sediments Density: 2.4 g/cm3
Poissons Ratio: 0.25
Youngs Modulus: 100e9 Pascal
Initial constant Te: 25 km

Does the mechanical model


fit the geology?

YES!
NO

Compare results to basement configuration


and calculate residual topography
Compute 3D Flexure of a thin plate using
centered finite differences (Cardozo, 2009)

End
modelling!
Modify
Loads
and/or Te

Input model
Fig. 5. Flowchart describing modeling approach. First, an initial model is created with assumed parameters (downsampled topography and mechanical parameters) and constant
elastic thickness (25 km). Then, the model is tested in the nite difference code (Cardozo, 2009). Flexure of the plate is compared to the basement morphology and residual
topography is compared to the real downsampled topography. The model is updated and tested until it ts the geological data. The gravimetric response of the best model is
computed and compared to the regional gravimetric anomaly; if the model roughly ts the data, the modeling is nished, if not the model is modied again.

M.S. Arnaiz-Rodrguez, F. Audemard / Journal of South American Earth Sciences 56 (2014) 251e264

Topographic
Load

257

Flexure due to Load


100
0

Elastic Thickness

500

10

10

00

100

1500

0
2000

(I)

00

20

00

1000

1500

10

00

15
2500

0
00

0
00

2500

74

72

70

2000
2500

15

1000

00

200

1500

3000

10

25
00

24

3000
20

3500

00

(II)

350
0

30
00

2500

0
400
0

2500

25

00

400

2000

24

1500

450

00

35

00

45

4500

00

20

5000

30

00

00
1

40

30

00

0
50

3500

74

72

70

200

2000
4000

10

(III)

6000

10000
80

00

8000

10000

00

40

60

74

10

0
00

12

00

60

00

Loads Height

Elastic Thickness (Te)

(km)

(km)

15

20

00

6 4 2

400

6000

4000

25

72

70

30

Fig. 6. Some steps of the exural model. From top to bottom three examples are presented: (I) is the initial model where the load is the same as the down-sample topography, Te is
25 km, the exure does not t the basement conguration. (II) is an intermediate step where the load is larger than the topography in the mountains and the same in the basins, Te
is different for SA (24e26 km) and for the MB (18e22 km), the exure has a similar shape to the basement conguration but the depth does not t. (III) is the nal model where the
topographic load is larger than the topography (dark red squares represents regions where the load is at least 3 km larger than the topography), Te gradients exist in all the area and
the exure roughly ts the depth and shape of the basement conguration. Quaternary faults from.Audemard et al. (2000). (For interpretation of the references to colour in this
gure legend, the reader is referred to the web version of this article).

rida Andes, which is the


may support the northern half of the Me
highest one, while South America supports the southern half. If we
consider that a low angle thrust is more efcient for overthusting
than shortening, and because shortening in the northern foothills
rida Andes (~40 km) is far greater than in the southern
of the Me
foothills (10e12 km; Audemard and Audemard, 2002), then we

may expect that the overthrusting in the North to be much smaller


than the one in the South. The asymmetry implied by the shortening/overthrusting relation, previously proposed by Colletta et al.
(1997), induces more uplift on the northern side than in the
southern side. Therefore the relatively weak lithosphere of the
Maracaibo block is holding a tall and narrow load that produces a

258

M.S. Arnaiz-Rodrguez, F. Audemard / Journal of South American Earth Sciences 56 (2014) 251e264

10

1
0

1
0

10

50

(b)

(a)
74

72

70

74

72

Residual Topography
(km)

70

Gravity Anomaly
(mGal)
4

100

Fig. 7. (a) Residual topography of the nal model; the largest value is 172 m, which represents 3.7% of the real topography. (b) Gravimetric anomalies due to the exure beam and
the masses of the modeled exural loads. The anomaly produced by the exure is similar to the regional gravimetric anomalies (Fig. 4a).

rida Andes northern foothills), while


large exure (9 km at the Me
the strong South American lithosphere holds a widespread load
rida Andes southern
that causes a minor exure (4.5 km at the Me
foothills). This interpretation supports the asymmetry of the
11

24

mountain chain masses, rst proposed by De Cizancourt (1933); the


asymmetric orogenic oat model anticipated by Audemard and
rida Andes load
Audemard (2002); and the unevenness of the Me
distribution over the Maracaibo Block and South America. It is also
consistent with the 20e30 dipping blind thrust and 10 km uplift of
basement rock described by De Toni and Kellogg (1993). Furthermore, Fig. 10 shows the relation between crustal thickness
(modeled from the regional Bouguer anomaly, Fig. 4a) and elastic
thickness. The region with thicker crust is related to the weakest
lithosphere, but they are also associated to the highest sections of
rida Andes, as well as the deepest sections of the Maracaibo
the Me

10

20

25

24

20

Frequency

20

24

15
10

20

5
0

7
74

73

72

71

15

20

Elastic Thickness (Te)

20

25

30

Elastic Thickness (Te)


(km)

(km)

15

25

Te (km)

70

30

15
Fig. 8. Contour map showing the lateral variations of the elastic thickness (Te) of the
MB. Largest values (>26 km) are associated with the undeformed shield to the SE and
the SMM to the NW. Small values (<20 km) area associated with the MA Northern
foothills, the MBa depocenter and the Uribante Trough. Quaternary faults from.Audemard et al. (2000).

20

25

30

Fig. 9. Frequency histogram of the elastic thicknesses in the studied region. Te values
range from 30 km to 18 km, with a mean value of 23.7 km and a mode of 26 km. The Te
values appear to be bimodal with one group ranging from 18 to 23 km and another
from 24 to 30 km.

M.S. Arnaiz-Rodrguez, F. Audemard / Journal of South American Earth Sciences 56 (2014) 251e264

259

Fig. 10. 3-D crustal image of the Maracaibo block showing, from top to bottom: topography, elastic thickness variations (Te) and Moho depth modeled from the regional gravity map
(Fig. 4a). (a) 3-D crustal view from N45E. (b) 3-D crustal view from S45W.

basin, both acting as large but narrow loads over a weakened


lithosphere with small elastic thickness.
 Range and the Northern
Other important loads are the Perija
Colombian Andes, which limits the basin on its western margin.
The Perij
a Range hardly distorts the pattern of the elastic thickness
contours, which implies that it is in isostatic equilibrium. Since
there is no gravimetric evidence (Figs. 2 and 4) that shows the
existence of an isostatic root (e.g. Kellogg and Bonini, 1982), and
elastic thickness values are not particularly large (Fig. 8), another
mechanism must be present for this equilibrium to exist. Generally,
when considering isostasy, other stresses beyond those produced
by a vertical load are not taken under consideration in the model.
One could think that horizontal stresses due to compression and
plate interaction, particularly related to the convergence between
the Maracaibo block, the Caribbean plate and South America, could
 Range in dynamic equibe enough to hold such load as the Perija
librium. Another possibility is that, given the fact that the Maracaibo block is a small plate, the large subsidence of the lithosphere
rida Andes could force the crustal block to tilt
caused by the Me
 Range and
towards the SE, uplifting the NW side. Since the Perija
the Santa Marta Mountains are on this side, they might hold the
plate from rising, and therefore be on some state of dynamic
equilibrium driven by a vertical (upward vs. downward) stress.

Most likely a blend between both cases exists, causing the Perija
Range not to have an isostatic root.
Larger elastic thickness values in the northwest, going from
26 km to 30 km, particularly near Santa Marta Mountains, would
explain why this mountain does not have sign of been isostatically
compensated (as free air and Bouguer anomalies are positive,
Fig. 2). This could be related to the convergence and coupling between the South American and the Caribbean plates. We cannot

rule out that a process similar to the one associated with the Perija
Range state could also play a signicant role in the Santa Marta
Mountains isostatic state.
The residual Bouguer anomaly map combined with the basement topography can be used to differentiate the four sections of

the Maracaibo basin (Fig. 11). The wedge-top depozone is located


rida Andes. This depozone is expected to be narrow
next to the Me
due to the relationship between overthrusting and shortening in
rida Andes northern foothill. The foredeep depozone is
the Me
present further to the NW and is characterized by a steep basement
from 9 km to 4.5 km depth. The Position of the forebulge crest
(shown in a bold line in Fig. 11) is indicated by a cluster of 4.5 km
contours and by a positive residual Bouguer anomaly. The forebulge
of the basin has not been located before, and was thought to be
absent (Mann et al., 2006; Arnaiz-Rodrguez et al., 2011). Because
this particular forebulge shows no topographic expression (is
buried by 4 km of sediments), the Maracaibo basin is in an overlled state (DeCelles, 2012). Furthermore, the relative proximity
between the forebulge and the thrust front supports low elastic
thickness values found within the basin (<20 km). Lastly, the
backbulge depozone is found between the forebulge and the Perij
a
Range. It could also be considered as the wedge-top and foredeep
 Range thrust-belt.
depozones of the Perija
The 20 km contour on the elastic thickness map has a similar
 fault system suggesting that the crustal
orientation than Bocono
and mechanical limit between the Maracaibo block and South
America is in some way associated with this structure, even though
it is improbable the fault displaces the Moho, as shown by crustal
rida Andes (Monod et al., 2010)
scale balanced sections of the Me
and in a similar mountain range (Laramide Wind River Range;
rida Andes
Smithson et al., 1987). It has been suggested that the Me
formation is related to lithospheric deformation parallel to the
 fault (Masy et al., 2011). Upper Mantle ow parallel to the
Bocono
rida Andes does not necessarily reect a Bocono
 fault of lithoMe
spheric scale, nor it reects mantle ow inuence in the formation
of this mountain range. It could be due to the presence of an
incipient continental slab that pushes the lithospheric mantle
down toward the asthenosphere, as suggested by the results of
Burgos et al. (2011) that show a thicker lithosphere (from 70 km in
the Barinas-Apure basin to 80 km beneath the Maracaibo block) in
the region, where a deep Moho is expected (Fig. 10). Interestingly,

260

M.S. Arnaiz-Rodrguez, F. Audemard / Journal of South American Earth Sciences 56 (2014) 251e264

11

3.0

E
4.5

3.0

BA

ON

4.5

10

LG

U
KB

DE

NE

O
OZ

D
RE

FO
6.0

1.0

P
EE

EP

D
OP
-T

7.5

2.0

Z
PO

4.5

FOREBULGE

3.0

GE

ED

3.0

4.5

1.5

7.5

7
74

73

72

6.0

71

4.5

70

Gravity Anomaly
(mGal)

100

50

50

100

Fig. 11. Forebuldge position and depozones of the Maracaibo basin on top of the residual Bouguer anomaly and major structures in the study area. The forebulge is located within a
set of 4.5 km contours and associated to some positive residual anomalies within the basin.

the Icotea fault, a left-lateral strike slip fault, also distorts the
pattern of the elastic thickness contours in the direction of its
displacement. Therefore, the Icotea fault has, at least, a crustal inuence, as suggested by seismicity in its vicinities, with events up
to 40 km deep (Audemard and Audemard, 2002). Moreover, this
distortion has the appearance of continuing to the southeast, which
 fault
is compatible with the idea of its convergence with the Bocono
n, 1994), or at least that it does not end in
at some point (e.g. Beltra
the Maracaibo Lake area as proposed by Castillo and Mann (2006).
As the loads within the Maracaibo block have different ages, we
will briey discuss the time dependent exure of the lithosphere.
Based on the viscoelastic plate model (Walcott, 1970), two characteristics of the load are important: the age and the width (Watts,
2001). Young loads are mostly correlated to high values of exural
rigidity, while older ones tend to produce lower exural rigidity
values. Wide loads cause the lithosphere to approach faster to a
hydrostatic state (Airy's isostasy model) than a narrow load. The
rida Andes, within the Maracaibo block exural system, can be
Me
considered as a relatively young and narrow load that produces a
short wavelength and deep exure over the lithosphere. Such circumstances would suggest that the instantaneous exural rigidity
of the plate (and therefore its elastic thickness) should be less than
the standards values for continental lithosphere. In fact, there is no
simple relationship between instantaneous exural rigidity, the
elastic thickness variations and the age of a load (Watts, 2001), so
we cannot directly determine how much of the subsidence and
rida Andes load,
elastic thickness variations are produce by the Me
nor how much is inherited from previous process.

6.1. Flexural history of the Maracaibo Block


Based on the results of this research and on previous interpretations of the geodynamic evolution of the Maracaibo Block,
its orogens and basins, we present a schematic portrayal of the
geodynamic history of the region (Fig. 12), with particular emphasis
on the different exural stages deforming this microplate:
a. Originally, the Maracaibo block was part of the South American
plate. This lithosphere probably had a relatively large and constant elastic thickness (>30 km), as suggested by regional exural studies about South America (e.g. Watts et al., 1995; Stewart
rez-Gussinye, 2007).
and Watts, 1997; Pe
b. In the Late Jurassic, rifting between North and South America
created the Proto-Caribbean seaway (e.g. Pindell and Barrett,
1990), as well as a passive margin along northern South America. This process would have reduced the elastic thickness towards the divergent margin. Eventually, extension of the
lithosphere created a series of grabens and half grabens in the
South American crust (Parnaud et al., 1995); their formation
would have weakened the lithosphere and signicantly reduced
the elastic thickness near these structures, as suggested by
Audemard and Audemard (2002).
c. During the Cretaceous, sediments were deposited over the
continental platform causing subsidence in the lithosphere
(Duerto, 1998). In the late Maastrichtian the uplift of the Santa
Marta Mountains began, which might have affected the elastic
thickness in an uncertain way. During this period, thermal

M.S. Arnaiz-Rodrguez, F. Audemard / Journal of South American Earth Sciences 56 (2014) 251e264

261

Fig. 12. Geodynamic evolution of the MB, its orogens and basins, based on the reconstruction proposed by different authors, see text for details. The dotted red line shows the
variation of the elastic thickness (Te) through time (not at true vertical scale). MA current structure is based on Arnaiz-Rodrguez et al. (2011) and Monod et al. (2010) models. The
age of the stages described are as follow: (a) Pre-Jurassic, (b) Late Jurassic, (c) Cretaceous, (d) Paleocene, (e) Oligocene, (f) Middle Moicene, (g) Pliocene. In Fig. 12g, S stands for
shortening and Ot for overthusting. Thick arrows show the direction of stress (either compression or extension) and the small arrow marks the forebulge. (For interpretation of the
references to colour in this gure legend, the reader is referred to the web version of this article).

equilibrium of the lithosphere drove thermal subsidence in the


Maracaibo Block, similar to the process described in the Eastern
Cordillera in Colombia (e.g. Sarmiento, 2002), certainly reducing
the elastic thickness.
d. The rst exural deformation stage of the Maracaibo basin was
associated with the collision of the The Great Caribbean Arc
with northern Venezuela from the Paleocene to the early Eocene
(Lugo and Mann, 1995). Throughout this period, some portions
of The Great Caribbean Arc collided and overthrust the passive

margin. Shortening related to this process led to the Lara nappe


emplacement (Stephan, 1985) that caused subsidence in
northwestern Venezuela, reducing the elastic thickness in
northern South America, similar to the lithospheric weakening
produced by nappe emplacement described in the East Carpathians (e.g. Artyushkov et al., 1996).
e The Oligocene represents an important orogenic stage because
 Range, as
of the uplift of the Colombian Andes and the Perija
well as a second pulse of the Santa Marta Mountains uplift. The

262

M.S. Arnaiz-Rodrguez, F. Audemard / Journal of South American Earth Sciences 56 (2014) 251e264

rst and second were related to the Nazca Plate subduction,


while the third to the at subduction of the Caribbean Plate
(Kellogg, 1984; Van der Hilst and Mann, 1994; Taboada et al.,
2000). Flat subduction of the Caribbean plate that started in
the NW at this period might have help support the load in the
 Range and the Santa Marta
Maracaibo block along the Perija
Mountains. It is worth noting that the crustal structure of the
 Range is not well known, therefore, in our reconstruction
Perija
we take the one proposed by Audemard and Audemard (2002).
f In the Middle Miocene, stress produced by the Panama Arc
collision with northern South America forced the inversion of a
rida Andes (e.g.
Jurassic graben that led to the uplift of the Me
Audemard and Audemard, 2002; Monod et al., 2010) and drove
the exural subsidence of the region, as well as the creation of
rida
the large depocenter on the northern foothills of the Me
Andes (Audemard, 2003). The formation of this foreland basin is
recorded by normal faults within it with an average trend of
rida
S37E (Castillo and Mann, 2006). The large load of the Me
Andes might have reduced the elastic thickness in the Maracaibo basin and the Barinas-Apure basin through its uplift to
some extent, and formed the incipient forebulge of the Maracaibo basin.
g. Ultimately, in the last 5 Ma, the compression generated by the
Panama arc collision and the subduction of the Carnegie Ridge at
the Ecuador trench in northern South America produced the
escape of the Maracaibo block and the northern Andes and the
Bonaire block (Egbue and Kellogg, 2010). As both overrode the
Caribbean Plate, a south-dipping amagmatic at oceanic subduction was created in the Southern Caribbean deformation belt
rida Andes
(e.g Audermard, 2009). The current uplift of the Me
and the Perij
a Range is driven by oblique convergence and
resulting transpression between South America and the Maracaibo block (Audemard and Audemard, 2002) affecting the
exural system in an uncertain way.
7. Conclusions
Numerical modeling of the complex load system within the
Maracaibo block has allowed us to estimate the lateral variations of
the exural thickness in the region. Based on the elastic thickness
variations, we can draw the following conclusions:
1. The use of a 3D numerical approach is valid to roughly estimate
the variations of the elastic thickness of the continental lithosphere. This method is applicable as long as the boundary
conditions and limitations expressed by the equations are
respected, and satisfy the overall geodynamic setting.
2. The elastic thickness in the study area ranges from 30 km to
18 km, with a mean value of 23.7 km, a mode of 26 km. The
orientation of the elastic thickness contours is roughly N45E,
rida Andes, indicating that this is the most
similar to the Me
important load within the Maracaibo block. Large elastic thickness values (higher than 26 km) are associated with the location
of the Santa Marta Mountains and with the deformed Guayana
Shield. The smallest values (less than 20 km) coincide with the
rida Andes-Maracaibo basin exural system. Estimated erMe
rors range from 1.0 km to 2.5 km.
3. The basement conguration within the Maracaibo basin seems
rida Andes (which is clearly the
to be controlled by the Me
largest load) and by the Perij
a Range-Northern Colombian
Andes (which limits the basin on its western margin).
4. The elastic thickness map shows that the 20 km contour has a
 fault system; this could
similar orientation than the Bocono
imply that the mechanical and geodynamic limit between the
Maracaibo block and South America is in some way associated

5.

6.

7.

8.

with this structure, even though it is improbable the fault displaces the Moho.
 Range barely distorts the pattern of the elastic
The Perija
thickness contours and lacks an isostatic root. There may be two
possible explanations for this: (a) horizontal stresses due to
 Range load, or
compression and plate interaction hold the Perija
rida
(b) the large subsidence of the lithosphere caused by the Me
Andes could force the Maracaibo Block to tilt towards the SE;
this would have caused the west side of the block to be uplifted
 Range and the Santa Marta Mountains prevent it
but the Perija
from rising. A mixture of both should not be discarded.
Contiguous 4.5 km contours and positive residual Bouguer
anomaly within the basin indicate the forebulge crest of the
Maracaibo basin. It shows no topographic expression as it is
buried by 4 km of sediments, which implies that the basin is in
an overlled state. Moreover, the distance between the forebulge and the northern thrust front support low elastic thickness values found within the basin.
The Santa Marta Mountains region has larger elastic thickness
values (from 26 km to 30 km). This could be related to the
convergence of South America and the Caribbean plate (CP) and
the coupling related to this process. The scenarios proposed for
 Range might play a role on the Santa Marta Mountains
the Perija
isostatic equilibrium as well. Moreover, the fact that the Santa
Marta Mountains was uplifted far from the region affected by
graben formation suggests that elastic thickness values in this
region were unaffected by pre-orogenic processes.
When looking at the full picture of the geodynamic evolution of
rida
the Maracaibo block, it is clear that, even though the Me
Andes is the most important load in the system, its orogenesis is
not the only process that produced the current elastic thickness
gradients within it. Consequently, the Te values within the
Maracaibo basin (from 24 to 18 km) are likely the response to
different stages in the Maracaibo block history. Particularly
Jurassic extension could have affected and weakened the lithosphere. Subsequently, the uplift and overthrusting of the
rida Andes over the weak Maracaibo block lithosphere proMe
duced the deep Maracaibo block.

Further work
As noted previously in this paper, classic exural studies take
into consideration vertical stresses related to the loads over the
lithosphere. Horizontal (either compression or extension) stresses
are not, but are often mentioned in the interpretation of the results obtained. As the Maracaibo block represents a region with a
relatively complex stress eld, compressional stress must be
considered. Further work will include modeling the geodynamic
situation with the nite element method so the horizontal stress
eld can be taken into account, as well as viscoelastic
deformation.
Acknowledgments
The authors would like to thank the Project GIAME and its team
for support during the research, Nestor Cardozo for designing algorithms freely available to the geoscientic community, and
Michael Schmitz and James Kellogg for their thoughts on previous
stages of the research and their motivation to nish the work.
Finally, we would like to thank the reviewers whose notes, comments, corrections, recommendations and thoughts greatly
improved the quality of the original manuscript, and once again to
James Kellogg Editor-in-Chief for his comments that helped us
improve the manuscript and interpretations.

M.S. Arnaiz-Rodrguez, F. Audemard / Journal of South American Earth Sciences 56 (2014) 251e264

References
n, Y., 2012. Anomalas gravime
tricas del Caribe.
Arnaiz-Rodrguez, M.S., Garzo
Interciencia 37 (3), 172e182.
Arnaiz-Rodriguez, M.S., Orihuela, N., 2013. Curie point depth in Venezuela and the
Eastern Caribbean. Tectonophysics 590, 38e51.
Arnaiz-Rodriguez, M.S., Rodriguez-Millan, I., Audemard, F., 2011. Analisis gravimetrico y exural del occidente de Venezuela. Rev. Mex. Cs. Geol. 28, 420e443.
Artyushkov, E.V., Baer, M.A., Morner, N.-A., 1996. The East-Carpathians: indications
of phase transitions, lithospheric failure and decoupled evolution of a thrust
belt and its foreland. Tectonophysics 262, 101e132.
Audemard, F.A., 2009. Key Issues on the Post-Mesozoic Southern Caribbean Plate
Boundary: Geological Society, London, Special Publications vol. 328 (1),
569e586. http://dx.doi.org/10.1144/SP328.23.
Audemard, F.A., 1993. Neotectonique, Sismotectonique et Ale a Sismique du Nordouest du Ve ne zue la (Syste'me de failles dOca-Anco n). PhD Thesis. Universite Montpellier II, France, p. 369. appendix.
Audemard, F.A., 2003. Geomorphic and geologic evidence of ongoing uplift and
rida Andes, Venezuela. Quat. Int. 101e102, 43e65.
deformation in the Me
Audemard, F.A., Machette, M.N., Cox, J.W., Dart, R.L., Haller, K.M., 2000. Map and
Database of Quaternary Faults in Venezuela and Its Offshore Regions. US
Geological Survey Open-File Report 00-0018. include map at scale 1:2,000,000,
p. 78.
Audemard, F.A., Romero, G., Rendon, H., Cano, V., 2005. Quaternary fault kinematics
and stress tensors along the southern Caribbean from faulteslip data and focal
mechanism solutions. Earth Sci. Rev. 69, 181e233. http://dx.doi.org/10.1016/
j.earscirev.2004.08.001.
rida Andes, Venezuela:
Audemard, F.E., Audemard, F.A., 2002. Structure of the Me
relations with the South AmericaeCaribbean geodynamic interaction. Tectonophysics 345, 299e327.
Audemard, F.E., 1991. Tectonics of Western Venezuela. Doctorate Thesis. Rice University, Houston-Texas.

2010. Subduction in the southern

Bezada, M.J., Levander, A., Schmandt, B.,


Caribbean: images from nite frequency P wave tomography. J. Geophys. Res. 115
(B12333), 1e19.
n, N.C., 1994. Trazas activas y sntesis neotectonica de Venezuela a escala 1:72
Beltra
000000. In: Proceedings of VII Congreso Venezolano de Geof sica, Caracas,
pp. 541e547.
Bodine, J.H., 1981. Numerical Computation of Plate Flexure in Marine Geophysics.
Lamont Doherty Geological Observatory of Columbia University, p. 153. Tech

nical Report 1.
J., Beucler, E., Trampert, J., Ritzwoller,
M.H., Capdeville, Y.,
Burgos, G., Montagner,
N.M., 2011. Proxies of Lithosphere/Asthenosphere Boundary from
Shapiro,
global surface wave tomography. In: American Geophysical Union, Fall Meeting
2011, Abstract #DI41A-2050.
Cardona, A., Valencia, V., Reiners, P., Duque, J., Montes, C., Nicoleus, S., Ojeda, G.,
Cruz, J., 2008. Cenozoic Exhumation of the Sierra Nevada de santa marta,
Colombia: implications on the interactions between the caribbean and south
american plate. In: 2008 Joint Annual Meeting. Houston, TX, GSA.
Cardozo, N., 2009. ex3dv [on-line] available at: http://www.ux.uis.no/~nestor/
Public/ex3dv.zip.
Castillo, M.V., Mann, P., 2006. Cretaceous to Holocene structural and stratigraphic
development in south Lake Maracaibo, Venezuela, inferred from well and
three-dimensional seismic data. AAPG Bull. 90, 529e564.
Ceron-Abril, J., 2008. Crustal Structure of the Colombian Caribbean Basin and
Margins. Doctorate Thesis. University of South Carolina, p. 182.
Ceron, J., Kellogg, J., Ojeda, G., 2007. Basin conguration of the northwestern South
America-Caribbean margin from recent geophysical data. CTandF, Cienc. Tecnol.
Futuro 3 (3), 25e49. Bucaramanga, Colombia.
come, M., Izarra, C., 2005. Flexural and gravity modelling of the me
rida
Chacn, L., Ja
Andes and Barinas-Apure Basin, western Venezuela. Tectonophysics 405,

155e167.

Colletta, B., Roure, F., de Toni, B., Loureiro, D., Passalacqua, H., Gou, Y., 1997. Tectonic
inheritance, crustal architecture and contrasting structural styles in the
Venezuela Andes. Tectonics 16 (5), 777e794.
De Cizancourt, H., 1933. Tectonic structure of northern Andes in Colombia and
Venezuela. AAPG Bull. 17, 211e228.
DeCelles, P.G., Giles, K.A., 1996. Foreland basin systems. Basin Res. 8, 105e123.
DeCelles, P.G., 2012. Foreland basin systems revisited: variations in response to
rez, A. (Eds.), Tectonics of Sedimentary
tectonic settings. In: Busby, C., Azor Pe
Basins: Recent Advances. Blackwell Publishing Ltd, pp. 405e426.
De Toni, B., Kellogg, J.N., 1993. Seismic evidence for blind thrusting of the northwestern ank of the Venezuelan Andes. Tectonics 12, 1,393e1,409.
Di Croce, J., 1995. Eastern Venezuela Basin: Sequence Stratigraphy and Structural
Evolution. Doctorate Thesis. Rice University, Houston, Texas, p. 225.
rida Andes and
Duerto, L., Escalona, A., Mann, P., 2006. Deep structure of the Me
 mountain fronts, Maracaibo Basin, Venezuela. Am. Assoc. Pet.
Sierra de Perija
Geolo. Bull. 90, 505e528.
Duerto, L., 1998. Principales zonas triangulares de Venezuela. M.Sc. thesis. Universidad Central de Venezuela, Caracas, p. 176. dx.doi.org/10.1029/
2006JB004802.
Egbue, O., Kellogg, J., 2010. Pleistocene to present North andean escape. Tectonophysics 489, 248e257. http://dx.doi.org/10.1016/j.tecto.2010.04.021.

263

Erlich, R., Macsotay, O., Nederbragt, A., Lorente, M., 1999. Palaecology, palaeogeography and depositional environments of Upper Cretaceous rocks of
western Venezuela. Palaeogeogr. Palaeoclimatol. Palaeoecol. 153, 203e238.
 n geofsica integrada de un transecto
Escobar, I.A., Rodrguez, I., 1995. Interpretacio
s de Los Andes venezolanos. In: Ponencia presentada en el I
NW-SE a trave
Latinoamerican Geophysical Congress, Rio de Janeiro, Brazil, pp. 273e276.
Folinsbeei, R.A., 1972. The Gravity Field and Plate Boundaries in Venezuela.
Doctorate thesis. Mass. Inst. Techn. y Wodds Hole Oceanog. Inst., USA.
French, C.D., Schenk, C.J., 2004. Map Showing Geology, Oil and Gas Fields, and
Geologic Provinces of the Caribbean Region: U.S. Geological Survey Open-File
Report 97-470-K, 1 map. [CD-ROM]. [on-line] available at: http://pubs.usgs.
gov/of/1997/ofr-97-470/OF97-470K/.
Freymueller, J.T., Kellogg, J.N., Vega, V., 1993. Plate motions in the North Andean
region. J. Geophys. Res. 98 (B12), 21,853e21,863.
Garrity, C., Hackley, P., Urbani, F., 2004. Digital Shaded-relief Map of Venezuela: U.S.
Geological Survey Open-File Report 2004e1322.
Geosoft Inc, 2007. Oasis Montaj [on-line] available at: http://www.geosoft.com/
products/oasis-montaj.
lez de Juana, C., 1952. Introduccio
n al estudio de la geologa de Venezuela. (4
Gonza
Parte). Bol. Geol. Caracas 2 (5), 311e330.
Jordan, T.E., 1995. Retroarc foreland and related basins. In: Busby, C.J., Ingersoll, R.V.
(Eds.), Tectonics of Sedimentary Basins. Blackwell Science, Oxford, pp. 331e362.
Kaniuth, K., Drewes, H., Stuber, K., Tremel, H., Hernandez, N., Hoyer, M.,
Wildermann, E., Kahle, H.G., Geiger, A., Straub, C., 1999. Position changes due to
recent crustal deformations along the Caribbean - south American plate
boundary derived from the CASA GPS Project. In: General Assembly of International Union of Geodesy and Geophysics (IUGG), Bitmingham, U. K. Poster at
Symposium G1 of International Association of Geodesy.
Karner, G., Watts, A., 1983a. Gravity anomalies and exure of the lithosphere at
mountain ranges. J. Geophys. Res. 88 (B12), 10.449e10.477.
Karner, G.D., Watts, A.B., 1983b. Gravity anomalies and exure of the lithosphere at
mountain ranges. J. Geophys. Res. 88, 10449e10477.
, VenezuelaKellogg, J.N., 1984. Cenozoic tectonic history of the Sierra de Perija
Colombia, and adjacent basins. In: Bonini, W.E., Hargraves, R.B., Shagam, y R.
(Eds.), The Caribbean Ruth America Plate Boundary and Regional Tectonics, 162.
Geological Society of America Memoir, pp. 239e261.
Kellogg, J., Vega, V., 1995. Tectonic development of Panama, Costa Rica, and the
Colombian Andes: constraints from global positioning system geodetic studies
and gravity. Spec. Pap. Geol. Soc. Am. 295, 75e90.
n of the Caribbean plate and basement
Kellogg, J.N., Bonini, W.E., 1982. Subduccio
uplifts in the overriding South American plate. Tectonics 1, 251e276.
Laske, G., Masters, G., 1997. A global Digital map of sediment thickness. EOS Trans.
AGU 78, F483.
Lugo, J., Mann, P., 1995. Jurassic e eocene tectonic evolution of maracaibo Basin,
Venezuela. In: Tankard, A., Suarez, S., Welsink, H. (Eds.), Petroleum Basins of
South America, vol. 62. AAPG Memoir, pp. 699e725.
Mann, P., Burke, K., 1984. Neotectonics of the Caribbean. Rev. Geophys. Space Phys.
22 (4), 309e362.
Mann, P., Escalona, A., Castillo, V., 2006. Regional geologic and tectonic setting of
the Maracaibo supergiant basin, western Venezuela. AAPG Bull. 90, 445e477.
Masy, J., Niu, F., Levander, A., Schmitz, M., 2011. Mantle ow beneath northwestern
Venezuela: seismic evidence for a deep origin of the Merida Andes. Earth
Planet. Sci. Lett. 305, 396e404. http://dx.doi.org/10.1016/j.epsl.2011.03.024.
n de espesor ela
stico efectivo de la litosfera en zonas de
Medina, O., 2009. Estimacio
n
cuencas cntepas: Cuenca Barinas e Apure. M.Sc thesis. Universidad Simo
Bolivar, p. 236.
Monod, B., Dhont, D., Hervouet, Y., 2010. Orogenic oat of the Venezuelan Andes.
Tectonophysics 490, 123e135. http://dx.doi.org/10.1016/j. tecto.2010.04.036.
Niu, F., Bravo, T., Pavlis, G., Vermon, F., Rendon, H., Bezada, M., Levander, A., 2007.
Receiver function study of the crustal structure of the southeastern Caribbean
plate boundary and Venezuela. J. Geophys. Res. 112, B11308.
Parnaud, F., Gou, Y., Pascual, J.C., Truskowski, I., Gallango, O., Passalacqua, H., 1995.
Petroleum geology of the central part of the Eastern Venezuela Basin. In:
Tankard, A.J., Suarez, R., Welsink, H.J. (Eds.), Petroleum Basins of South America,
vol. 62. American Association of Petroleum Geologists, AAPG, Tulsa, Oklahoma,
pp. 741e756.
rez-Gussinye
, M., Lowry, A.R., Watts, A.B., 2007. Effective elastic thickness of
Pe
South America and its implications for intracontinental deformation. G3 8.
http://dx.doi.org/10.1029/2006GC001511.
Pindell, J.L., y Kennan, L., 2009. Tectonic evolution of the Gulf of Mexico, Caribbean
and northern South America in the mantle reference frame: an update. In:
James, K., Lorente, M.A., Pindell, J. (Eds.), Origin and evolution of the Caribbean
Region, vol. 328. Geological Society of London, Special Publication, pp. 1e55.
Pennington, W., 1981. Subduction of the eastern Panama basin and seismotectonics
of northwestern South America. J. Geophys. Res. B 86, 10,753e10,770.
Pindell, J.L., Barrett, S.F., 1990. Geological evolution of the Caribbean regions; a plate
tectonic perspective. In: Dengo, G., Case, J.E. (Eds.), The Caribbean Region. Geological
Society of America. The Geology of North America, Boulder, Colorado, pp. 405e432.
Sandwell, D.T., Smith, W.H.F., 2009. Global marine gravity from retracked Geosat
and ERS-1 altimetry: ridge Segmentation versus spreading rate. J. Geophys. Res.
114, B01411. http://dx.doi.org/10.1029/2008JB006008.
Sarmiento, L.F., 2002. Relationships between stratigraphy, deformation and thermal
history in sedimentary basins; impact of geodynamic concepts in petroleum
exploration. CT&F e Cienc. Tecnol. Futuro 2 (3), 7e21.

264

M.S. Arnaiz-Rodrguez, F. Audemard / Journal of South American Earth Sciences 56 (2014) 251e264

Schmitz, M., Audemard, F.A., Orihuela, N., Klarica, S., Gil, E., Levander, A., Mazuera, F.,
Avila, J., 2013. Lithospheric scale model of Merida Andes, Venezuela (GIAME
project). In: Agu-meeting of the Americas, Cancn, Abstract T23B-08.
Smithson, S.B., Brewer, J., Hurich, C., Kaufman, S., Oliver, J., 1987. Nature of the Wind
River thrust, Wyoming, from COCORP deep - reection data and from gravity
data. Geology 6, 648e652.
Spector, A., Grant, F.S., 1970. Statistical models for interpreting aeromagnetic data.
Geophysics 35, 293e302.
Stephan, J.F., 1985. Andes et chan Caraibe sur la transversale de Barquisimeto
odinamique. In: Symposium Geodynamique des Car(Venezuela). Evolution ge
ditions Technip, Paris, pp. 505e529.
aibes. e
Stewart, J., Watts, A.B., 1997. Gravity anomalies and spatial variations of exural
rigidity at mountain ranges. J. Geophys. Res. 102, 5327e5352.
Taboada, A., Rivera, L.A., Fuenzalida, A., Cisternas, A., Philip, H., Bijwaard, H., Olaya, J.,
Rivera, C., 2000. Geodynamic of the northern Andes: subductions and intracontinental deformation (Colombia). Tectonics 19 (5), 787e813.

Turcotte, D.L., Schubert, G., 2002. Geodynamics. Cambridge University Press, New
York.
Van der Hilst, R.D., Mann, P., 1994. Tectonic implications of tomographic images of
subducted lithosphere beneath northwestern South America. Geology 22,
451e454.
Ventsel, E., Krauthammer, T., 2001. Thin Plates and Shells: Theory, Analysis and
Applications. M Dekker, New York, p. 666.
Walcott, R.I., 1970. Flexural rigidity, thickness, and viscosity of the lithosphere.
J. Geophys. Res. 75, 3941e3953.
Watts, A.B., 2001. Isostasy and Flexure of the Lithosphere. Oxford University Press,
Cambridge.
Watts, A.B., Lamb, S., Fairhead, J.D., Dewey, J.F., 1995. Lithospheric exure and
bending of the Central Andes. Earth Planet. Sci. Lett. 134, 9e21.
Watts, A.B., ten Brink, U., Buhl, P., Brocher, T., 1985. A multi-channel seismic study of
lithospheric exure across the Hawaiian-Emperor seamount chain. Nature 315
(6015), 105e111.

Das könnte Ihnen auch gefallen