Sei sulla pagina 1di 22

Bulletin of the Seismological Society of America, Vol. 103, No. 3, pp. 20252046, June 2013, doi: 10.

1785/0120120328

Tearing and Breaking Off of Subducted Slabs as the Result of Collision


of the Panama Arc-Indenter with Northwestern South America
by Carlos A. Vargas and Paul Mann

Abstract

We present two regional, lithospheric cross sections that illustrate eastward- and southeastward-dipping, subducted slabs to depths of 315 km beneath the
surface of Colombia in northwestern South America. These cross-sectional interpretations are based on relocated earthquake hypocentral solutions, models supported on
gravity and magnetic regional data, and coda-Q (Qc ) tomography. The method of
tomographic imaging based on spatial inversion of the coda wave has advantages
of providing information on the lateral variations of the anelastic properties and thermal structure of the lithospheric system. Mapping of earthquake-defined Benioff
zones combined with tomographic imaging reveals the presence of an 240 km long
eastwest-striking slab tear, named here the Caldas tear. The proposed Caldas tear
separates a zone of shallow, 2030-dipping, southeastward subduction in the area
of Colombia adjacent to Panama and the Caribbean Sea, which is not associated with
subduction-related volcanism, from an area of steeper, 3040-dipping, slab adjacent
to the eastern Pacific Ocean that is associated with an active northsouth chain of
active arc volcanoes. We propose that the Caldas slab tear separating these two distinct
subducted slabs originally formed as the southern boundary of the Panama indenter,
an extinct island arc that began subducting beneath northwestern South America about
12 Ma. The area south of the Panama indenter is Miocene oceanic crust of the Nazca
plate, which subducts eastward beneath northwestern South America at normal angles
and melts to form a northsouth-trending active volcanic arc. In addition to the formation of the Caldas tear, we propose that impedance of the thicker crustal area of the
Panama arc-indenter over the past 12 Ma may have led to down-dip break-off of
previously subducted oceanic crust that is marked by an extremely concentrated
and active earthquake swarm of intermediate-depth earthquakes beneath east-central
Colombia.

Introduction and Tectonic Setting


Hypocentral solutions recorded by the Colombian National Seismological Network (CNSN) show an 240 km
long, right-lateral offset of intermediate to deep events with
azimuth of 102 (Fig. 1a,b). We infer this discontinuity in
earthquakes to be a major slab tear which we have named
the Caldas tear based on the location in the Caldas department
of Colombia and the alignment of fault-related surface features
(e.g., volcanism, faulting, mineral deposits, geothermal anomalies, etc.). Using the distribution of earthquakes > 80 km,
Ojeda and Havskov (2001) proposed that the discontinuity
along the Caldas tear represented a boundary between two subducted slabs with differing dips and strikes: the northern subduction zone, called the Bucaramanga subduction zone, has
a shallower dip (27) and more northeasterly strike, and the
southern, called the Cauca subduction zone, has a steeper
dip (3540) and a more northerly strike (Fig. 1a).

Regional compilations of Global Positioning Systems


(GPS) data provide a quantitative tectonic framework for
understanding the widespread crustal effects of the Panama
arc collision on large areas of northwestern South America
(Calais and Mann, 2009; Fig. 1a). GPS vectors in western
Colombia show a marked decrease in velocities consistent
with the ongoing collision of the Panama arc with northwestern South America along a northsouth-trending suture
zone roughly parallel to the international boundary between
Panama and Colombia (Adamek et al., 1988; Trenkamp
et al., 2002; Corredor, 2003; Fig. 1a). The eastwest direction of GPS vectors shows that the effects of eastwest
shortening and indentation related to the collision of the Panama arc remains relatively constant over a large, V-shaped,
fault-bounded area of Colombia due east of the Panama arcindenter (Fig. 1b). GPS vectors on the Maracaibo block of

2025

2026

C. A. Vargas and P. Mann

Figure 1. (a) Tectonic map of northwestern South America and Panama showing plate boundaries, neotectonic fault systems, and selective distribution of hypocentral solutions of 30;000 earthquakes extracted from the entire catalog of the CNSN (102;000 events) during
19932012 with these criteria: mL 0:5; GAP 200; rms 0:5; error in latitude 10:0 km; error in longitude 10:0 km; and error in depth
10:0 km. Color scale indicates depth of earthquakes. The north and south profiles symbolize the tomographic sections presented in this
study. SMM, Santa Marta massif; CB, Choco block; WC, Western Cordillera; CC, Central Cordillera; EC, Eastern Cordillera; PR, Perija Range;
GB, Guajira basin; LB, Llanos foreland basin; MMVB, Middle Magdalena Valley basin; RFZ, Romeral fault zone; SMBF, Santa Marta
Bucaramanga fault; PF, Palestina fault; CF, Cimitarra fault; MGF, MulatoGetudo fault; HF, Honda fault; SFS, Salinas fault system; GF,
Garrapatas fault; LFS, Llanos fault system; IF, Ibague fault; SR, Sandra ridge; BN, Bucaramanga nest; CN, Cauca nest; MN, Murindo nest;
PIVC, PaipaIza volcanic complex; RSDV, Romeral and San Diego volcanoes. Yellow stars correspond to (1) the Tauramena earthquake (19
January 1995, Mw 6.5); (2) the Armenia earthquake (25 January 1999, M w 6.2); and (3) the Quetame earthquake (24 May 2008, mL  5:7).
Sections AA0 and BB0 correspond to tomographic profiles presented in Figures 5 and 6. (b) Crustal isochron pattern of the Sandra ridge; pinkcolored line, Caldas tear zone; arrows, station velocity GPS vectors relative to stable South America (after Calais and Mann, 2009). CHEP and
BOGO are GPS stations used as reference to estimate the onset of the Panama-arc and South American plate collision. Other GPS stations in
the Panama-arc collision area are MANZ, RION, BUCM, MONT, and CART. Faded blue arrow enclosing 102 azimuth of the approximately
240 km long, right-lateral offset of intermediate to deep events associated with the Caldas tear.

Colombia and Venezuela show a more northerly direction of


plate motions related to northward tectonic escape of the
Maracaibo block into the southern Caribbean (Trenkamp
et al., 2002). In contrast to this fairly uniform GPS velocity
field of deformed crustal rocks produced by the Panama collision, underlying, eastward-dipping slabs change abruptly
across the Caldas tear from dip angles of 3040 between
latitudes 3:05:6 N in southern Colombia, to dip angles of

2030 in the area north of 5:6 N (Ojeda and Havskov,


2001; Vargas et al., 2007).
Two nests of concentrated intermediate-depth earthquakes are present beneath Colombia (Fig. 1a). The Bucaramanga earthquake nest (BN) is found at a depth of 160 km
on the down-dip extension of the southern (Bucaramanga)
subduction zone and has an estimated volume dimension
of 13 18 12 km (Schneider et al., 1987; Frohlich et al.,

Tearing and Breaking Off of Subducted Slabs as the Result of Collision of the Panama Arc-Indenter
1995). Previous tectonic interpretations of the origin of the
Bucaramanga nest vary from a zone of two slabs in contact
(van der Hilst and Mann, 1994), two slabs overlapping
(Taboada et al., 2000), or a single slab undergoing extreme
bending (Corts and Angelier, 2005) all occurring in the
boundary area of the subducted northern (Bucaramanga)
and southern (Cauca) subduction zones (Fig. 1a). The Cauca
intermediate-depth earthquake nest (CN) is located 400 km
southwest of the Bucaramanga nest on the trend of our proposed Caldas tear and has been previously interpreted
by Corts and Angelier (2005) as a bend in the slab in this
area (Fig. 1a). There is no clear consensus among seismologists for the tectonic interpretation of the two concentrated
Colombian intermediate earthquake nests (Frohlich, 2006;
Zarifi, 2006).
The Caldas tear defines the northern limit of the active
volcanic front of the northern Andes that has formed as a
consequence of the steeper subduction of oceanic slab of
normal thickness of the Nazca plate (Fig. 1a). Moreover,
associated with active and inactive volcanoes, the eastwest
projected surface trace of the Caldas tear localizes an east
west alignment of some unusual volcanic rocks including
adakites (Borrero et al., 2009; Fig. 1a). Other volcanic rocks
in the vicinity of the eastwest-trending Caldas tear include
the Plio-Pleistocene Paipa-Iza volcanic complex in the
Eastern Cordillera of Colombia and the Romeral and San
Diego volcanoes (Pardo et al., 2005). The presence of these
eastwest aligned volcanic rocks along with locally elevated
geothermal gradient values (Vargas et al., 2009) suggests that
the Caldas tear may penetrate the upper crust as a fault zone
and provide a conduit for the upward rise of magmas and
hydrothermal fluids produced by melting of the slabs on either side of the Caldas tear (Fig. 2). Furthermore, recent,
shallow-focus, strong motion events such as the Tauramena
earthquake (19 January 1995; M w 6.5, h  25  10 km),
the Quindio earthquake (25 January 1999; M w 6.2,
h  18:6 km), and the Quetame earthquake (24 May 2008;
mL 5.7; h  superficial) are all in alignment with the surface
trace of the Caldas tear.
Previous tomographic studies using both local and
regional earthquakes of varying resolution have produced
differing tectonic interpretations for slabs in this area (van
der Hilst and Mann, 1994; Taboada et al., 2000; Vargas
et al., 2007). In this paper, we present the results of an integrated geophysics and geologic study that improves the 3D
imaging of the interactions between the eastward-moving
Panama indenter and its collisional area in northwestern
South America.

Data and Methods


The following sections describe data and procedures
used to estimate hypocentral solutions, the attenuation and its
spatial distribution, the simultaneous 2D inversion of gravity
and magnetic data, and the correlation of these results with
focal mechanisms, geothermal gradients, geological maps,

2027

and a high-resolution seismic profile, seeking to define the


geometry of the Caldas tear and its geotectonic implications
in the northwestern corner of South America.
Hypocentral Solutions and Estimation
of the Coda-Wave Attenuation
A catalog has been compiled of 102;000 earthquake
locations calculated by the CNSN during the period 1993
2012 (mL 6:8). Hypocentral solutions were estimated by
using a seismological array of 17 short-period instruments
(T  1 s) of the CNSN and complemented by 13 stations
associated with local volcanic monitoring systems and also
foreign networks (Panama, Ecuador, and Venezuela). Final
solutions were calculated with the HYPOCENTER program
and the velocity model proposed by Ojeda and Havskov
(2001). Then, 9338 waveforms associated with 7645 regional
earthquakes (3:0 mL 6:5; 19932012) were selected for
estimating the decay rate of the coda amplitudes (Q1
c , coda
attenuation). The selected events, on basis of a significant
number of stations that recorded them (Table 1), have epicentral distances to stations ranging between 22.6 and 690.0 km
and depths varying between 0 and 222.0 km. Figure 2 shows
all events used for the Q1
c plotted on a map of northwestern
South America along with tectonically significant earthquake focal mechanisms including the Quindio, Quetama,
and Tauramena events aligned along the Caldas tear and intermediate-depth focal mechanisms from the Bucaramanga
and Cauca nests.
Estimations of the Q1
c were done using the Single
Backscattering model proposed by Aki and Chouet (1975).
This model assumes that the coda of a local earthquake is
composed of the sum of secondary S waves produced by
heterogeneities distributed randomly and uniformly within
the lithosphere. The coda is the portion of a seismogram corresponding to back-scattered S-waves. The estimation of Q1
c
used the following equation:
 
P; t 

2gjSj
e
t2

t
Qc

(1)

where P; t is the time-dependent coda power spectrum,


is the angular frequency, is the shear-wave velocity, jSj is
the source spectrum, and g represents the directional scattering coefficient. The g term has been defined as 4 times
the fractional loss of energy by scattering, per unit travel distance of primary waves, and per unit solid angle of the radiation direction measured from the direction of primary wave
propagation. Using these assumptions, the geometrical spreading is assumed to be proportional to r1 , which only applies
to body waves in a uniform medium. The source factor can be
treated as a constant value for single frequency. According to
equation (1), Q1
c values can be obtained as the slope of the
least-squares fit of Lnt2 P; t versus t, for t > t , where
t represents the S-wave travel time (Haskov et al., 1989). The

2028

C. A. Vargas and P. Mann

Figure 2. Epicenter projection of events used during the coda-wave-attenuation (Q1


c ) estimation. Colored circles, earthquakes; blue
squares, locations of all seismological stations used in this paper; gray stars are shown with large focal mechanisms, and the most recent and
surficial strong-motion events occurring along the Caldas tear are shown by banded-gray polygon. The main focal mechanisms reported by
the NEIC-USGS (mb 4:0) defining the Bucaramanga nest to the northeast and the Cauca nest to the southwest are shown; pink areas
identified in the epicentral location of these nests are two main geothermal gradient anomalies reported by Vargas et al. (2009).

time-dependent coda power spectrum was calculated using the


mean squared amplitudes of the coda Aobs ; t from bandpass-filtered seismograms around a center frequency.
In order to take into account the deep structure using
coda waves,  4:64 km=s was assumed and calculated as
a weighted average of S-wave speeds in the whole earth volume covered by the scattered waves (Badi et al., 2009). All
records were filtered in a chosen frequency band and then
used a coda-wave time window (W) of 20 s, starting from
2 t ststart . The average lapse time, defined as tc  tstart 
W=2 ranges between 11.0 and 384.0 s. These large tc values
ensure the sampling of regional structures. Attenuation esti-

mates were performed with short-period records (T  1 s) at


several frequencies (Table 1). Then we chose estimates in the
frequency band 13 (2  1) Hz because of the high availability and geographical distribution of observations regarding
other frequencies; and also best values of correlation coefficients, the root mean square (rms), and signal-to-noise ratio.
In addition, it has been reported that the study region presents
Q1
c values in this frequency band with errors 5% (e.g., see
Vargas et al. (2004)). In general, errors seen along Q1
c
estimations are acceptable, for example, the rms of all estimations vary between 0.07 and 1.79 (  0:24,  0:07)
and the coefficients of correlation are oscillating between

Tearing and Breaking Off of Subducted Slabs as the Result of Collision of the Panama Arc-Indenter

2029

Table 1
Estimated Values of Coda-Wave Attenuation (Q1
c ) at Various Frequencies
3
Q1
c 10

Coefficient of Correlation

rms

Signal/Noise

Frequency

Waveforms
Analyzed

Min

MeanStd

Max

Min

MeanStd

Max

Min

MeanStd

Max

Min

MeanStd

Max

2
8
12
16

9338
5421
4441
3741

1.7
0.8
0.6
0.5

7.12.7
1.80.7
1.20.4
0.90.3

47.6
20.4
1.2
9.4

0.97
0.96
0.94
0.95

0.670.11
0.600.08
0.590.08
0.580.07

0.5
0.5
0.5
0.5

0.07
0.18
0.15
0.17

0.240.07
0.320.06
0.350.06
0.350.09

1.79
2.67
1.89
4.13

2.0
2.0
2.0
2.0

16.247.5
11.025.6
9.721.1
9.018.1

985.5
842.6
600.9
315.7

Also presented are extreme values, averages, standard deviations, and quality parameters. Q1
c values at 2 Hz were used to estimate tomograms because of
the high availability of observations regarding other frequencies, best values of correlation coefficients, rms, and signal-to-noise ratio. A power law equation
1
3 0:970:06 .
for all Q1
c observations suggested a high-frequency dependence of the attenuation in this region: Qc f  13:2  0:6 10 f

0:5 and 0:97 (  0:67,  0:11). Table 2 presents a


statistical summary of the main parameters related with the
values for 30 seismological stations.
estimation of Q1
c
Figure 3a shows an example of typical waveform used during this analysis, as well the corresponding record filtered for
the chosen frequency band. The attenuation factor (Q1
c ) is
suggested as a decay factor for the coda-wave amplitudes.
Figure 3b presents histograms for Q1
c values and their correlation coefficients, as well as distributions for the epicentral
distances, focal depths, and local magnitudes of the events
analyzed. Figures 2 and 3b emphasize the presence of attenuation contrasts in the region and at least two sources of
events, one of them surficial and dispersed, and the other
located at an intermediate depth (linked to the nests of Bucaramanga and Cauca).
Tomographic Imaging Using Coda-Wave Attenuation
Mukhopadhyay and Sharma (2010) have proposed that
the variation of Q1
c with tc shows a direct relationship with
depth. These authors interpreted that Q1
c values related to
scattering processes that penetrate > 200 km depth are controlled by a crust and a relatively more transparent mantle.
These results support the idea that Q1
c estimated with a large
tc is representative of a large sampled volume and large
sampled depths. A corollary of this hypothesis is that the
1
Q1
c value must be near to the intrinsic absorption (Qi ) controlled mainly by the mantle. Following these ideas, Vargas
et al. (2004) developed a regional tomographic study using
stations of the CNSN with relative large tc , (up to 180 s) and
1 values for
found that the Q1
c values are near to the Qi
almost all stations, meaning that a large portion of the upper
mantle is being sampled. Other studies have suggested a
direct relation between the thermal field and anelastic attenuation (Faul and Jackson, 2005; Priestley and McKenzie,
2006; Yang et al., 2007). The physical meaning of this relationship is not been completely understood, but Karato and
Jung (1998) proposed that the higher water content in the
asthenosphere significantly reduces the seismic-wave velocities through anelastic relaxation and increasing temperature.
Convergent margins such as Colombia, which involve large
amounts of sediments and water mobilized during the sub-

duction processes, are a likely site of large contrasts in


anelastic attenuation in the subducted lithospheric slabs.
Given the ease to estimate Q1
c , we can use this observation for highlighting regional structures related to contrasts
in rigidity (e.g., crust or lithospheric plates). One way to
regionalize Q1
c is based on the work of Malin (1978) who,
expanding on the work of Aki (1969) and Aki and Chouet
(1975), realized that the first-order scatterers responsible for
the generation of coda waves at any given tc can be located
on the surface of an ellipsoid with earthquake and station
locations as foci (Singh and Herrmann, 1983). In the ellipsoidal volume sampled by coda waves at any time t, Pulli
(1984) defined the large semi-axis as a1  t=2, and defined
the small semi-axis as a2  a3  a21 r2 =41=2 , where r is
the sourcereceiver distance of the ellipsoid. The horizontal
projection of this volume is coincident with the elliptical envelope proposed by several authors as the area occupied by
the scattered energy of the coda-wave record (Mitchell et al.,
1997; Mitchell and Cong, 1998; Xie, 2002; Vargas et al.,
2004). Following these observations and knowing the values
of tc , W, and , it is possible to deduce the volumes of the
ellipsoidal shells where the seismic energy is scattered.
Hence Q1
c values estimated with large tc correspond to large
sampled volumes, and vice versa. Based on these hypotheses
we can perform a generalized inversion for regionalizing
Q1
c . For the purpose of the inversion, we define a geographic grid around the seismic station that also encloses the
hypocenter. We recognize that each measured Q1
c is an aver1
age estimate Q1
(or
Q
)
for
the
volume
as
sampled by
av
apparent
the ellipsoidal shell given by
V TOTAL X V Block-j

;
Qav
Qj
j

(2)

where V Block-j is the fraction of volume (block) sampled by


the ellipsoidal shell with the true attenuation coefficient Q1
j
(or Q1
true ). Assuming a constant S-wave velocity of propagation, the volume traveled by a ray that leaves the hypocenter
moves outward to the ellipsoidal shell as defined by the
observation time of the coda and is scattered to the receiver,
can be determined. Equation (2) can be written as

2030

C. A. Vargas and P. Mann

(a)

(b)

=-0.67

1600

=0.11

3000

1200

3500

2000
1500

Frequency

1400

4000

2500

1000
800
600

2500
2000
1500

400

1000

500

200

500

10

20

30

40

0
-1

50

-0.9

-0.8

-0.7

-0.6

0
0 100 200 300 400 500 600 700 800 900

-0.5

Coefficient of correlation
=87.7

=65.6

3500

3000

3000

2500

2500
2000
1500

1500
1000

500

500
0

50

100
150
Depth (km)

200

250

=0.7

2000

1000
0

Epicentral distance

=3.1

3500

Frequency

Frequency

4000

=96.0

3000

1000

=155.0

4500

3500
Frequency

Frequency

4000

Figure 3.

(a) Example of waveform used for estimating the Qc values. Upper trace represents the original record of an earthquake
recorded by a short-period seismological station of the CNSN. Middle trace represents the filtered record in frequency band 13
(2  1) Hz. Lower trace represents the decay envelope of the coda wave in a window of 20 s, starting from 2 t s. Qc value was obtained
as the slope of the least-squares fit of Lnt2 P; t versus t (dashed line with arrowheads), for t > t , where t represents the S-wave
travel time (Haskov et al., 1989). (b) Histograms for Q1
c values and their correlation coefficients, as well as distributions for the epicentral
distances, focal depths, and local magnitudes of all events analyzed.

1
1 V Block-1
1 V Block-1

  
 
Qav Q1 V TOTAL
Q1 V TOTAL
1 V Block-n
;

Qn V TOTAL

where


(3)

where the ratio V Block-j =V TOTAL is the volume fraction


associated with the total scattered-wave travel path spent
in the jth block. If the process is repeated for each station
hypocentral pair, the entire region is sampled. Equation (3) is
of the form
a1 x1      ai xi      an xn  y;

(4)

y



 

1
1
V Block-i
xi 
ai 
:
V TOTAL
Qav
Qi

Then, a least-squares estimation of the xi is given by the


compact matrix equation AX  Y where A is a (k n)
coefficient matrix, X is a (n 1) vector, Y is a (k 1) vector,
and k is the number of stationhypocenter pairs. A linear
inversion of the matrix equation was formulated as an iteratively damped least-squares technique (Levenberg, 1944;
Marquardt, 1963). The damping factor (), which adds
to the diagonal parameters of the matrix, was computed

4.49
6.58
9.17
2.68
7.72
4.63
0.94
9.94
1.22
1.57
5.01
5.20
0.94
1.21
1.23
2.54
6.19
4.01
2.84
5.57
8.24
0.24
2.33
3.71
4.86
4.9
5.89
8.88
6.23
4.59

75.40
73.18
79.84
75.44
72.79
73.73
77.89
73.44
77.25
76.95
74.2
74.04
77.83
77.42
77.35
72.63
75.53
77.34
75.95
74.87
73.32
78.45
76.4
74.89
74.33
75.35
73.08
70.63
77.41
75.32

Station

ANIL
BAR
BCIP
BET
BRI
CHI
CLIM
COD
CPAS
CRU
CTAB
CTAU
CUM
GCAL
GCUF
GUA
HEL
MAL
MARA
NOR
OCA
OTAV
PCON
PRA
ROSC
RREF
RUS
SDV
SOL
TOL

2300
1864
61
540
1427
3140
4232
108
2620
2761
3500
3868
3420
2353
3800
217
2815
75
2207
536
1264
3492
4294
468
3020
4743
3697
1620
38
2577

Altitude
(masl)

248
754
5
66
46
724
17
104
8
330
2
5
234
8
56
11
216
391
78
596
3304
26
120
313
149
205
150
64
850
258

Waveforms
Analyzed

2.9
1.7
5.5
2.9
2.9
2.7
2.5
3.4
5.1
2.5
5.6
4.3
3.0
7.9
3.7
3.6
3.0
2.4
3.0
2.5
2.0
4.3
2.0
2.5
2.9
2.5
2.6
2.8
3.1
2.5

Min

6.81.6
6.12.5
6.92.0
6.63.0
6.03.5
6.83.0
7.05.0
7.33.1
8.84.6
6.72.9
8.18.2
8.06.7
7.03.2
10.31.8
7.42.7
6.22.9
6.12.1
6.20.0
5.92.4
5.92.1
6.02.3
7.62.1
6.63.2
6.12.7
5.92.4
6.12.3
6.02.7
5.82.0
7.33.1
6.32.6

MeanStd

3
Q1
c 10

Max

17.2
32.3
12.2
14.9
47.6
20.0
12.0
18.2
15.9
16.1
14.5
28.6
22.7
13.5
18.2
10.2
19.6
15.9
18.2
20.4
21.7
13.9
16.9
19.6
33.3
12.7
20.4
12.5
25.6
20.0

12.1
11.8
104.7
16.7
13.1
11
31.7
19.3
16.6
17.6
132.4
52.8
12.9
14.6
22.0
20.3
21.0
14.9
17.9
16.3
16.4
23.6
12.5
17.6
17.9
113.7
17.5
114.5
17.2
19.8

Min

89.3 51.5
71.219.2
126.415.7
68.156.3
81.636.7
89.453.3
78.935.5
47.348.7
14.110.2
79.260.2
1330.85
96.231.2
82.757.9
22.821.12
52.340.1
174.485.5
110.240.9
65.137.5
87.665.2
113.435.5
102.823.1
70.225.7
72.452.8
101.161.3
88.247.6
118.148.8
72.639.9
176.323.7
64.639.8
100.456.1

MeanStd

tc
Max

231
227
146
246
150
263
132
180
38.4
339
134
119.1
384
69.4
152
227
190
323
259
297
287
147
289
259
204
220
240
262
305
248

0.94
0.94
0.86
0.91
0.97
0.97
0.91
0.95
0.91
0.95
0.86
0.9
0.97
0.91
0.93
0.86
0.96
0.95
0.92
0.94
0.96
0.88
0.96
0.95
0.95
0.92
0.94
0.90
0.97
0.95

Min

0.670.11
0.670.11
0.690.10
0.680.11
0.660.13
0.690.11
0.690.11
0.700.12
0.720.12
0.690.11
0.700.22
0.730.12
0.690.12
0.830.05
0.700.11
0.670.11
0.650.10
0.670.11
0.650.11
0.660.11
0.660.11
0.730.12
0.670.11
0.670.11
0.660.12
0.650.10
0.660.11
0.650.11
0.710.11
0.670.11

MeanStd

0.50
0.50
0.59
0.50
0.50
0.50
0.53
0.50
0.56
0.50
0.55
0.6
0.50
0.74
0.50
0.51
0.50
0.50
0.51
0.50
0.50
0.53
0.50
0.50
0.50
0.50
0.50
0.50
0.50
0.50

Max

Coefficients of Correlation

26.7
26.6
200.7
46.7
24.7
24.5
73.0
33.8
29.1
30.8
249.2
109.9
22.6
25.6
38.5
53.0
36.8
43.6
31.3
28.5
28.7
58.8
39.4
30.8
31.3
41.5
30.6
217.9
30.1
34.7

Min

166.390.2
142.033.7
238.727.5
136.898.5
159.665.8
173.593.8
155.562.1
100.285.3
42.217.8
156.0105.3
250.21.48
185.954.5
162.3101.3
57.437.0
108.170.5
322.7149.7
210.071.5
131.465.7
170.8114.1
215.862.1
197.440.4
140.344.9
144.192.4
194.5107.4
171.883.3
224.185.4
144.569.7
326.041.4
130.669.6
193.198.2

MeanStd

Epicentral Distances
Max

413.6
414.6
272.8
448.0
280.2
477.6
247.6
332.2
84.7
610.4
251.3
225.9
690.0
139.0
283.5
415.3
350.2
582.4
469.9
537.4
520.3
274.1
522.6
470.1
373.6
402.7
437.9
476.5
551.1
451.9

The stations detected 7645 earthquakes (19932012) that were used for estimating 9338 Q1
c values in frequency band 13 (2  1) Hz and coda-wave time window (W) of 20 s. tc values and relatively
large epicentral distances allowed us to estimate the coda-wave tomography regionally.

Latitude
()

Longitude
()

Table 2
Seismological Stations of the CNSN Used in this Study

Tearing and Breaking Off of Subducted Slabs as the Result of Collision of the Panama Arc-Indenter
2031

2032

C. A. Vargas and P. Mann

automatically for each iteration (Hoerl and Kennard, 1970;


Hoerl et al., 1975). According to this technique, the solution
and resolution matrixes can be found for the following
equations:
X  AT A  2 I1 AT Y;

(5)

R  AT A  2 I1 AT A:

(6)

and

Similar procedures for the Q1


c imaging have been used
in previous works in order to establish a deterministic characterization of the heterogeneity in the lithosphere as an
alternative technique for traditional tomographic measurements (ODoherty et al., 1997; Lacruz et al., 2009).
Resolution and Reliability
A spatial inversion of attenuation of 32 32 8
blocks with block dimensions of 60 km latitude
50 km longitude 40 km thickness was designed in
order to detect relevant structures in the region. We qualified
the tomographic inversion by means of three approaches:
(a) hit count of ellipsoidal shells; (b) solving controlled tests;
and (c) mapping the diagonal elements of the resolution matrix (RDE) by using equation (6). The hit count is a very
rough quality estimation that only tells about summing up
the number of ellipsoids that contribute to the solution of
a block. Based on this discretized volume, we mapped the
hit count with the available data in eight layers (0, 45, 90,
135, 180, 225, 270, and 315 km; Fig. 4a). Although a large
part of northwestern South America (including Colombia,
western Venezuela, eastern Panama, and northern Ecuador)
is covered by ellipsoidal shells (over 500 crossings), it is in
northern Colombia and northwestern Venezuela (71 W to
76 W; 5 N to 10 N; 0180 km depth) where the largest
number of shells run through the blocks (based on more than
5000 hits per block). This approach emphasizes the importance of the Bucaramanga nest data in the estimation of tomographic images.
To incorporate the second approach, we evaluated the
efficiency of the method described above solving the 3D direct problem. The ellipsoidal shell volume associated with all
foci (pairs of earthquake and station) were used to relate the
Q1
c values in two controlled test boards. The first one was
for appraising large domains of attenuation, for example, a
zone with flat subduction in the north, and the other zone
with normal subduction in the south (Fig. 4b). The second
test board evaluated the ability of the method to detect small
anomalies by use of a typical chessboard (Fig. 4c). In tests,
we assigned two values of Q1
true that represent attenuation
contrasts (1=70 and 1=200) into the 3D grid. Then we esti1
mated the Q1
av values (or theoretical Qc values) for all ellipsoids (each one related to foci [earthquakestation]) by
estimating the weighted average of Q1
true involved in the volume of each ellipsoid.

For the 3D inverse problem, we estimated the fraction


of volumes associated with each Q1
av in order to establish
equation (3). Using all foci related to the events selected in
this study, we assembled the compact matrix of equation (4)
and then we inverted the Q1
av values using equation (5).
Finally, a spatial interpolation of the Q1
av values was done
based on the Kriging method (Oliver and Webster, 1990)
and presented on Mercator projection. Figure 4d shows
the results of the inversion for the synthetic experiment based
on two domains of contrasting attenuation (Fig. 4b). This
experiment is comparable to a slab-tearing model for which
two zones with different angles of subduction, are related to
different attenuations. This hypothetical model linked a flat
subduction zone in the north (lower attenuation) and a normal subduction zone in the south (higher attenuation). In
general, the available data may allow detection of large structures with significant contrasts of attenuation as much as
270 km depth. On the other hand, Figure 4e presents
the inversion for a chessboard based on two areas of contrasting attenuation (Fig. 4c). This experiment suggests that
the available data may allow detection of smaller bodies
(e.g., 100 km 100 km 60 km) with significant contrasts
of attenuation, mainly in Colombia, and as much as 180 km
depth.
After several trials of accurate resolution and stability, the spatial inversion of attenuation with real data
was performed with the same grid (32 32 8 blocks).
Figure 4f,g shows results of the tomographic estimation
and their maps of the RDE at different depths. Because each
RDE shows the amount of independence in the solution of
one model parameter (RDE oscillates between 0 and 1), the
larger value of the RDE for one model parameter suggests a
more independent solution for this parameter. 3D inversion
presents higher RDE values (e.g., > 0:4) limited by the availability and geographical concentration of Q1
c values, indicating that the method is useful for areas for which a large
stacking of attenuation observations is present. From the
available earthquake data, the tomographic solution of the
attenuation efficiently images large areas of the crust and
upper mantle of northwestern South America including
Colombia, eastern Panama, and western Venezuela with
sampling depths reaching > 315 km (Fig. 4f). Because the
ellipsoids related to deeper hypocentral solutions can sample
profounder volumes, the 3D inversion may detect the thermal
influx from the mantle adequately.
In order to infer the geometry of the Caldas tear and its
relationships with the adjacent Nazca and Caribbean plates,
we made two regional cross sections: (1) a northern section (AA0 , Fig. 1a) from the Caribbean plate to the Llanos
foreland basin of eastern Colombia and crossing the intermediate-depth earthquakes of the Bucaramanga nest; (2) a
southern section designed for imaging the corridor between
the Nazca plate and the Llanos basin and crossing the intermediate-depth earthquakes of the Cauca nest (BB0 , Fig. 1a).
As discussed subsequently, it is essential to incorporate all

Tearing and Breaking Off of Subducted Slabs as the Result of Collision of the Panama Arc-Indenter
available geophysical data for proper interpretation of the
tomograms along these sections.
Integrating Earthquake Data with Regional
Seismic-Reflection Lines
Hypocentral solutions of the CNSN (rms < 0:3 s;
GAP < 200; stations 6; error in latitude 10:0 km; error in

2033

longitude 10:0 km; error in depth 5:0 km) were plotted on


the tomographic profiles along two 60 km wide corridors
(Figs. 5 and 6). Because seismicity in a corridor parallel to
the northern section is sparse, we have included an interpretation of the Trans-Andean megaregional seismic-reflection line
that extends from the Caribbean coast to the Eastern Cordillera of Colombia (Vargas et al., 2010) and to the northern tomographic section. This 383 km long reflection line is a 20 s

Figure 4. Resolution, reliability, and results of the spatial inversion of attenuation based on a geometry of 32 32 8 blocks with
dimensions of 60 km latitude 50 km longitude 40 km thickness. Coda-wave tomograms were estimated with 9338 Q1
c observations associated with 7645 regional earthquakes (mL 6:7; 19932012) in the frequency band 13 (2  1) Hz. (a) Hit count of ellipsoidal shells, suggesting that the 3D inversion of Q1
c may solve large areas of Colombia, eastern Panama, western Venezuela, and northern
Ecuador. (b) Synthetic model that represents two large domains of attenuation (e.g., a zone with flat subduction in the north and normal
subduction in the south, limited by a slab tear). The contrasts of attenuation incorporated into the model to evaluate the effectiveness of the
1
method were Q1
c  1=200 and Qc  1=70. (c) Chessboard with smaller and regular distribution of attenuation contrasts. As in the previous
1  1=70. (d) 3D inversion of the synthetic model
case, the contrasts of attenuation incorporated into the model were Q1
c  1=200 and Qc
presented in (b) suggesting that the distribution of the available data may allow detection of large structures with significant contrasts of
attenuation as much as 270 km depth. (e) 3D inversion of the chessboard model presented in (c) suggesting that the distribution of the
available data may permit detection of smaller bodies (e.g., 100 km 100 km 60 km) with significant contrasts of attenuation, mainly in
Colombia, and as much as 180 km depth. (f) Results of the tomographic inversion with the available data. (g) Maps of the RDE at different
depths. Higher RDE values (e.g., 0:5) indicate zones efficiently solved. However, these higher RDE values were limited by the geographical
concentration of Q1
c values, indicating that the method is useful for areas where a large stacking of attenuation observations is present.
Tomographic solution of the attenuation efficiently images large areas of the crust and upper mantle of northwestern South America including
Colombia, eastern Panama, and western Venezuela with sampling depths reaching > 315 km. High-attenuation anomalies suggest that Bucaramanga and Cauca seismic nests may be related to asthenospheric emplacements.
(Continued)

2034

C. A. Vargas and P. Mann

Figure 4.

Continued.

Tearing and Breaking Off of Subducted Slabs as the Result of Collision of the Panama Arc-Indenter

Figure 4.

Continued.

2035

2036

C. A. Vargas and P. Mann

Figure 5. Section crossing the northern Panama-arc indenter and its down-dip Bucaramanga nest (Fig. 1a, AA0 ). (a) Geologic and geothermal observations; (b) gravity and magnetic data; (c) interpreted tomographic section. Green dots with vertical bars that represent vertical
errors, hypocentral solutions in a 60 km wide corridor. Plotted events have the following selection criteria: rms < 0:3 s; GAP 200; stations
6; error in latitude 10:0 km; error in longitude 10:0 km; error in depth 5:0 km. Some representative focal mechanisms (beach balls)
have been also plotted.

record and 200-fold, and shows the subduction geometry of


northern (Bucaramanga) slab dipping at a shallow angle
to the southeast beneath northwestern Colombia (Fig. 7).
Although reflectors are difficult to distinguish in deeper areas
of the line, the overall distribution of deep reflectors dips eastward in same amount as the subducted slab on the gravity and
magnetic model in Figure 5c. Deep reflections are concentrated within what we interpret as the lower crust whereas
the upper mantle appears more transparent (Tittgemeyer et al.,
1999). Prominent reflections in the upper crust can be correlated with major sedimentary basins such as the SinuSan Jacinto, Lower Magdalena, Middle Magdalena, and Eastern
Cordillera, as well as major faults such as the Romeral fault
zone (RFZ). This major fault separates oceanic crustal rocks in
the western terranes of Colombia and continental basement in
eastern Colombia (Cediel et al., 2003). In general, seismicity
east of the Romeral is more concentrated in the older and

more anisotropic continental crust of northwestern South


America.
Gravity and Magnetic Modeling
Coincident gravity and magnetic models were completed
for this study based on regional information (Maus et al.,
2007; Sandwell and Smith, 2009; National Hydrocarbons
Agency of Colombia, 2010; Figs. 5b and 6b). The gravity
and magnetic data was merged with the 90 m elevation topographic information available from NASA (Jarvis et al.,
2008), with corrections from the International Gravity Standardization Net 1971 (IGSN71), the World Geodetic System
1984 (WGS-84), the International Geomagnetic Reference
Field (IGRF) and the Observed Magnetic Intensity (Hinze
et al., 2005; Maus et al., 2005). The final database allowed
us to estimate free air and magnetic anomalies. We then calculated Bouguer anomalies using densities of 2:67 g=cm3 for

Tearing and Breaking Off of Subducted Slabs as the Result of Collision of the Panama Arc-Indenter

2037

Figure 6. Coda-wave-attenuation section crossing the southern part of the Panama indenter and the Cauca nest (Fig. 1b, BB0 ). (a) Geologic and geothermal observations; (b) gravity and magnetic data; (c) interpreted section. Green dots with vertical bars that represent vertical
errors, hypocentral solutions in a 60 km wide corridor. Plotted events have the following selection criteria: rms < 0:3 s; GAP 200; stations
6; error in latitude 10:0 km; error in longitude 10:0 km; error in depth 5:0 km.

land and 1:03 g=cm3 for marine water. In order to estimate


gravity and magnetic model responses (Talwani et al.,
1959; Talwani and Heirtzler, 1964; Geosoft, 2010) and comparing with the observed data, we used values of density, magnetization, and magnetic susceptibility shown on Table 3. In
addition, the gravity and magnetic models were constrained
with the seismological and seismic data, as well as geologic
transects compiled with superficial cartography and representative seismic lines (Section AA0 , Figs. 1a and 5; and Section
BB0 , Figs. 1a and 6; Lopez, 2004).
Because of restrictions on the data use of the TransAndean megaregional seismic-reflection line, it was not
possible to estimate refraction travel-time tomography for
correlating with the profiles presented in this paper. However,
in order to interpret the thermal and tectonic structure in this
region, we used the velocity anomalies reported by Vargas
et al. (2007) and van der Hilst and Mann (1994) that show
similar distribution anomalies. In general, trends of high

velocities match with slabs suggested by the gravity and


magnetic models.
Other Geophysical Information for Constraining
the Interpretation
Seismic attenuation has been used for examining the
acoustic contrast between the upper mantle and the lithosphere because it is believed that this seismic factor is a
physical parameter closely related to the thermal state of
the volume sampled by the waves (Faul and Jackson, 2005;
Priestley and McKenzie, 2006; Yang et al., 2007). Therefore,
in order to evaluate the empirical superficial response of the
lithospheric thermal field, we plotted geothermal anomalies
reported from oil wells in Colombia (Vargas et al., 2009)
onto the two sections (Figs. 5a and 6a). Furthermore, topographic response and focal mechanisms compiled from NEIC
are plotted on the two profiles.

2038

C. A. Vargas and P. Mann

Figure 7. (a) Deep seismic profile nearly parallels the tomographic section AA0 (see inset map). Seismic image was assembled with three
segments of the Trans-Andean Seismic Line acquired by the National Hydrocarbons Agency of Colombia (ANH; Vargas et al., 2010). (b) An
interpretation of the seismic line. Black lines, suggested sedimentary basins and deeper reflections; blue lines, faults and main tectonic
features (e.g., the Romeral fault zone in bold line); yellow line, suggested detachment surface associated with the Caribbean plate subduction.
Red dots with vertical bars that represent vertical errors, hypocenter solutions in a 60 km wide corridor. Plotted events have the following
selection criteria: rms < 0:3 s; GAP 200; stations 6; error in latitude 10:0 km; error in longitude 10:0 km; error in depth 5:0 km.
Depthtime relation has been estimated by using several oil wells in the area (Vargas et al., 2010).
To support our interpretation, we have presented a total
of eight variables along the profiles: hypocenter solutions,
focal mechanisms, coda-wave attenuation, gravity, magnetic
and geothermal anomalies, and geologic and topographic
data derived from seismic-reflection profiles. The two sec-

tions flank the intersection of the Panama arc-indenter and


the Caldas tear to the north (Section AA0 , Figs. 1a and 5)
and south (Section BB0 , Figs. 1a and 6). The profiles are extended to the west from the Caribbean Sea and Pacific Ocean
to the Llanos foreland basin in the east. Both images cross

Table 3
Physical Properties Expressed in SI for the Materials Used in Gravity and Magnetic Models
Unit

Caribbean and Nazca plates


Lower continental crust
Upper continental crust
Mantle
Accreted sediments and oceanic crust
Oceanic sediments
Water

Density (kg=m3 )
3

3:30 10
2:78 103
2:31 103
3:15 103
2:07 103 2:35 103
1:68 103 2:00 103
1:03 103

Susceptibility
1

Magnetization (A=m)

1:26 10 1:38 10
1:26 105
3:21 102 4:90 102

1:00 103 1:45 100


1:26 105 1:90 100
5:51 101 1:00 100

1:26 101
1:26 105

1:00 103 8:01 101


1:00 103 4:50 101

Tearing and Breaking Off of Subducted Slabs as the Result of Collision of the Panama Arc-Indenter

Longitude ()

(a)

2039

-72.8
-73
-73.2

y(x) = a x + b
a = -0.00529
b = 17.428
R = 0.36776

-73.4

(b)

7.4

Latitude ()

1994

7.2
7

1996

1998

2000

2002

2004

2006

2008

2010

2012

2014

1998

2000

2002

2004

2006

2008

2010

2012

2014

1998

2000

2002

2004

2006

2008

2010

2012

2014

y(x) = a x + b
a = -0.006479
b = -60.081
R = 0.32692

6.8
6.6

Depth (km)

(c)

1994

1996

-80
-100
-120
-140
-160

y(x) = a x + b
a = 0.23732
b = -621.43
R = 0.11894

-180
-200
-220

1994

1996

Time (year)

Figure 8.

Temporal evolution of the hypocentral parameters of the Bucaramanga nest. Earthquake information provided by the CNSN
from 1993 to 2012. Events of the Bucaramanga nest were selected around the point 73.1 W, 6.9 N. Parameters of selection were
radii  0:2; 0 rms 0:2 s; 120 h 160 km; GAP 200; (a) error in longitude 10:0 km; (b) error in latitude 10:0 km; (c) error
in depth 10:0 km. Error bars have been associated with each event. Dashed polygon, the linear trend of occurrences estimated by leastsquares means.

the northern Andes, the Bucaramanga and Cauca seismic


nests of intermediate depth earthquakes, and major faults
including the Romeral, the Llanos, the Santa Marta
Bucaramanga, the Garrapatas, and the Ibague faults.

Results and Discussion


Northern Transect Crossing the Caribbean Plate
and the Bucaramanga Nest
Our northern transect shows the presence of an oceanic
crust with thicknesses greater than 20 km related to the Caribbean oceanic plateau with a shallow subduction angle of
< 15 over a distance of 100 km that extends from the
South Caribbean deformed belt to the Romeral fault zone
(Figs. 1a and 5). At this point the slab dip steepens to 20
30 into the mantle. Following our interpretation of the
Trans-Andean seismic-reflection line (Fig. 7), the interaction
between the continental crust and the South Caribbean Deformed Belt (SinuSan Jacinto basin) suggests that the oceanic crust is sinking with a higher dip and is reaching the
intermediate to deep seismicity zone (160 km), where

the sinking oceanic crust further increases its dip to penetrate


the mantle.
Intensive intermediate seismicity of the Bucaramanga
nest is concentrated within the subducted Caribbean slab
at depths of 130160 km (Figs. 1a and 5). Furthermore, there
are at least two clusters of events concentrated at 80 and
110 km. Focal mechanisms of the Bucaramanga nest for
events reported by NEIC (19772012) with mb 4:0 include
mainly reverse and strike-slip events (Fig. 2). We relate this
earthquake nest to resistance to subduction of the Panama
indenter and consequent initiation of tearing and slab break
off at intermediate depths of the down-dip part of the slab
(Corts and Angelier, 2005). In order to show the temporal
trend of the nest using seismicity located by the CNSN between 1993 and 2012, we have compiled data from the Bucaramanga nest (Zarifi, 2006; Fig. 8). Although errors related
to hypocentral solutions and temporal variation in the array
configuration may blur the dimensions of the nest (Schneider
et al., 1987; Frohlich et al., 1995; Zarifi, 2006), and also
nearby events could be associated erroneously, the continuous recording of seismicity for a period as long as 19 years
shows a progressive southward and westward displacement

2040

C. A. Vargas and P. Mann

Longitude()

(a) -75.5
-76

-76.5

-77

Latitude()

(b)

y(x) = a (x - b)
a = 0.021514
b = 5553.5
R = 0.40882

1994

1996

1998

2000

2002

2004

2006

2008

2010

2012

2014

1996

1998

2000

2002

2004

2006

2008

2010

2012

2014

1996

1998

2000

2002

2004

2006

2008

2010

2012

2014

5.5
5
4.5

y(x) = a x + b
a = 0.0085141
b = -12.525
R = 0.12753

(c)

Depth(km)

3.5

-50
-100

1994

y(x) = a x + b
a = 1.0488
b = -2211.1
R = 0.19102

-150
-200

1994

Time (year)

Figure 9. Temporal evolution of the hypocentral parameters of the Cauca nest. Earthquake information provided by the CNSN from 1993
to 2012. The Cauca nest events were selected around the point 76.3 W, 4.5 N. Parameters of selection were Radii  0:5; 0 rms 0:3 s;
h 60 km; GAP 200; (a) error in longitude 10:0 km; (b) error in latitude 10:0 km; (c) error in depth 10:0 km. Error bars have been
associated with each event. Dashed polygon, the linear trend of trend of occurrences estimated by least-squares means.
of events. Although lineal regression of the temporal series of
events is of low confidence, this unidirectional, westward
displacement of events may be caused by down-dip and
southwestward propagation of tearing of the subducted
Caribbean slab.
Southern Tomographic Transect Crossing the Nazca
Plate and the Cauca Nest
The Nazca oceanic slab has been modeled with an
1522 km deep crustal thickness and a constant dip angle
of 3040 to a depth > 150 km beneath the active volcanic
line (Figs. 1a and 6). The volcanic line is underlain by
high-attenuation anomalies indicative of a normal melting
range for the subducted oceanic slab (Figs. 2, 4c, 6). Highattenuation anomalies and the presence of shallow to intermediate seismicity around the Romeral fault zone suggest
this major strike-slip provides another major upward conduit
for the release of upper mantle heat. A large low-attenuation
anomaly corresponds to the low geothermal gradient observed between the Colombian trench along the Pacific margin and the Central Cordillera. The low geothermal gradient
coincides with thick volcanic and sedimentary material accreted to western Colombia, mainly in the Western Cordillera
and the Baudo Range. The accretion of this area may have

accompanied a proposed westward jump in subduction from


the Romeral fault zone in the Central Cordillera to the
present Colombia trench (Cediel et al., 2003). An additional
attenuation anomaly observed on tomographic data indicates
a prominent high thermal inflow from the mantle that is
located beneath the forebulge of the Llanos basin and is approximately coincident with the largest geothermal anomaly
measured in this basin.
Seismicity of the Cauca nest is highly dispersed in depth
(70150 km) and its time evolution seems more complex
than the Bucaramanga nest (Fig. 9). Even with the low confidence of the lineal regression, seismicity from the CNSC
catalog shows an eastward displacement of events. The Cauca
nest exhibits earthquakes with focal mechanisms ranging from
pure gravitational collapse to strike-slip in the north and
reverse with strike-slip component in the south (Fig. 2).
Nature of the Caldas Tear
In addition to the hypocenter solutions that initially
revealed the 240 km long offset of intermediate to deep
seismicity of the Caldas tear, we found that differential
displacements between the southern of the Caldas tear in the
South American plate and the Panama arc-indenter, based on
GPS observations, suggest a trend of decreasing displacement

Tearing and Breaking Off of Subducted Slabs as the Result of Collision of the Panama Arc-Indenter
toward the east (Trenkamp et al., 2002; e.g., BOGO versus
MZAL, RION, BUCM, MONT, and CART GPS stations). Assuming the BOGO station as the reference point south of the
Caldas tear, and the CHEP station as the reference point on the
Panama indenter, we estimate 24 mm=year of active rightlateral displacement across the Caldas tear. The hypothesis
of lateral homogeneity of the crust, constant displacement rate,
and a seismic offset along Caldas tear of 240 km, would
suggest an 10:0 Ma initiation of the Panama arcColombia
collision (240 km=24 mm=year). Geologic field observations in the Panama-arc (Coates et al., 2004) suggest that the
age for the initiation of the Panama arc collision with northern
South America occurred between 12.8 and 7.1 Ma, which is
consistent with our estimated 10:0 Ma initiation of the tear
propagation. In addition, because the origin of the eastwest
Sandra ridge occurs between 9 and 12 Ma (Defant et al., 1992;
Lonsdale, 2005), and this structure is collinear with the Caldas
tear, we propose that the right-lateral lineament defined by the
Caldas tear and the Sandra ridge, constitutes a major area of
lithospheric weakness along the southern flank of the Panama
arc-indenter. Although there is no evidence for recent activity
of the Sandra ridge due to lack of near-bottom instrumentation
in the Pacific offshore of Colombia and Panama, recent earthquakes and the adakite magmatism along the Caldas tear may
indicate that this lineament localizes upper crustal fault conduits that allowed the upward migration of magmatic fluids
and are associated with elevated geothermal anomalies. Offset
of intermediate- to deep-seismicity that defines the Caldas tear
is also coincident with inactive volcanoes of adakite composition and geothermal anomalies (Fig. 2). The adakites of the
Ruiz volcanic complex with ages ranging from 0.970.05 and
1.80.6 Ma (Borrero et al., 2009) and the PaipaIza volcanic
complex dated 1.92.5 Ma (Pardo et al., 2005) are a likely
consequence of magmatism related to this progressive
slab tear. Low ratio 87 Sr=87 Sr (0.705) and the presence of
xenoliths of metamorphic rocks in this last volcanic complex
(J. M. Jaramillo, personal comm., 2012) support the proposed
break-off interpretation south of the Bucaramanga nest and
east of the Caldas tear.
Surficial evidence of this lithospheric tear are restricted
to presence of mineral deposits, hydrocarbon occurrences,
and some geomorphological anomalies. High-grade mineral
deposits of platinum, gold, and copper exploited in the mining areas called Condoto (Tistl, 1994), Marmato (Ordoez,
2001), Quinchia (INGEOMINAS, 1999), La Colosa (GilRodriguez, 2010), and Cerro de Cobre (McLaughlin and
Arce, 1970) are near, or collinear with, the Caldas tear and
exhibit ages ranging between 6 and 20 Ma (see blue hexagons on map in Fig. 10a). In addition, significant changes in
distribution and trend of oil and gas seeps, as well as the
hydrocarbon fields on both sides of the Caldas tear, suggest
that this structure may also affect the geometry of several
sedimentary basins (e.g., Llanos foreland, Eastern Cordillera,
Middle Magdalena Valley). But likely the most prominent
geomorphological evidence is coming from hydraulic behavior of the main rivers that cross the south-to-north-flowing

2041

Magdalena and Cauca rivers of northwestern South America.


After flowing 200 km from their sources, these rivers occupy
broad river valleys. Downstream, rivers passing the Caldas
tear lineament change their morphology from broad valleys
to steeper relief gorges near the surface projection of the
Caldas tear. The Cauca River near Supia town (Fig. 10a,b)
reduces to a narrow channel only 150 m wide whereas the
Magdalena River near the town of Honda reduces to a 250 m
wide channel (Fig. 10a,c). In contrast to the broader, 42 km
wide valleys observed upstream (e.g., Bolivar and Guamo
towns in Fig. 10a,d), in both areas the steeper gradient and
more narrow rivers produce rapids that are an impediment to
navigation. The narrowness of these rivers favored the downstream economic development of urban settlements such as
Honda and Supia from Spanish colonial times. However, the
quake that occurred on 16 June 1805 that destroyed Honda
and other nearby towns shows us that the Caldas tear could
form a major source of earthquakes and seismic hazard in
this region.
The Caldas tear may localize large crustal earthquakes,
including the recent strong-motion activity associated with
the Tauramena earthquake (19 January 1995, M w 6.5).
Two types of focal mechanisms have been proposed for this
event, one of which suggests an eastwest-oriented right
lateral movement (Dimate et al., 2003; Fig. 2). Similarly,
seismic activity in the area of the Armenia earthquake (25
January 1999, M w 6.2) shows eastwest alignment of aftershocks with the Caldas tear. This long right Caldas tear does
not explain the Quetame earthquake (24 May 2008, mL 5.7)
with left-lateral slip, in which the Caldas tear could have
played an important role for controlling the propagation
of northsouth-trending faults (such as this event or any related to the Llanos fault system) in a similar way as has been
suggested in the southwest Colombia margin where transverse faults reduce coupling between adjacent segments
(Collot et al., 2004).
We removed the crustal earthquakes from the database
of events located by the CNSN between 1993 and 2012 in
order to illustrate the upper surface of the subducted slabs
beneath Colombia (Fig. 11a). The 3D model of this surface
images the Caldas tear and flat-slab subduction geometry
that we relate to the presence of the Panama arc-indenter (Ramos and Folguera, 2009). The northern border of the indenter
becomes diffuse and does not appear to form a distinctive
tear as seen south of the indenter (see dashed line in Fig. 11a).
The Caribbean plate changes from flat to steep subduction
toward the northeast. South of the Caldas tear (2:5 N),
there is a shift to a new pattern of intermediate and deep
seismicity associated with flat subduction along with the
development of a broad area of active volcanoes in southern
Colombia and northern Ecuador. Figure 11b presents a conceptual model that explains the kinematic role of the
Panama arc-indenter whose southern boundary is defined
by the Sandra ridge and the Caldas tear. In this model the
coupling of the Panama arc with the Caribbean plate could
generate a change in buoyancy of the lithospheric system and

2042

C. A. Vargas and P. Mann

Figure 10.

Surficial evidences of the Caldas tear related to mineral deposits, hydrocarbon occurrences, and geomorphological anomalies.
(a) Blue hexagons, map of distribution of high-grade mineral deposits of platinum, gold, and copper: (1) Condoto, (2) Marmato, (3) Quinchia,
(4) La Colosa, and (5) Cerro de Cobre. Black dots, oil and gas seepages. Purple circles are giant hydrocarbon fields. Blue stars are other oil
and gas fields. These hydrocarbon occurrences suggest that the Caldas tear also is affecting the geometrical configuration of several sedimentary basins. White stars, hydraulic anomalies of the Cauca and Magdalena rivers on the Caldas tear (rapids on the Supia and the Honda).
White circle and square are places upstream the rivers where there are broad valleys (Bolivar and Guamo). (b) Rapids of the Cauca River near
Supia town that overlies the Caldas tear. (c) Rapids of the Magdalena River near Honda town that overlies the Caldas tear. (d) Broad valley
observed upstream of the Cauca River near Bolivar town. Similar landscape is observed in Guamo town where the valley width of the
Magdalena River reaches > 40 km wide.

consequently the northern region of indentation has conditions that favor flat subduction. In the east, the Caribbean
plate suddenly changes its subduction angle and produces
a break off of the slab around the location of the Bucaramanga nest. South of the Caldas tear, the Nazca plate is subducting beneath the South American plate with a steeper
angle and a faster rate. The Cauca nest is a combined product
of eastward decoupling of plates along the Caldas tear and
flexure during the subduction process.
A corollary of our model for the Panama indenter and
the formation of the Caldas tear is the eastward indentation of
geologic features. Inspection of the Map of Quaternary faults
and folds of Colombia (Paris et al., 2000) suggests that some

branches of the Romeral fault zone south of the Caldas tear


show right-lateral offset from parallel faults including the
Palestina, Cimitarra, MulatoGetudo, Honda, or Bituima
faults to the north. Other faults with southwestnortheast
trend, such as the Ibague and Garrapatas, could be part of
a transverse strike-slip fault at the surface level overlying
the Caldas tear (Fig. 1a). These results raise new questions
about the regional evolution in northwestern South America.
For example, the Great Arc of the Caribbean has been defined along the south Caribbean region but disappears once
it enters the Guajira basin and the Santa Marta Massif north
of Colombia (Ostios et al., 2005). As a consequence of the
Panama-arc collision, it is possible that this regional feature

Tearing and Breaking Off of Subducted Slabs as the Result of Collision of the Panama Arc-Indenter

2043

Figure 11.

(a) Seismic surface estimated by interpolation and filtering of 68;000 local earthquakes (h 10:0 km). Blue lines, shore line
of northwest South America. Bold black lines, limits of the convergent margins. Bold red lines, the southern border of the Panama indenter
that includes the Sandra ridge and the Caldas tear. Bucaramanga nest is related to a break-off process that is propagating toward the southwest. Triangles, red (active) and green (inactive) volcanoes. Orange dashed lines, wireframe model suggested for indicating the subduction
geometry of the Caribbean plate. (b) Schematic 3D model suggesting flat subduction on the northern side of the weakness zone formed by the
Sandra ridge and the Caldas tear. Caribbean plate suddenly changes its subduction angle and promotes a break off of the slab around the
location of the Bucaramanga nest. South of the weakness zone, the Nazca plate subducts beneath the South American plate with a steeper
angle and faster displacement. Probably the Cauca nest is the combined product of eastward decoupling of plates along the Caldas tear as well
as flexion and discrete movements of the plate during subduction. The Murind earthquake nest located in proximity to the Panamanian and
Colombian border, may be response to convergent accommodation between the Panama arc-indenter and the Caribbean plate.

has been offset and displaced eastward by the Panama


indenter.

Conclusions
The eastward-directed collision of the buoyant Panama
arc-indenter with northwestern South America produces
a distinctive V-shaped pattern of crustal deformation and

widens the northern Andes in Colombia and Venezuela


(Fig. 1b). The Panama collision initiated 10 Ma and continues to be active as shown by GPS data.
The southern edge of the Panama indenter is associated
with the proposed Caldas slab tear at latitude 5:6 N. This
tear extends for 240 km in an eastwest direction and
is collinear with the 912 Ma, now extinct, eastwestoriented Sandra oceanic spreading ridge on the unsubducted

2044

C. A. Vargas and P. Mann

oceanic Nazca plate to the west (Fig. 1b). We postulate that


the Caldas tear may have formed as a zone of lithospheric
weakness along the now subducted part of the inactive
Sandra spreading ridge.
The 240 km long Caldas tear is a narrow, eastwesttrending boundary between two subducted slabs of different dip. The northern zone is the down-dip extension of the
Panama arc, has a shallower dip, and is not associated with
active arc volcanism. The southern zone has a steeper dip
and is associated with an active volcanic front (Fig. 11a,b).
The Caldas tear also localizes angular difference of the
subduction geometry in both geophysical sections presented in this work (Figs. 5 and 6). The lineament defined
by the 240 km long offset of the deep seismicity along
5:6 N; the eruption of the northsouth belt of active
volcanism and presence of extinct magmatic bodies with
adakite composition along the lineament; the occurrence
of high-grade mineral deposits and geothermal gradient
anomalies; the different patterns associated with oil and
gas manifestations; the distribution of major oil and gas
deposits north and south of the Caldas tear as well as the
GPS measurements and strong-motion events with rightlateral movements, support the existence of the Caldas tear.

Data and Resources


Waveforms and preliminary hypocentral solutions of
the Colombian territory were supplied by the Geological Survey of Colombia (INGEOMINAS). Bouguer and magnetic
data provided for the National Hydrocarbons Agency of
Colombia (http://www.anh.gov.co/es/index.php?id=82, last
accessed November 2012) were used for modeling geologic
sections using the GM-SYS profile module of the Oasis
Montaj software (Geosoft, 2010). This software calculates
the gravity and magnetic model response based on the methods of Talwani et al. (1959), and Talwani and Heirtzler
(1964). GM-SYS uses a 2D, flat-earth model for the gravity
and magnetic calculations. Each structural unit or block
extends to plus and minus infinity in the direction perpendicular to the profile. The earth is assumed to have topography but no curvature. The model also extends plus and minus
30,000 km along the profile to eliminate edge effects. The
90 m elevation topographic information used for the gravity
modeling is available from the CGIAR-CSI SRTM 90 m
database (http://srtm.csi.cgiar.org, last accessed November
2012). Focal mechanisms reported by NEIC were used in
this work (http://earthquake.usgs.gov/earthquakes/eqarchives/
sopar/, last accessed November 2012).

Acknowledgments
This work was partially funded by the industry sponsors of the CBTH
project of the University of Houston and by fellowship support from the
University of Texas at Austin. Earthquake, gravity, magnetic, seismic,
and geothermal data were kindly provided by Agencia Nacional de Hidrocarburos, Universidad Nacional de Colombia, INGEOMINAS, and the following
research projects: 1233-333-18664, Contract 201-2006 (COLCIENCIAS);

1233-487-25728, Contract 589-2009 (COLCIENCIAS); CGL2005-04541C03-02 and CGL2008-00869/BTE (UPC, MICCIN, FEDER). We also thank
the Associate Editor Heather DeShon, and two anonymous reviewers for their
helpful reviews of this paper.

References
Adamek, S., C. Frohlich, and W. Pennington (1988). Seismicity of the
Caribbean-Nazca boundary: Constraints on microplate tectonics of
the Panama region, J. Geophys. Res. 93, 20532075.
Aki, K. (1969). Analysis of the seismic coda of local earthquakes as
scattered waves, J. Geophys. Res. 74, 615631.
Aki, K., and B. Chouet (1975). Origin of coda waves: Source, attenuation
and scattering effects, J. Geophys. Res. 80, 615631.
Badi, G., E. Del Pezzo, J. M. Ibanez, F. Bianco, N. Sabbione, and M. Araujo
(2009). Depth dependent seismic scattering attenuation in the Nuevo
Cuyo region (southern central Andes), Geophys. Res. Lett. 36,
no. L24307, doi: 10.1029/2009GL041081.
Borrero, C., L. M. Toro, M. Alvaran, and H. Castillo (2009). Geochemistry
and tectonic controls of the effusive activity related with the
ancestral Nevado del Ruiz volcano, Colombia, Geofsc. Int. 48,
no. 1, 149169.
Calais, E., and P. Mann (2009). A combined GPS velocity field for the
Caribbean plate and its margins (abstract #G33B-0657), American
Geophysical Union, (Fall Meet.), G33B-0657.
Cediel, F., R. P. Shaw, and C. Caceres (2003). Tectonic assembly of the
northern Andean block, in The circum-Gulf of Mexico and the
Caribbean: Hydrocarbon habitats, basin formation, and plate tectonics,
AAPG Memoir 79, 134.
Coates, A. G., L. S. Collins, M. P. Aubry, and W. A. Berggren (2004). The
geology of the Darien, Panama, and the late Miocene-Pliocene collision of the Panama arc with northwestern South America, Bull. Geol.
Soc. Am. 116, nos. 1112, 13271344.
Collot, J. Y., B. Marcaillou, F. Sage, F. Michaud, W. Agudelo, P. Charvis, D.
Graindorge, M. A. Gutscher, and G. D. Spence (2004). Are rupture
zone limits of great subduction earthquakes controlled by upper plate
structures? Evidence from multichannel seismic reflection data acquired across the northern Ecuadorsouthwest Colombia margin, J.
Geophys. Res. 109, B11103, doi: 10.1029/2004JB003060.
Corredor, F. (2003). Seismic strain rates and distributed continental deformation in the northern Andes and three-dimensional seismotectonics
of northwestern South America, Tectonophysics 372, 147166.
Corts, M., and J. Angelier (2005). Current states of stress in the northern
Andes as indicated by focal mechanisms of earthquakes, Tectonophysics 403, 2958.
Defant, M. J., T. E. Jackson, M. S. Drummond, J. Z. De Boer, H. Bellon, M.
D. Feigenson, R. C. Maury, and R. H. Stewart (1992). The geochemistry of young volcanism throughout western Panama and southeastern
Costa Rica: An overview, J. Geol. Soc. 149, 569579.
Dimate, C., L. A. Rivera, A. Taboada, B. Delouis, A. Osorio, E. Jimenez, A.
Fuenzalida, A. Cisternas, and I. Gomez (2003). The 19 January 1995
Tauramena (Colombia) earthquake: Geometry and stress regime,
Tectonophysics 363, 159180.
Faul, U. H., and I. Jackson (2005). The seismological signature of temperature and grain size variations in the upper mantle, Earth Planet. Sci.
Lett. 234, 119134.
Frohlich, C. (2006). Deep Earthquakes, Cambridge Univ. Press, Cambridge,
U.K., 574 pp.
Frohlich, C., K. Kadinsky-Cade, and S. D. Davis (1995). A reexamination of
the Bucaramanga, Colombia, earthquake nest, Bull. Seismol. Soc. Am.
85, 16221634.
Geosoft (2010). GM-SYS profile module, Oasis montaj processing and
mapping.
Gil-Rodriguez, J. (2010). Igneous Petrology of the Colosa Gold-Rich Porphyry
System (Tolima, Colombia), M. Sc. Thesis, University of Arizona, 51 pp.
Haskov, J., S. Malone, D. McClury, and R. Crosson (1989). Coda-Q for the
State of Washington, Bull. Seismol. Soc. Am. 79, 10241038.

Tearing and Breaking Off of Subducted Slabs as the Result of Collision of the Panama Arc-Indenter
Hinze, W. J., C. Aiken, J. Brozena, B. Coakley, D. Dater, G. Flanagan,
R. Forsberg, T. Hildenbrand, G. R. Keller, J. Kellogg, R. Kucks,
X. Li, A. Mainville, R. Morin, M. Pilkington, D. Plouff, D. Ravat,
D. Roman, J. Urrutia-Fucugauchi, M. Veronneau, M. Webring, and
D. Winester (2005). New standards for reducing gravity data: The
North American gravity database, Geophysics 70, no. 4, J25J32.
Hoerl, A. E., and R. W. Kennard (1970). Ridge regression: Biased estimation
for nonorthogonal problems, Technometrics 12, 5567.
Hoerl, A. E., R. W. Kennard, and K. F. Baldwin (1975). Ridge regression:
Some simulation, Comm. Stat. 4, 105123.
INGEOMINAS (1999). Risaralda, Department Mining Inventory, Internal
report (in Spanish).
Jarvis, A., H. I. Reuter, A. Nelson, and E. Guevara (2008). Hole-filled SRTM
for the globe version 4, available from the CGIAR-CSI SRTM 90m
Database (http://srtm.csi.cgiar.org; last accessed July 2012).
Karato, S., and H. Jung (1998). Water, partial melting and the origin of the
seismic low velocity and high attenuation zone in the upper mantle,
Earth Planet. Sci. Lett. 157, 193207.
Lacruz, J., A. Ugalde, C. A. Vargas, and E. Carcol (2009). Coda-wave
attenuation imaging of Galeras Volcano, Colombia, Bull. Seismol.
Soc. Am. 99, no. 6, 35103515.
Levenberg, K. (1944). A method for the solution of certain nonlinear
problems in least squares, Q. Appl. Math. 2, 164168.
Lonsdale, P. (2005). Creation of the Cocos and Nazca plates by fission of the
Farallon plate, Tectonophysics 404, 237264.
Lopez, C. (2004). Upper crust models of Colombia, INGEOMINAS, Bogota,
internal Rept., 64 pp.
Malin, P. E. (1978). A first order scattering solution for modeling lunar and
terretrial seismic coda, Ph.D. Dissertation, Princeton University,
Princeton, New Jersey.
Marquardt, D. W. (1963). An algorithm for least squares estimation of
nonlinear parameters, J. Soc. Ind. Appl. Math. 11, 5567.
Maus, S., S. Macmillan, T. Chernova, S. Choi, D. Dater, V. Golovkov, V. Lesur,
F. Lowes, H. Luhr, W. Mai, S. McLean, N. Olsen, M. Rother, T. Sabaka,
A. Thomson, and T. Zvereva (2005). The 10th generation international
geomagnetic reference field, Phys. Earth Planet. Int. 151, 320322.
Maus, S., T. Sazonova, K. Hemant, J. D. Fairhead, and D. Ravat (2007).
National Geophysical Data Center candidate for the World Digital
Magnetic Anomaly Map, Geochem. Geophys. Geosyst. 8, Q06017,
doi: 10.1029/2007GC001643.
McLaughlin, D. H., and M. Arce (1970). Economic geology of the Zipaquira
quadrangle and adjoining area, Department of Cundinamarca,
Colombia, U.S. Geol. Surv. Open-File Rept., 141 pp.
Mitchell, B. J., and L. Cong (1998). Lg coda Q and its relation to the
structure and evolution of continents: A global perspective, Pure Appl.
Geophys. 153, 655663.
Mitchell, B. J., Y. Pan, J. Xie, and L. Cong (1997). Lg coda variation across
Eurasia and its relation to crustal evolution, J. Geophys. Res. 102,
22,76722,779.
Mukhopadhyay, S., and J. Sharma (2010). Attenuation characteristics of
GarwhalKumaun Himalayas from analysis of coda of local earthquakes, J. Seismol. 14, doi: 10.1007/s10950-010-9192-9.
National Hydrocarbons Agency of Colombia (2010). Total Bouguer Anomalies Map (MABT) of Colombia, National Hydrocarbons Agency of
Colombia (ANH).
ODoherty, K. B., C. J. Bean, and J. McCloskey (1997). Coda wave imaging
of the Long Valley caldera using a spatial stacking technique, Geophys.
Res. Lett. 24, 15451550.
Ojeda, A., and J. Havskov (2001). Crustal structure and local seismicity in
Colombia, J. Seismol. 5, no. 4, 575593.
Oliver, M. A., and R. Webster (1990). Kriging: A method of interpolation
for geographical information system, Int. J. Geogr. Inf. Syst. 4,
no. 3, 313332.
Ordoez, C. O. (2001). Caracterizao isotpica Rb-Sr e Sm-Nd dos
principais eventos magmticos nos Andes Colombianos, Ph.D. Thesis,
Instituto de Geocincias, Universidade de Brasilia, Brasilia, 197 pp. (in
Portuguese).

2045

Ostios, M., F. Yoris, and H. G. Av Lallemant (2005). Overview of the southeast Caribbean-South American plate boundary zone, Spec. Pap. Geol.
Soc. Am. 394, 5389.
Pardo, N., H. Cepeda, and J. M. Jaramillo (2005). The Paipa volcano,
Eastern Cordillera of Colombia, South America: Volcanic stratigraphy,
Earth Sci. Res. J. 9, no. 1, 318.
Paris, G., R. L. Dart, and M. N. Manchette (2000). Map of Quaternary
faults and folds of Colombia and its offshore regions, U.S. Geol. Surv.
Open-File Rept., Scale 1:2,500,000.
Priestley, K., and D. McKenzie (2006). The thermal structure of the
lithosphere from shear wave velocities, Earth Planet. Sci. Lett. 244,
285301.
Pulli, J. J. (1984). Attenuation of coda waves in New England, Bull. Seismol.
Soc. Am. 74, 11491166.
Ramos, V. A., and A. Folguera (2009). Andean flat-slab subduction through
time, Spec. Publ. Geol. Soc. Lond. 327, 3154.
Sandwell, D. T., and W. H. F. Smith (2009). Global marine gravity
from retracked Geosat and ERS-1 altimetry: Ridge segmentation
versus spreading rate, J. Geophys. Res. 114, B01411, doi: 10.1029/
2008JB006008.
Schneider, J. F., W. D. Pennington, and R. P. Meyer (1987). Microseismicity
and focal mechanisms of the intermediate-depth Bucaramanga
nest, Colombia, J. Geophys. Res. 92, no. B13, 13,91313,926, doi:
10.1029/JB092iB13p13913.
Singh, S., and R. B. Herrmann (1983). Regionalization of crustal coda Q in
the continental United States, J. Geophys. Res. 88, 527538.
Taboada, A., L. A. Rivera, A. Fuenzalida, A. Cisternas, H. Philip,
H. Bijwaard, J. Olaya, and C. Rivera (2000). Geodynamics of the
Northern Andes; subductions and intracontinental deformation
(Colombia), Tectonics 19, 787813.
Talwani, M., and J. R. Heirtzler (1964). Computation of magnetic anomalies
caused by two-dimensional bodies of arbitrary shape, in Computers in
the mineral industries, Part 1: Stanford Univ. Publ., Geological
Sciences, G. A. Parks (Editor), Vol. 9, 464480.
Talwani, M., J. L. Worzel, and M. Landisman (1959). Rapid gravity
computations for two-dimensional bodies with application to the
Mendocino submarine fracture zone, J. Geophys. Res. 64, 4959.
Tistl, M. (1994). Geochemistry of platinum-group elements of the zoned
ultramafic Alto Condoto complex, northwest Colombia, Econ. Geol.
89, no. 1, 158167.
Tittgemeyer, M., F. Wenzel, T. Ryberg, and K. Fuchs (1999). Scales of
heterogeneities in the continental crust and upper mantle, Pure Appl.
Geophys. 156, 2952.
Trenkamp, R., J. N. Kellogg, J. T. Freymueller, and H. Mora (2002). Wide
plate margin deformation, southern Central America and northwestern
South America, CASA GPS observations, J. S. Am. Earth Sci. 15,
no. 2, 157171.
van der Hilst, R. D., and P. Mann (1994). Tectonic implications of tomographic images of subducted lithosphere beneath northwestern South
America, Geology 22, 451454.
Vargas, C. A., C. Alfaro, L. A. Briceo, I. Alvarado, and W. Quintero (2009).
Mapa Geotrmico de Colombia, 2009, in Proceedings of X Simposio
Bolivariano Exploracin Petrolera en Cuencas Subandinas, Colombia
(in Spanish).
Vargas, C. A., P. Mann, C. Gomez, L. A. Briceo, and C. Rey (2010). TransAndean mega-regional seismic reflection line extending from the
Caribbean coast to Cordillera Oriental of Colombia: Implications for
hydrocarbon exploration, AAPG Annual Convention: Unmasking the
Potential of Exploration & Production, New Orleans, Abstract 90104.
Vargas, C. A., L. G. Pujades, and L. A. Montes (2007). Seismic structure of
south-central Andes of Colombia by tomographic inversion, Geofsc.
Int. 46, no. 2, 117127.
Vargas, C. A., A. Ugalde, L. G. Pujades, and J. A. Canas (2004). Spatial
variation of coda wave attenuation in northwestern Colombia,
Geophys. J. Int. 158, 609624.
Xie, J. (2002). Lg Q in the eastern Tibetan Plateau, Bull. Seismol. Soc. Am.
92, 871876.

2046
Yang, Y., D. W. Forsyth, and D. S. Weeraratne (2007). Seismic attenuation
near the East Pacific Rise and the origin of the low-velocity zone, Earth
Planet. Sci. Lett. 258, 260268, doi: 10.1016/j.epsl.2007.03.040.
Zarifi, Z. (2006). Unusual subduction zones: Case studies in Colombia and
Iran, Ph.D. Thesis, University of Bergen, Norway, 78 pp.
Department of Geosciences
Universidad Nacional de ColombiaSede Bogot
Carrera 45 No 26-85Edificio Manuel Ancizar
Bogot D.C.Colombia, 111321
cavargasj@unal.edu.co
(C.A.V.)

C. A. Vargas and P. Mann


Department of Earth and Atmospheric Sciences
University of Houston
4800 Calhoun Boulevard
Houston, Texas 77004
pmann@uh.edu
(P.M.)

Manuscript received 3 November 2012

Potrebbero piacerti anche