You are on page 1of 22

Bulletin of the Seismological Society of America, Vol. 103, No. 3, pp. 2025–2046, 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 east-
ward- and southeastward-dipping, subducted slabs to depths of 315 km beneath the
surface of Colombia in northwestern South America. These cross-sectional interpre-
tations 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 ther-
mal structure of the lithospheric system. Mapping of earthquake-defined Benioff
zones combined with tomographic imaging reveals the presence of an ∼240 km long
east–west-striking slab tear, named here the Caldas tear. The proposed Caldas tear
separates a zone of shallow, 20°–30°-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, 30°–40°-dipping, slab adjacent
to the eastern Pacific Ocean that is associated with an active north–south 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 north–south-trending active volcanic arc. In addition to the for-
mation 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 Na- Regional compilations of Global Positioning Systems
tional Seismological Network (CNSN) show an ∼240 km (GPS) data provide a quantitative tectonic framework for
long, right-lateral offset of intermediate to deep events with understanding the widespread crustal effects of the Panama
azimuth of 102° (Fig. 1a,b). We infer this discontinuity in arc collision on large areas of northwestern South America
earthquakes to be a major slab tear which we have named (Calais and Mann, 2009; Fig. 1a). GPS vectors in western
the Caldas tear based on the location in the Caldas department Colombia show a marked decrease in velocities consistent
of Colombia and the alignment of fault-related surface features with the ongoing collision of the Panama arc with north-
(e.g., volcanism, faulting, mineral deposits, geothermal anoma- western South America along a north–south-trending suture
lies, etc.). Using the distribution of earthquakes > 80 km, zone roughly parallel to the international boundary between
Ojeda and Havskov (2001) proposed that the discontinuity Panama and Colombia (Adamek et al., 1988; Trenkamp
along the Caldas tear represented a boundary between two sub- et al., 2002; Corredor, 2003; Fig. 1a). The east–west direc-
ducted slabs with differing dips and strikes: the northern sub- tion of GPS vectors shows that the effects of east–west
duction zone, called the Bucaramanga subduction zone, has shortening and indentation related to the collision of the Pan-
a shallower dip (27°) and more northeasterly strike, and the ama arc remains relatively constant over a large, V-shaped,
southern, called the Cauca subduction zone, has a steeper fault-bounded area of Colombia due east of the Panama arc-
dip (35°–40°) and a more northerly strike (Fig. 1a). indenter (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 se-
lective distribution of hypocentral solutions of ∼30;000 earthquakes extracted from the entire catalog of the CNSN (∼102;000 events) during
1993–2012 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, Mulato–Getudo 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, Paipa–Iza 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; pink-
colored 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 20°–30° in the area north of ∼5:6° N (Ojeda and Havskov,
plate motions related to northward tectonic escape of the 2001; Vargas et al., 2007).
Maracaibo block into the southern Caribbean (Trenkamp Two nests of concentrated intermediate-depth earth-
et al., 2002). In contrast to this fairly uniform GPS velocity quakes are present beneath Colombia (Fig. 1a). The Bucara-
field of deformed crustal rocks produced by the Panama col- manga earthquake nest (BN) is found at a depth of ∼160 km
lision, underlying, eastward-dipping slabs change abruptly on the down-dip extension of the southern (Bucaramanga)
across the Caldas tear from dip angles of 30°–40° between subduction zone and has an estimated volume dimension
latitudes ∼3:0°–5:6° N in southern Colombia, to dip angles of 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 2027

1995). Previous tectonic interpretations of the origin of the and a high-resolution seismic profile, seeking to define the
Bucaramanga nest vary from a zone of two slabs in contact geometry of the Caldas tear and its geotectonic implications
(van der Hilst and Mann, 1994), two slabs overlapping in the northwestern corner of South America.
(Taboada et al., 2000), or a single slab undergoing extreme
bending (Cortés and Angelier, 2005) all occurring in the
boundary area of the subducted northern (Bucaramanga) Hypocentral Solutions and Estimation
and southern (Cauca) subduction zones (Fig. 1a). The Cauca of the Coda-Wave Attenuation
intermediate-depth earthquake nest (CN) is located ∼400 km A catalog has been compiled of ∼102;000 earthquake
southwest of the Bucaramanga nest on the trend of our pro- locations calculated by the CNSN during the period 1993–
posed Caldas tear and has been previously interpreted 2012 (mL ≤ 6:8). Hypocentral solutions were estimated by
by Cortés and Angelier (2005) as a bend in the slab in this using a seismological array of 17 short-period instruments
area (Fig. 1a). There is no clear consensus among seismol- (T  1 s) of the CNSN and complemented by 13 stations
ogists for the tectonic interpretation of the two concentrated associated with local volcanic monitoring systems and also
Colombian intermediate earthquake nests (Frohlich, 2006; foreign networks (Panama, Ecuador, and Venezuela). Final
Zarifi, 2006). solutions were calculated with the HYPOCENTER program
The Caldas tear defines the northern limit of the active and the velocity model proposed by Ojeda and Havskov
volcanic front of the northern Andes that has formed as a (2001). Then, 9338 waveforms associated with 7645 regional
consequence of the steeper subduction of oceanic slab of earthquakes (3:0 ≤ mL ≤ 6:5; 1993–2012) were selected for
normal thickness of the Nazca plate (Fig. 1a). Moreover, estimating the decay rate of the coda amplitudes (Q−1c , coda
associated with active and inactive volcanoes, the east–west attenuation). The selected events, on basis of a significant
projected surface trace of the Caldas tear localizes an east– number of stations that recorded them (Table 1), have epicen-
west alignment of some unusual volcanic rocks including tral distances to stations ranging between 22.6 and 690.0 km
adakites (Borrero et al., 2009; Fig. 1a). Other volcanic rocks and depths varying between 0 and 222.0 km. Figure 2 shows
in the vicinity of the east–west-trending Caldas tear include all events used for the Q−1c plotted on a map of northwestern
the Plio-Pleistocene Paipa-Iza volcanic complex in the South America along with tectonically significant earth-
Eastern Cordillera of Colombia and the Romeral and San quake focal mechanisms including the Quindio, Quetama,
Diego volcanoes (Pardo et al., 2005). The presence of these and Tauramena events aligned along the Caldas tear and in-
east–west aligned volcanic rocks along with locally elevated termediate-depth focal mechanisms from the Bucaramanga
geothermal gradient values (Vargas et al., 2009) suggests that and Cauca nests.
the Caldas tear may penetrate the upper crust as a fault zone Estimations of the Q−1 c were done using the Single
and provide a conduit for the upward rise of magmas and Backscattering model proposed by Aki and Chouet (1975).
hydrothermal fluids produced by melting of the slabs on ei- This model assumes that the coda of a local earthquake is
ther side of the Caldas tear (Fig. 2). Furthermore, recent, composed of the sum of secondary S waves produced by
shallow-focus, strong motion events such as the Tauramena heterogeneities distributed randomly and uniformly within
earthquake (19 January 1995; M w 6.5, h  25  10 km), the lithosphere. The coda is the portion of a seismogram cor-
the Quindio earthquake (25 January 1999; M w 6.2, responding to back-scattered S-waves. The estimation of Q−1c
h  18:6 km), and the Quetame earthquake (24 May 2008; used the following equation:
mL 5.7; h  superficial) are all in alignment with the surface
 
trace of the Caldas tear.
−ωt
Previous tomographic studies using both local and 2
2gθjSωj Qc
Pω; t  e ; (1)
regional earthquakes of varying resolution have produced βt2
differing tectonic interpretations for slabs in this area (van
der Hilst and Mann, 1994; Taboada et al., 2000; Vargas where Pω; t is the time-dependent coda power spectrum, ω
et al., 2007). In this paper, we present the results of an in- is the angular frequency, β is the shear-wave velocity, jSωj is
tegrated geophysics and geologic study that improves the 3D the source spectrum, and gθ represents the directional scat-
imaging of the interactions between the eastward-moving tering coefficient. The gθ term has been defined as 4π times
Panama indenter and its collisional area in northwestern the fractional loss of energy by scattering, per unit travel dis-
South America. tance of primary waves, and per unit solid angle of the radi-
ation direction θ measured from the direction of primary wave
Data and Methods propagation. Using these assumptions, the geometrical spread-
ing is assumed to be proportional to r−1 , which only applies
The following sections describe data and procedures to body waves in a uniform medium. The source factor can be
used to estimate hypocentral solutions, the attenuation and its treated as a constant value for single frequency. According to
spatial distribution, the simultaneous 2D inversion of gravity equation (1), Q−1 c values can be obtained as the slope of the
and magnetic data, and the correlation of these results with least-squares fit of Lnt2 × Pω; t versus ωt, for t > tβ , where
focal mechanisms, geothermal gradients, geological maps, tβ 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 (Q−1 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 mates were performed with short-period records (T  1 s) at
mean squared amplitudes of the coda Aobs ω; t from band- several frequencies (Table 1). Then we chose estimates in the
pass-filtered seismograms around a center frequency. frequency band 1–3 (2  1) Hz because of the high availabil-
In order to take into account the deep structure using ity and geographical distribution of observations regarding
coda waves, β  4:64 km=s was assumed and calculated as other frequencies; and also best values of correlation coeffi-
a weighted average of S-wave speeds in the whole earth vol- cients, the root mean square (rms), and signal-to-noise ratio.
ume covered by the scattered waves (Badi et al., 2009). All In addition, it has been reported that the study region presents
records were filtered in a chosen frequency band and then Q−1
c values in this frequency band with errors ≤ 5% (e.g., see
used a coda-wave time window (W) of 20 s, starting from Vargas et al. (2004)). In general, errors seen along Q−1 c
2 × tβ ststart . The average lapse time, defined as tc  tstart  estimations are acceptable, for example, the rms of all esti-
W=2 ranges between 11.0 and 384.0 s. These large tc values mations vary between 0.07 and 1.79 (μ  0:24, σ  0:07)
ensure the sampling of regional structures. Attenuation esti- 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 (Q−1
c ) at Various Frequencies

Q−1
c × 10
3 Coefficient of Correlation rms Signal/Noise
Waveforms
Frequency Analyzed Min Mean±Std Max Min Mean±Std Max Min Mean±Std Max Min Mean±Std Max

2 9338 1.7 7.1±2.7 47.6 −0.97 −0.67±0.11 −0.5 0.07 0.24±0.07 1.79 2.0 16.2±47.5 985.5
8 5421 0.8 1.8±0.7 20.4 −0.96 −0.60±0.08 −0.5 0.18 0.32±0.06 2.67 2.0 11.0±25.6 842.6
12 4441 0.6 1.2±0.4 1.2 −0.94 −0.59±0.08 −0.5 0.15 0.35±0.06 1.89 2.0 9.7±21.1 600.9
16 3741 0.5 0.9±0.3 9.4 −0.95 −0.58±0.07 −0.5 0.17 0.35±0.09 4.13 2.0 9.0±18.1 315.7

Also presented are extreme values, averages, standard deviations, and quality parameters. Q−1 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
for all Q−1 −1
c observations suggested a high-frequency dependence of the attenuation in this region: Qc f  13:2  0:6 × 10 f
3 −0:970:06 .

−0:5 and −0:97 (μ  −0:67, σ  0:11). Table 2 presents a duction processes, are a likely site of large contrasts in
statistical summary of the main parameters related with the anelastic attenuation in the subducted lithospheric slabs.
estimation of Q−1 c values for 30 seismological stations. Given the ease to estimate Q−1 c , we can use this obser-
Figure 3a shows an example of typical waveform used dur- vation for highlighting regional structures related to contrasts
ing this analysis, as well the corresponding record filtered for in rigidity (e.g., crust or lithospheric plates). One way to
the chosen frequency band. The attenuation factor (Q−1 c ) is regionalize Q−1c is based on the work of Malin (1978) who,
suggested as a decay factor for the coda-wave amplitudes. expanding on the work of Aki (1969) and Aki and Chouet
Figure 3b presents histograms for Q−1 c values and their cor- (1975), realized that the first-order scatterers responsible for
relation coefficients, as well as distributions for the epicentral the generation of coda waves at any given tc can be located
distances, focal depths, and local magnitudes of the events on the surface of an ellipsoid with earthquake and station
analyzed. Figures 2 and 3b emphasize the presence of attenu- locations as foci (Singh and Herrmann, 1983). In the ellip-
ation contrasts in the region and at least two sources of soidal volume sampled by coda waves at any time t, Pulli
events, one of them surficial and dispersed, and the other (1984) defined the large semi-axis as a1  βt=2, and defined
located at an intermediate depth (linked to the nests of Buca- the small semi-axis as a2  a3  a21 − r2 =41=2 , where r is
ramanga and Cauca). the source–receiver distance of the ellipsoid. The horizontal
projection of this volume is coincident with the elliptical en-
Tomographic Imaging Using Coda-Wave Attenuation velope proposed by several authors as the area occupied by
the scattered energy of the coda-wave record (Mitchell et al.,
Mukhopadhyay and Sharma (2010) have proposed that 1997; Mitchell and Cong, 1998; Xie, 2002; Vargas et al.,
the variation of Q−1
c with tc shows a direct relationship with 2004). Following these observations and knowing the values
depth. These authors interpreted that Q−1 c values related to of tc , W, and β, it is possible to deduce the volumes of the
scattering processes that penetrate > 200 km depth are con- ellipsoidal shells where the seismic energy is scattered.
trolled by a crust and a relatively more transparent mantle. Hence Q−1 c values estimated with large tc correspond to large
These results support the idea that Q−1
c estimated with a large sampled volumes, and vice versa. Based on these hypotheses
tc is representative of a large sampled volume and large we can perform a generalized inversion for regionalizing
sampled depths. A corollary of this hypothesis is that the Q−1
c . For the purpose of the inversion, we define a geo-
Q−1 −1
c value must be near to the intrinsic absorption (Qi ) con- graphic grid around the seismic station that also encloses the
trolled mainly by the mantle. Following these ideas, Vargas hypocenter. We recognize that each measured Q−1 c is an aver-
et al. (2004) developed a regional tomographic study using age estimate Q−1 (or Q −1 ) for the volume as sampled by
av apparent
stations of the CNSN with relative large tc , (up to 180 s) and the ellipsoidal shell given by
found that the Q−1 c values are near to the Qi
−1 values for

V TOTAL X V Block-j
almost all stations, meaning that a large portion of the upper
mantle is being sampled. Other studies have suggested a  ; (2)
Qav j
Qj
direct relation between the thermal field and anelastic attenu-
ation (Faul and Jackson, 2005; Priestley and McKenzie,
2006; Yang et al., 2007). The physical meaning of this rela- where V Block-j is the fraction of volume (block) sampled by
tionship is not been completely understood, but Karato and the ellipsoidal shell with the true attenuation coefficient Q−1
j
Jung (1998) proposed that the higher water content in the (or Q−1true ). Assuming a constant S-wave velocity of propaga-
asthenosphere significantly reduces the seismic-wave veloc- tion, the volume traveled by a ray that leaves the hypocenter
ities through anelastic relaxation and increasing temperature. moves outward to the ellipsoidal shell as defined by the
Convergent margins such as Colombia, which involve large observation time of the coda and is scattered to the receiver,
amounts of sediments and water mobilized during the sub- can be determined. Equation (2) can be written as
2030 C. A. Vargas and P. Mann

(a)

(b)
σ
4000 µ=-0.67 σ=0.11 µ=155.0 σ=96.0
1600 4500
3500 1400 4000
3000 1200 3500
Frequency

Frequency

Frequency
2500 3000
1000
2500
2000 800
2000
1500 600
1500
1000 400 1000
500 200 500
0 0 0
0 10 20 30 40 50 -1 -0.9 -0.8 -0.7 -0.6 -0.5 0 100 200 300 400 500 600 700 800 900
Coefficient of correlation Epicentral distance
µ=87.7 σ=65.6 µ=3.1 σ=0.7
4000 3500
3500 3000
3000 2500
Frequency
Frequency

2500
2000
2000
1500
1500
1000 1000

500 500
0 0
0 50 100 150 200 250 0 1 2 3 4 5 6 7
Depth (km)

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 1–3
(2  1) Hz. Lower trace represents the decay envelope of the coda wave in a window of 20 s, starting from 2 × tβ s. Qc value was obtained
as the slope of the least-squares fit of Lnt2 × Pω; t versus ωt (dashed line with arrowheads), for t > tβ , where tβ represents the S-wave
travel time (Haskov et al., 1989). (b) Histograms for Q−1 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 where


           
Qav Q1 V TOTAL Q1 V TOTAL
1 1 V Block-i
1 V Block-n y xi  ai  :
 ; (3) Qav Qi V TOTAL
Qn V TOTAL
Then, a least-squares estimation of the xi is given by the
where the ratio V Block-j =V TOTAL is the volume fraction compact matrix equation AX  Y where A is a (k × n)
associated with the total scattered-wave travel path spent coefficient matrix, X is a (n × 1) vector, Y is a (k × 1) vector,
in the jth block. If the process is repeated for each station– and k is the number of station–hypocenter pairs. A linear
hypocentral pair, the entire region is sampled. Equation (3) is inversion of the matrix equation was formulated as an iter-
of the form atively damped least-squares technique (Levenberg, 1944;
Marquardt, 1963). The damping factor (σ), which adds
a1 x1      ai xi      an xn  y; (4) to the diagonal parameters of the matrix, was computed
Table 2
Seismological Stations of the CNSN Used in this Study
3
Q−1
c × 10 tc Coefficients of Correlation Epicentral Distances
Longitude Latitude Altitude Waveforms
Station (°) (°) (masl) Analyzed Min Mean±Std Max Min Mean±Std Max Min Mean±Std Max Min Mean±Std Max

ANIL −75.40 4.49 2300 248 2.9 6.8±1.6 17.2 12.1 89.3 ± 51.5 231 −0.94 −0.67±0.11 −0.50 26.7 166.3±90.2 413.6
BAR −73.18 6.58 1864 754 1.7 6.1±2.5 32.3 11.8 71.2±19.2 227 −0.94 0.67±0.11 −0.50 26.6 142.0±33.7 414.6
BCIP −79.84 9.17 61 5 5.5 6.9±2.0 12.2 104.7 126.4±15.7 146 −0.86 −0.69±0.10 −0.59 200.7 238.7±27.5 272.8
BET −75.44 2.68 540 66 2.9 6.6±3.0 14.9 16.7 68.1±56.3 246 −0.91 −0.68±0.11 −0.50 46.7 136.8±98.5 448.0
BRI −72.79 7.72 1427 46 2.9 6.0±3.5 47.6 13.1 81.6±36.7 150 −0.97 −0.66±0.13 −0.50 24.7 159.6±65.8 280.2
CHI −73.73 4.63 3140 724 2.7 6.8±3.0 20.0 11 89.4±53.3 263 −0.97 −0.69±0.11 −0.50 24.5 173.5±93.8 477.6
CLIM −77.89 0.94 4232 17 2.5 7.0±5.0 12.0 31.7 78.9±35.5 132 −0.91 −0.69±0.11 −0.53 73.0 155.5±62.1 247.6
COD −73.44 9.94 108 104 3.4 7.3±3.1 18.2 19.3 47.3±48.7 180 −0.95 −0.70±0.12 −0.50 33.8 100.2±85.3 332.2
CPAS −77.25 1.22 2620 8 5.1 8.8±4.6 15.9 16.6 14.1±10.2 38.4 −0.91 −0.72±0.12 −0.56 29.1 42.2±17.8 84.7
CRU −76.95 1.57 2761 330 2.5 6.7±2.9 16.1 17.6 79.2±60.2 339 −0.95 −0.69±0.11 −0.50 30.8 156.0±105.3 610.4
CTAB −74.2 5.01 3500 2 5.6 8.1±8.2 14.5 132.4 133±0.85 134 −0.86 −0.70±0.22 −0.55 249.2 250.2±1.48 251.3
CTAU −74.04 5.20 3868 5 4.3 8.0±6.7 28.6 52.8 96.2±31.2 119.1 −0.9 −0.73±0.12 −0.6 109.9 185.9±54.5 225.9
CUM −77.83 0.94 3420 234 3.0 7.0±3.2 22.7 12.9 82.7±57.9 384 −0.97 −0.69±0.12 −0.50 22.6 162.3±101.3 690.0
GCAL −77.42 1.21 2353 8 7.9 10.3±1.8 13.5 14.6 22.8±21.12 69.4 −0.91 −0.83±0.05 −0.74 25.6 57.4±37.0 139.0
GCUF −77.35 1.23 3800 56 3.7 7.4±2.7 18.2 22.0 52.3±40.1 152 −0.93 −0.70±0.11 −0.50 38.5 108.1±70.5 283.5
GUA −72.63 2.54 217 11 3.6 6.2±2.9 10.2 20.3 174.4±85.5 227 −0.86 −0.67±0.11 −0.51 53.0 322.7±149.7 415.3
HEL −75.53 6.19 2815 216 3.0 6.1±2.1 19.6 21.0 110.2±40.9 190 −0.96 −0.65±0.10 −0.50 36.8 210.0±71.5 350.2
MAL −77.34 4.01 75 391 2.4 6.2±0.0 15.9 14.9 65.1±37.5 323 −0.95 −0.67±0.11 −0.50 43.6 131.4±65.7 582.4
MARA −75.95 2.84 2207 78 3.0 5.9±2.4 18.2 17.9 87.6±65.2 259 −0.92 −0.65±0.11 −0.51 31.3 170.8±114.1 469.9
NOR −74.87 5.57 536 596 2.5 5.9±2.1 20.4 16.3 113.4±35.5 297 −0.94 −0.66±0.11 −0.50 28.5 215.8±62.1 537.4
OCA −73.32 8.24 1264 3304 2.0 6.0±2.3 21.7 16.4 102.8±23.1 287 −0.96 −0.66±0.11 −0.50 28.7 197.4±40.4 520.3
OTAV −78.45 0.24 3492 26 4.3 7.6±2.1 13.9 23.6 70.2±25.7 147 −0.88 −0.73±0.12 −0.53 58.8 140.3±44.9 274.1
PCON −76.4 2.33 4294 120 2.0 6.6±3.2 16.9 12.5 72.4±52.8 289 −0.96 −0.67±0.11 −0.50 39.4 144.1±92.4 522.6
PRA −74.89 3.71 468 313 2.5 6.1±2.7 19.6 17.6 101.1±61.3 259 −0.95 −0.67±0.11 −0.50 30.8 194.5±107.4 470.1
ROSC −74.33 4.86 3020 149 2.9 5.9±2.4 33.3 17.9 88.2±47.6 204 −0.95 −0.66±0.12 −0.50 31.3 171.8±83.3 373.6
RREF −75.35 4.9 4743 205 2.5 6.1±2.3 12.7 113.7 118.1±48.8 220 −0.92 −0.65±0.10 −0.50 41.5 224.1±85.4 402.7
RUS −73.08 5.89 3697 150 2.6 6.0±2.7 20.4 17.5 72.6±39.9 240 −0.94 −0.66±0.11 −0.50 30.6 144.5±69.7 437.9
SDV −70.63 8.88 1620 64 2.8 5.8±2.0 12.5 114.5 176.3±23.7 262 −0.90 −0.65±0.11 −0.50 217.9 326.0±41.4 476.5
SOL −77.41 6.23 38 850 3.1 7.3±3.1 25.6 17.2 64.6±39.8 305 −0.97 −0.71±0.11 −0.50 30.1 130.6±69.6 551.1
TOL −75.32 4.59 2577 258 2.5 6.3±2.6 20.0 19.8 100.4±56.1 248 −0.95 −0.67±0.11 −0.50 34.7 193.1±98.2 451.9

The stations detected 7645 earthquakes (1993–2012) that were used for estimating 9338 Q−1
c values in frequency band 1–3 (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.
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; For the 3D inverse problem, we estimated the fraction
Hoerl et al., 1975). According to this technique, the solution of volumes associated with each Q−1 av in order to establish
and resolution matrixes can be found for the following equation (3). Using all foci related to the events selected in
equations: this study, we assembled the compact matrix of equation (4)
and then we inverted the Q−1 av values using equation (5).
X  AT A  σ2 I−1 AT Y; (5) Finally, a spatial interpolation of the Q−1 av values was done
and based on the Kriging method (Oliver and Webster, 1990)
and presented on Mercator projection. Figure 4d shows
R  AT A  σ2 I−1 AT A: (6) the results of the inversion for the synthetic experiment based
on two domains of contrasting attenuation (Fig. 4b). This
Similar procedures for the Q−1
c imaging have been used experiment is comparable to a slab-tearing model for which
in previous works in order to establish a deterministic char- two zones with different angles of subduction, are related to
acterization of the heterogeneity in the lithosphere as an different attenuations. This hypothetical model linked a flat
alternative technique for traditional tomographic measure- subduction zone in the north (lower attenuation) and a nor-
ments (O’Doherty et al., 1997; Lacruz et al., 2009). mal subduction zone in the south (higher attenuation). In
general, the available data may allow detection of large struc-
Resolution and Reliability tures with significant contrasts of attenuation as much as
∼270 km depth. On the other hand, Figure 4e presents
A spatial inversion of attenuation of 32 × 32 × 8 the inversion for a chessboard based on two areas of contrast-
blocks with block dimensions of ∼60 km latitude× ing attenuation (Fig. 4c). This experiment suggests that
∼50 km longitude × ∼40 km thickness was designed in the available data may allow detection of smaller bodies
order to detect relevant structures in the region. We qualified (e.g., 100 km × 100 km × 60 km) with significant contrasts
the tomographic inversion by means of three approaches:
of attenuation, mainly in Colombia, and as much as ∼180 km
(a) hit count of ellipsoidal shells; (b) solving controlled tests;
depth.
and (c) mapping the diagonal elements of the resolution ma-
After several trials of accurate resolution and sta-
trix (RDE) by using equation (6). The hit count is a very
bility, the spatial inversion of attenuation with real data
rough quality estimation that only tells about summing up
was performed with the same grid (32 × 32 × 8 blocks).
the number of ellipsoids that contribute to the solution of
Figure 4f,g shows results of the tomographic estimation
a block. Based on this discretized volume, we mapped the
and their maps of the RDE at different depths. Because each
hit count with the available data in eight layers (0, 45, 90,
RDE shows the amount of independence in the solution of
135, 180, 225, 270, and 315 km; Fig. 4a). Although a large
one model parameter (RDE oscillates between 0 and 1), the
part of northwestern South America (including Colombia,
larger value of the RDE for one model parameter suggests a
western Venezuela, eastern Panama, and northern Ecuador)
is covered by ellipsoidal shells (over 500 crossings), it is in more independent solution for this parameter. 3D inversion
northern Colombia and northwestern Venezuela (71° W to presents higher RDE values (e.g., > 0:4) limited by the avail-
76° W; 5° N to 10° N; 0–180 km depth) where the largest ability and geographical concentration of Q−1 c values, indi-

number of shells run through the blocks (based on more than cating that the method is useful for areas for which a large
5000 hits per block). This approach emphasizes the impor- stacking of attenuation observations is present. From the
tance of the Bucaramanga nest data in the estimation of to- available earthquake data, the tomographic solution of the
mographic images. attenuation efficiently images large areas of the crust and
To incorporate the second approach, we evaluated the upper mantle of northwestern South America including
efficiency of the method described above solving the 3D di- Colombia, eastern Panama, and western Venezuela with
rect problem. The ellipsoidal shell volume associated with all sampling depths reaching > 315 km (Fig. 4f). Because the
foci (pairs of earthquake and station) were used to relate the ellipsoids related to deeper hypocentral solutions can sample
Q−1c values in two controlled test boards. The first one was
profounder volumes, the 3D inversion may detect the thermal
for appraising large domains of attenuation, for example, a influx from the mantle adequately.
zone with flat subduction in the north, and the other zone In order to infer the geometry of the Caldas tear and its
with normal subduction in the south (Fig. 4b). The second relationships with the adjacent Nazca and Caribbean plates,
test board evaluated the ability of the method to detect small we made two regional cross sections: (1) a northern sec-
anomalies by use of a typical chessboard (Fig. 4c). In tests, tion (AA0 , Fig. 1a) from the Caribbean plate to the Llanos
we assigned two values of Q−1 true that represent attenuation foreland basin of eastern Colombia and crossing the inter-
contrasts (1=70 and 1=200) into the 3D grid. Then we esti- mediate-depth earthquakes of the Bucaramanga nest; (2) a
mated the Q−1 −1
av values (or theoretical Qc values) for all el- southern section designed for imaging the corridor between
lipsoids (each one related to foci [earthquake–station]) by the Nazca plate and the Llanos basin and crossing the inter-
estimating the weighted average of Q−1 true involved in the vol- mediate-depth earthquakes of the Cauca nest (BB0 , Fig. 1a).
ume of each ellipsoid. 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 2033

available geophysical data for proper interpretation of the longitude ≤ 10:0 km; error in depth ≤ 5:0 km) were plotted on
tomograms along these sections. 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 interpreta-
Integrating Earthquake Data with Regional tion of the Trans-Andean megaregional seismic-reflection line
Seismic-Reflection Lines
that extends from the Caribbean coast to the Eastern Cordil-
Hypocentral solutions of the CNSN (rms < 0:3 s; lera of Colombia (Vargas et al., 2010) and to the northern to-
GAP < 200; stations ≥ 6; error in latitude ≤ 10:0 km; error in mographic 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 Q−1 c ob-
servations associated with 7645 regional earthquakes (mL ≤ 6:7; 1993–2012) in the frequency band 1–3 (2  1) Hz. (a) Hit count of ellip-
soidal shells, suggesting that the 3D inversion of Q−1c 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
method were Q−1 −1
c  1=200 and Qc  1=70. (c) Chessboard with smaller and regular distribution of attenuation contrasts. As in the previous
case, the contrasts of attenuation incorporated into the model were Q−1c  1=200 and Qc
−1  1=70. (d) 3D inversion of the synthetic model

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 Q−1 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 Buca-
ramanga 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 2035

Figure 4. Continued.
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 geo-
thermal 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 more anisotropic continental crust of northwestern South
northern (Bucaramanga) slab dipping at a shallow angle America.
to the southeast beneath northwestern Colombia (Fig. 7).
Although reflectors are difficult to distinguish in deeper areas Gravity and Magnetic Modeling
of the line, the overall distribution of deep reflectors dips east-
Coincident gravity and magnetic models were completed
ward in same amount as the subducted slab on the gravity and for this study based on regional information (Maus et al.,
magnetic model in Figure 5c. Deep reflections are concen- 2007; Sandwell and Smith, 2009; National Hydrocarbons
trated within what we interpret as the lower crust whereas Agency of Colombia, 2010; Figs. 5b and 6b). The gravity
the upper mantle appears more transparent (Tittgemeyer et al., and magnetic data was merged with the 90 m elevation topo-
1999). Prominent reflections in the upper crust can be corre- graphic information available from NASA (Jarvis et al.,
lated with major sedimentary basins such as the Sinu–San Ja- 2008), with corrections from the International Gravity Stand-
cinto, Lower Magdalena, Middle Magdalena, and Eastern ardization Net 1971 (IGSN71), the World Geodetic System
Cordillera, as well as major faults such as the Romeral fault 1984 (WGS-84), the International Geomagnetic Reference
zone (RFZ). This major fault separates oceanic crustal rocks in Field (IGRF) and the Observed Magnetic Intensity (Hinze
the western terranes of Colombia and continental basement in et al., 2005; Maus et al., 2005). The final database allowed
eastern Colombia (Cediel et al., 2003). In general, seismicity us to estimate free air and magnetic anomalies. We then cal-
east of the Romeral is more concentrated in the older and culated 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) Geo-
logic 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 velocities match with slabs suggested by the gravity and
gravity and magnetic model responses (Talwani et al., magnetic models.
1959; Talwani and Heirtzler, 1964; Geosoft, 2010) and com-
paring with the observed data, we used values of density, mag-
Other Geophysical Information for Constraining
netization, and magnetic susceptibility shown on Table 3. In
the Interpretation
addition, the gravity and magnetic models were constrained
with the seismological and seismic data, as well as geologic Seismic attenuation has been used for examining the
transects compiled with superficial cartography and represen- acoustic contrast between the upper mantle and the litho-
tative seismic lines (Section AA0 , Figs. 1a and 5; and Section sphere because it is believed that this seismic factor is a
BB0 , Figs. 1a and 6; Lopez, 2004). physical parameter closely related to the thermal state of
Because of restrictions on the data use of the Trans- the volume sampled by the waves (Faul and Jackson, 2005;
Andean megaregional seismic-reflection line, it was not Priestley and McKenzie, 2006; Yang et al., 2007). Therefore,
possible to estimate refraction travel-time tomography for in order to evaluate the empirical superficial response of the
correlating with the profiles presented in this paper. However, lithospheric thermal field, we plotted geothermal anomalies
in order to interpret the thermal and tectonic structure in this reported from oil wells in Colombia (Vargas et al., 2009)
region, we used the velocity anomalies reported by Vargas onto the two sections (Figs. 5a and 6a). Furthermore, topo-
et al. (2007) and van der Hilst and Mann (1994) that show graphic response and focal mechanisms compiled from NEIC
similar distribution anomalies. In general, trends of high 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.
Depth–time 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 tions flank the intersection of the Panama arc-indenter and
of eight variables along the profiles: hypocenter solutions, the Caldas tear to the north (Section AA0 , Figs. 1a and 5)
focal mechanisms, coda-wave attenuation, gravity, magnetic and south (Section BB0 , Figs. 1a and 6). The profiles are ex-
and geothermal anomalies, and geologic and topographic tended to the west from the Caribbean Sea and Pacific Ocean
data derived from seismic-reflection profiles. The two sec- 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 Density (kg=m3 ) Susceptibility Magnetization (A=m)
−1
Caribbean and Nazca plates 3:30 × 10 3
1:26 × 10 –1:38 × 10 0
1:00 × 10−3 –1:45 × 100
Lower continental crust 2:78 × 103 1:26 × 10−5 1:26 × 10−5 –1:90 × 100
Upper continental crust 2:31 × 103 3:21 × 10−2 –4:90 × 10−2 5:51 × 10−1 –1:00 × 100
Mantle 3:15 × 103
Accreted sediments and oceanic crust 2:07 × 103 –2:35 × 103 1:26 × 10−1 1:00 × 10−3 –8:01 × 10−1
Oceanic sediments 1:68 × 103 –2:00 × 103 1:26 × 10−5 1:00 × 10−3 –4:50 × 10−1
Water 1:03 × 103
Tearing and Breaking Off of Subducted Slabs as the Result of Collision of the Panama Arc-Indenter 2039

(a)
-72.8
Longitude (°)

-73
y(x) = a x + b
a = -0.00529
b = 17.428
-73.2 R = 0.36776

-73.4

1994 1996 1998 2000 2002 2004 2006 2008 2010 2012 2014

(b) 7.4

7.2
Latitude (°)

y(x) = a x + b
a = -0.006479
7 b = -60.081
R = 0.32692
6.8

6.6
1994 1996 1998 2000 2002 2004 2006 2008 2010 2012 2014

(c) -80
Depth (km)

-100
-120
y(x) = a x + b
-140 a = 0.23732
-160
b = -621.43
R = 0.11894
-180
-200
-220
1994 1996 1998 2000 2002 2004 2006 2008 2010 2012 2014
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 least-
squares means.

the northern Andes, the Bucaramanga and Cauca seismic the sinking oceanic crust further increases its dip to penetrate
nests of intermediate depth earthquakes, and major faults the mantle.
including the Romeral, the Llanos, the Santa Marta– Intensive intermediate seismicity of the Bucaramanga
Bucaramanga, the Garrapatas, and the Ibague faults. nest is concentrated within the subducted Caribbean slab
at depths of 130–160 km (Figs. 1a and 5). Furthermore, there
are at least two clusters of events concentrated at 80 and
Results and Discussion 110 km. Focal mechanisms of the Bucaramanga nest for
events reported by NEIC (1977–2012) with mb ≥ 4:0 include
Northern Transect Crossing the Caribbean Plate mainly reverse and strike-slip events (Fig. 2). We relate this
and the Bucaramanga Nest
earthquake nest to resistance to subduction of the Panama
Our northern transect shows the presence of an oceanic indenter and consequent initiation of tearing and slab break
crust with thicknesses greater than 20 km related to the Car- off at intermediate depths of the down-dip part of the slab
ibbean oceanic plateau with a shallow subduction angle of (Cortés and Angelier, 2005). In order to show the temporal
< 15° over a distance of ∼100 km that extends from the trend of the nest using seismicity located by the CNSN be-
South Caribbean deformed belt to the Romeral fault zone tween 1993 and 2012, we have compiled data from the Buca-
(Figs. 1a and 5). At this point the slab dip steepens to 20°– ramanga nest (Zarifi, 2006; Fig. 8). Although errors related
30° into the mantle. Following our interpretation of the to hypocentral solutions and temporal variation in the array
Trans-Andean seismic-reflection line (Fig. 7), the interaction configuration may blur the dimensions of the nest (Schneider
between the continental crust and the South Caribbean De- et al., 1987; Frohlich et al., 1995; Zarifi, 2006), and also
formed Belt (Sinu–San Jacinto basin) suggests that the oce- nearby events could be associated erroneously, the continu-
anic crust is sinking with a higher dip and is reaching the ous recording of seismicity for a period as long as 19 years
intermediate to deep seismicity zone (∼160 km), where shows a progressive southward and westward displacement
2040 C. A. Vargas and P. Mann

(a) -75.5
Longitude(°)

-76 y(x) = a (x - b)
a = 0.021514
b = 5553.5
R = 0.40882
-76.5

-77
1994 1996 1998 2000 2002 2004 2006 2008 2010 2012 2014

(b) 5.5

5
Latitude(°)

y(x) = a x + b
4.5
a = 0.0085141
b = -12.525
R = 0.12753
4

3.5
1994 1996 1998 2000 2002 2004 2006 2008 2010 2012 2014

0
(c)
-50
Depth(km)

y(x) = a x + b
a = 1.0488
-100 b = -2211.1
R = 0.19102
-150

-200
1994 1996 1998 2000 2002 2004 2006 2008 2010 2012 2014
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 accompanied a proposed westward jump in subduction from
events is of low confidence, this unidirectional, westward the Romeral fault zone in the Central Cordillera to the
displacement of events may be caused by down-dip and present Colombia trench (Cediel et al., 2003). An additional
southwestward propagation of tearing of the subducted attenuation anomaly observed on tomographic data indicates
Caribbean slab. a prominent high thermal inflow from the mantle that is
located beneath the forebulge of the Llanos basin and is ap-
Southern Tomographic Transect Crossing the Nazca proximately coincident with the largest geothermal anomaly
Plate and the Cauca Nest measured in this basin.
Seismicity of the Cauca nest is highly dispersed in depth
The Nazca oceanic slab has been modeled with an (70–150 km) and its time evolution seems more complex
∼15–22 km deep crustal thickness and a constant dip angle than the Bucaramanga nest (Fig. 9). Even with the low con-
of 30°–40° to a depth > 150 km beneath the active volcanic fidence of the lineal regression, seismicity from the CNSC
line (Figs. 1a and 6). The volcanic line is underlain by catalog shows an eastward displacement of events. The Cauca
high-attenuation anomalies indicative of a normal melting nest exhibits earthquakes with focal mechanisms ranging from
range for the subducted oceanic slab (Figs. 2, 4c, 6). High- pure gravitational collapse to strike-slip in the north and
attenuation anomalies and the presence of shallow to inter- reverse with strike-slip component in the south (Fig. 2).
mediate seismicity around the Romeral fault zone suggest
this major strike-slip provides another major upward conduit
Nature of the Caldas Tear
for the release of upper mantle heat. A large low-attenuation
anomaly corresponds to the low geothermal gradient ob- In addition to the hypocenter solutions that initially
served between the Colombian trench along the Pacific mar- revealed the ∼240 km long offset of intermediate to deep
gin and the Central Cordillera. The low geothermal gradient seismicity of the Caldas tear, we found that differential
coincides with thick volcanic and sedimentary material ac- displacements between the southern of the Caldas tear in the
creted to western Colombia, mainly in the Western Cordillera South American plate and the Panama arc-indenter, based on
and the Baudo Range. The accretion of this area may have 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 2041

toward the east (Trenkamp et al., 2002; e.g., BOGO versus Magdalena and Cauca rivers of northwestern South America.
MZAL, RION, BUCM, MONT, and CART GPS stations). As- After flowing 200 km from their sources, these rivers occupy
suming the BOGO station as the reference point south of the broad river valleys. Downstream, rivers passing the Caldas
Caldas tear, and the CHEP station as the reference point on the tear lineament change their morphology from broad valleys
Panama indenter, we estimate ∼24 mm=year of active right- to steeper relief gorges near the surface projection of the
lateral displacement across the Caldas tear. The hypothesis Caldas tear. The Cauca River near Supia town (Fig. 10a,b)
of lateral homogeneity of the crust, constant displacement rate, reduces to a narrow channel only 150 m wide whereas the
and a seismic offset along Caldas tear of ∼240 km, would Magdalena River near the town of Honda reduces to a 250 m
suggest an ∼10:0 Ma initiation of the Panama arc–Colombia wide channel (Fig. 10a,c). In contrast to the broader, 42 km
collision (∼240 km=24 mm=year). Geologic field observa- wide valleys observed upstream (e.g., Bolivar and Guamo
tions in the Panama-arc (Coates et al., 2004) suggest that the towns in Fig. 10a,d), in both areas the steeper gradient and
age for the initiation of the Panama arc collision with northern more narrow rivers produce rapids that are an impediment to
South America occurred between 12.8 and 7.1 Ma, which is navigation. The narrowness of these rivers favored the down-
consistent with our estimated ∼10:0 Ma initiation of the tear stream economic development of urban settlements such as
propagation. In addition, because the origin of the east–west Honda and Supia from Spanish colonial times. However, the
Sandra ridge occurs between 9 and 12 Ma (Defant et al., 1992; quake that occurred on 16 June 1805 that destroyed Honda
Lonsdale, 2005), and this structure is collinear with the Caldas and other nearby towns shows us that the Caldas tear could
tear, we propose that the right-lateral lineament defined by the form a major source of earthquakes and seismic hazard in
Caldas tear and the Sandra ridge, constitutes a major area of this region.
lithospheric weakness along the southern flank of the Panama The Caldas tear may localize large crustal earthquakes,
arc-indenter. Although there is no evidence for recent activity including the recent strong-motion activity associated with
of the Sandra ridge due to lack of near-bottom instrumentation the Tauramena earthquake (19 January 1995, M w 6.5).
in the Pacific offshore of Colombia and Panama, recent earth- Two types of focal mechanisms have been proposed for this
quakes and the adakite magmatism along the Caldas tear may event, one of which suggests an east–west-oriented right
indicate that this lineament localizes upper crustal fault con- lateral movement (Dimate et al., 2003; Fig. 2). Similarly,
duits that allowed the upward migration of magmatic fluids seismic activity in the area of the Armenia earthquake (25
and are associated with elevated geothermal anomalies. Offset January 1999, M w 6.2) shows east–west alignment of after-
of intermediate- to deep-seismicity that defines the Caldas tear shocks with the Caldas tear. This long right Caldas tear does
is also coincident with inactive volcanoes of adakite compo- not explain the Quetame earthquake (24 May 2008, mL 5.7)
sition and geothermal anomalies (Fig. 2). The adakites of the with left-lateral slip, in which the Caldas tear could have
Ruiz volcanic complex with ages ranging from 0.97–0.05 and played an important role for controlling the propagation
1.8–0.6 Ma (Borrero et al., 2009) and the Paipa–Iza volcanic of north–south-trending faults (such as this event or any re-
complex dated 1.9–2.5 Ma (Pardo et al., 2005) are a likely lated to the Llanos fault system) in a similar way as has been
consequence of magmatism related to this progressive suggested in the southwest Colombia margin where trans-
slab tear. Low ratio 87 Sr=87 Sr (0.705) and the presence of verse faults reduce coupling between adjacent segments
xenoliths of metamorphic rocks in this last volcanic complex (Collot et al., 2004).
(J. M. Jaramillo, personal comm., 2012) support the proposed We removed the crustal earthquakes from the database
break-off interpretation south of the Bucaramanga nest and of events located by the CNSN between 1993 and 2012 in
east of the Caldas tear. order to illustrate the upper surface of the subducted slabs
Surficial evidence of this lithospheric tear are restricted beneath Colombia (Fig. 11a). The 3D model of this surface
to presence of mineral deposits, hydrocarbon occurrences, images the Caldas tear and flat-slab subduction geometry
and some geomorphological anomalies. High-grade mineral that we relate to the presence of the Panama arc-indenter (Ra-
deposits of platinum, gold, and copper exploited in the min- mos and Folguera, 2009). The northern border of the indenter
ing areas called Condoto (Tistl, 1994), Marmato (Ordoñez, becomes diffuse and does not appear to form a distinctive
2001), Quinchia (INGEOMINAS, 1999), La Colosa (Gil- tear as seen south of the indenter (see dashed line in Fig. 11a).
Rodriguez, 2010), and Cerro de Cobre (McLaughlin and The Caribbean plate changes from flat to steep subduction
Arce, 1970) are near, or collinear with, the Caldas tear and toward the northeast. South of the Caldas tear (∼2:5° N),
exhibit ages ranging between 6 and 20 Ma (see blue hexa- there is a shift to a new pattern of intermediate and deep
gons on map in Fig. 10a). In addition, significant changes in seismicity associated with flat subduction along with the
distribution and trend of oil and gas seeps, as well as the development of a broad area of active volcanoes in southern
hydrocarbon fields on both sides of the Caldas tear, suggest Colombia and northern Ecuador. Figure 11b presents a con-
that this structure may also affect the geometry of several ceptual model that explains the kinematic role of the
sedimentary basins (e.g., Llanos foreland, Eastern Cordillera, Panama arc-indenter whose southern boundary is defined
Middle Magdalena Valley). But likely the most prominent by the Sandra ridge and the Caldas tear. In this model the
geomorphological evidence is coming from hydraulic behav- coupling of the Panama arc with the Caribbean plate could
ior of the main rivers that cross the south-to-north-flowing 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 sedi-
mentary 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 condi- branches of the Romeral fault zone south of the Caldas tear
tions that favor flat subduction. In the east, the Caribbean show right-lateral offset from parallel faults including the
plate suddenly changes its subduction angle and produces Palestina, Cimitarra, Mulato–Getudo, Honda, or Bituima
a break off of the slab around the location of the Bucara- faults to the north. Other faults with southwest–northeast
manga nest. South of the Caldas tear, the Nazca plate is sub- trend, such as the Ibague and Garrapatas, could be part of
ducting beneath the South American plate with a steeper a transverse strike-slip fault at the surface level overlying
angle and a faster rate. The Cauca nest is a combined product the Caldas tear (Fig. 1a). These results raise new questions
of eastward decoupling of plates along the Caldas tear and about the regional evolution in northwestern South America.
flexure during the subduction process. For example, the Great Arc of the Caribbean has been de-
A corollary of our model for the Panama indenter and fined along the south Caribbean region but disappears once
the formation of the Caldas tear is the eastward indentation of it enters the Guajira basin and the Santa Marta Massif north
geologic features. Inspection of the Map of Quaternary faults of Colombia (Ostios et al., 2005). As a consequence of the
and folds of Colombia (Paris et al., 2000) suggests that some 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 south-
west. 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 Murindó 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 widens the northern Andes in Colombia and Venezuela
indenter. (Fig. 1b). The Panama collision initiated ∼10 Ma and con-
tinues to be active as shown by GPS data.
• The southern edge of the Panama indenter is associated
Conclusions with the proposed Caldas slab tear at latitude ∼5:6° N. This
• The eastward-directed collision of the buoyant Panama tear extends for ∼240 km in an east–west direction and
arc-indenter with northwestern South America produces is collinear with the ∼9–12 Ma, now extinct, east–west-
a distinctive V-shaped pattern of crustal deformation and oriented 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 1233-487-25728, Contract 589-2009 (COLCIENCIAS); CGL2005-04541-
the Caldas tear may have formed as a zone of lithospheric C03-02 and CGL2008-00869/BTE (UPC, MICCIN, FEDER). We also thank
the Associate Editor Heather DeShon, and two anonymous reviewers for their
weakness along the now subducted part of the inactive helpful reviews of this paper.
Sandra spreading ridge.
• The ∼240 km long Caldas tear is a narrow, east–west- References
trending boundary between two subducted slabs of differ-
ent dip. The northern zone is the down-dip extension of the Adamek, S., C. Frohlich, and W. Pennington (1988). Seismicity of the
Panama arc, has a shallower dip, and is not associated with Caribbean-Nazca boundary: Constraints on microplate tectonics of
the Panama region, J. Geophys. Res. 93, 2053–2075.
active arc volcanism. The southern zone has a steeper dip Aki, K. (1969). Analysis of the seismic coda of local earthquakes as
and is associated with an active volcanic front (Fig. 11a,b). scattered waves, J. Geophys. Res. 74, 615–631.
• The Caldas tear also localizes angular difference of the Aki, K., and B. Chouet (1975). Origin of coda waves: Source, attenuation
subduction geometry in both geophysical sections pre- and scattering effects, J. Geophys. Res. 80, 615–631.
Badi, G., E. Del Pezzo, J. M. Ibanez, F. Bianco, N. Sabbione, and M. Araujo
sented in this work (Figs. 5 and 6). The lineament defined
(2009). Depth dependent seismic scattering attenuation in the Nuevo
by the ∼240 km long offset of the deep seismicity along Cuyo region (southern central Andes), Geophys. Res. Lett. 36,
∼5:6° N; the eruption of the north–south belt of active no. L24307, doi: 10.1029/2009GL041081.
volcanism and presence of extinct magmatic bodies with Borrero, C., L. M. Toro, M. Alvaran, and H. Castillo (2009). Geochemistry
adakite composition along the lineament; the occurrence and tectonic controls of the effusive activity related with the
of high-grade mineral deposits and geothermal gradient ancestral Nevado del Ruiz volcano, Colombia, Geofísc. Int. 48,
no. 1, 149–169.
anomalies; the different patterns associated with oil and Calais, E., and P. Mann (2009). A combined GPS velocity field for the
gas manifestations; the distribution of major oil and gas Caribbean plate and its margins (abstract #G33B-0657), American
deposits north and south of the Caldas tear as well as the Geophysical Union, (Fall Meet.), G33B-0657.
GPS measurements and strong-motion events with right- Cediel, F., R. P. Shaw, and C. Caceres (2003). Tectonic assembly of the
lateral movements, support the existence of the Caldas tear. northern Andean block, in The circum-Gulf of Mexico and the
Caribbean: Hydrocarbon habitats, basin formation, and plate tectonics,
AAPG Memoir 79, 1–34.
Data and Resources 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 colli-
Waveforms and preliminary hypocentral solutions of sion of the Panama arc with northwestern South America, Bull. Geol.
Soc. Am. 116, nos. 11–12, 1327–1344.
the Colombian territory were supplied by the Geological Sur-
Collot, J. Y., B. Marcaillou, F. Sage, F. Michaud, W. Agudelo, P. Charvis, D.
vey of Colombia (INGEOMINAS). Bouguer and magnetic Graindorge, M. A. Gutscher, and G. D. Spence (2004). Are rupture
data provided for the National Hydrocarbons Agency of zone limits of great subduction earthquakes controlled by upper plate
Colombia (http://www.anh.gov.co/es/index.php?id=82, last structures? Evidence from multichannel seismic reflection data ac-
accessed November 2012) were used for modeling geologic quired across the northern Ecuador–southwest Colombia margin, J.
sections using the GM-SYS profile module of the Oasis Geophys. Res. 109, B11103, doi: 10.1029/2004JB003060.
Corredor, F. (2003). Seismic strain rates and distributed continental defor-
Montaj software (Geosoft, 2010). This software calculates mation in the northern Andes and three-dimensional seismotectonics
the gravity and magnetic model response based on the meth- of northwestern South America, Tectonophysics 372, 147–166.
ods of Talwani et al. (1959), and Talwani and Heirtzler Cortés, M., and J. Angelier (2005). Current states of stress in the northern
(1964). GM-SYS uses a 2D, flat-earth model for the gravity Andes as indicated by focal mechanisms of earthquakes, Tectonophy-
and magnetic calculations. Each structural unit or block sics 403, 29–58.
Defant, M. J., T. E. Jackson, M. S. Drummond, J. Z. De Boer, H. Bellon, M.
extends to plus and minus infinity in the direction perpen- D. Feigenson, R. C. Maury, and R. H. Stewart (1992). The geochem-
dicular to the profile. The earth is assumed to have topogra- istry of young volcanism throughout western Panama and southeastern
phy but no curvature. The model also extends plus and minus Costa Rica: An overview, J. Geol. Soc. 149, 569–579.
30,000 km along the profile to eliminate edge effects. The 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
90 m elevation topographic information used for the gravity
Tauramena (Colombia) earthquake: Geometry and stress regime,
modeling is available from the CGIAR-CSI SRTM 90 m Tectonophysics 363, 159–180.
database (http://srtm.csi.cgiar.org, last accessed November Faul, U. H., and I. Jackson (2005). The seismological signature of temper-
2012). Focal mechanisms reported by NEIC were used in ature and grain size variations in the upper mantle, Earth Planet. Sci.
this work (http://earthquake.usgs.gov/earthquakes/eqarchives/ Lett. 234, 119–134.
sopar/, last accessed November 2012). 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
Acknowledgments the Bucaramanga, Colombia, earthquake nest, Bull. Seismol. Soc. Am.
85, 1622–1634.
This work was partially funded by the industry sponsors of the CBTH Geosoft (2010). GM-SYS profile module, Oasis montaj processing and
project of the University of Houston and by fellowship support from the mapping.
University of Texas at Austin. Earthquake, gravity, magnetic, seismic, Gil-Rodriguez, J. (2010). Igneous Petrology of the Colosa Gold-Rich Porphyry
and geothermal data were kindly provided by Agencia Nacional de Hidrocar- System (Tolima, Colombia), M. Sc. Thesis, University of Arizona, 51 pp.
buros, Universidad Nacional de Colombia, INGEOMINAS, and the following Haskov, J., S. Malone, D. McClury, and R. Crosson (1989). Coda-Q for the
research projects: 1233-333-18664, Contract 201-2006 (COLCIENCIAS); State of Washington, Bull. Seismol. Soc. Am. 79, 1024–1038.
Tearing and Breaking Off of Subducted Slabs as the Result of Collision of the Panama Arc-Indenter 2045

Hinze, W. J., C. Aiken, J. Brozena, B. Coakley, D. Dater, G. Flanagan, Ostios, M., F. Yoris, and H. G. Avé Lallemant (2005). Overview of the south-
R. Forsberg, T. Hildenbrand, G. R. Keller, J. Kellogg, R. Kucks, east Caribbean-South American plate boundary zone, Spec. Pap. Geol.
X. Li, A. Mainville, R. Morin, M. Pilkington, D. Plouff, D. Ravat, Soc. Am. 394, 53–89.
D. Roman, J. Urrutia-Fucugauchi, M. Veronneau, M. Webring, and Pardo, N., H. Cepeda, and J. M. Jaramillo (2005). The Paipa volcano,
D. Winester (2005). New standards for reducing gravity data: The Eastern Cordillera of Colombia, South America: Volcanic stratigraphy,
North American gravity database, Geophysics 70, no. 4, J25–J32. Earth Sci. Res. J. 9, no. 1, 3–18.
Hoerl, A. E., and R. W. Kennard (1970). Ridge regression: Biased estimation Paris, G., R. L. Dart, and M. N. Manchette (2000). Map of Quaternary
for nonorthogonal problems, Technometrics 12, 55–67. faults and folds of Colombia and its offshore regions, U.S. Geol. Surv.
Hoerl, A. E., R. W. Kennard, and K. F. Baldwin (1975). Ridge regression: Open-File Rept., Scale 1:2,500,000.
Some simulation, Comm. Stat. 4, 105–123. Priestley, K., and D. McKenzie (2006). The thermal structure of the
INGEOMINAS (1999). Risaralda, Department Mining Inventory, Internal lithosphere from shear wave velocities, Earth Planet. Sci. Lett. 244,
report (in Spanish). 285–301.
Jarvis, A., H. I. Reuter, A. Nelson, and E. Guevara (2008). Hole-filled SRTM Pulli, J. J. (1984). Attenuation of coda waves in New England, Bull. Seismol.
for the globe version 4, available from the CGIAR-CSI SRTM 90m Soc. Am. 74, 1149–1166.
Database (http://srtm.csi.cgiar.org; last accessed July 2012). Ramos, V. A., and A. Folguera (2009). Andean flat-slab subduction through
Karato, S., and H. Jung (1998). Water, partial melting and the origin of the time, Spec. Publ. Geol. Soc. Lond. 327, 31–54.
seismic low velocity and high attenuation zone in the upper mantle, Sandwell, D. T., and W. H. F. Smith (2009). Global marine gravity
Earth Planet. Sci. Lett. 157, 193–207. from retracked Geosat and ERS-1 altimetry: Ridge segmentation
Lacruz, J., A. Ugalde, C. A. Vargas, and E. Carcolé (2009). Coda-wave versus spreading rate, J. Geophys. Res. 114, B01411, doi: 10.1029/
attenuation imaging of Galeras Volcano, Colombia, Bull. Seismol. 2008JB006008.
Soc. Am. 99, no. 6, 3510–3515. Schneider, J. F., W. D. Pennington, and R. P. Meyer (1987). Microseismicity
Levenberg, K. (1944). A method for the solution of certain nonlinear and focal mechanisms of the intermediate-depth Bucaramanga
problems in least squares, Q. Appl. Math. 2, 164–168. nest, Colombia, J. Geophys. Res. 92, no. B13, 13,913–13,926, doi:
Lonsdale, P. (2005). Creation of the Cocos and Nazca plates by fission of the 10.1029/JB092iB13p13913.
Farallon plate, Tectonophysics 404, 237–264. Singh, S., and R. B. Herrmann (1983). Regionalization of crustal coda Q in
Lopez, C. (2004). Upper crust models of Colombia, INGEOMINAS, Bogota, the continental United States, J. Geophys. Res. 88, 527–538.
internal Rept., 64 pp. Taboada, A., L. A. Rivera, A. Fuenzalida, A. Cisternas, H. Philip,
Malin, P. E. (1978). A first order scattering solution for modeling lunar and H. Bijwaard, J. Olaya, and C. Rivera (2000). Geodynamics of the
terretrial seismic coda, Ph.D. Dissertation, Princeton University, Northern Andes; subductions and intracontinental deformation
Princeton, New Jersey. (Colombia), Tectonics 19, 787–813.
Marquardt, D. W. (1963). An algorithm for least squares estimation of Talwani, M., and J. R. Heirtzler (1964). Computation of magnetic anomalies
nonlinear parameters, J. Soc. Ind. Appl. Math. 11, 55–67. caused by two-dimensional bodies of arbitrary shape, in Computers in
Maus, S., S. Macmillan, T. Chernova, S. Choi, D. Dater, V. Golovkov, V. Lesur, the mineral industries, Part 1: Stanford Univ. Publ., Geological
F. Lowes, H. Luhr, W. Mai, S. McLean, N. Olsen, M. Rother, T. Sabaka, Sciences, G. A. Parks (Editor), Vol. 9, 464–480.
A. Thomson, and T. Zvereva (2005). The 10th generation international Talwani, M., J. L. Worzel, and M. Landisman (1959). Rapid gravity
geomagnetic reference field, Phys. Earth Planet. Int. 151, 320–322. computations for two-dimensional bodies with application to the
Maus, S., T. Sazonova, K. Hemant, J. D. Fairhead, and D. Ravat (2007). Mendocino submarine fracture zone, J. Geophys. Res. 64, 49–59.
National Geophysical Data Center candidate for the World Digital Tistl, M. (1994). Geochemistry of platinum-group elements of the zoned
Magnetic Anomaly Map, Geochem. Geophys. Geosyst. 8, Q06017, ultramafic Alto Condoto complex, northwest Colombia, Econ. Geol.
doi: 10.1029/2007GC001643. 89, no. 1, 158–167.
McLaughlin, D. H., and M. Arce (1970). Economic geology of the Zipaquira Tittgemeyer, M., F. Wenzel, T. Ryberg, and K. Fuchs (1999). Scales of
quadrangle and adjoining area, Department of Cundinamarca, heterogeneities in the continental crust and upper mantle, Pure Appl.
Colombia, U.S. Geol. Surv. Open-File Rept., 141 pp. Geophys. 156, 29–52.
Mitchell, B. J., and L. Cong (1998). Lg coda Q and its relation to the Trenkamp, R., J. N. Kellogg, J. T. Freymueller, and H. Mora (2002). Wide
structure and evolution of continents: A global perspective, Pure Appl. plate margin deformation, southern Central America and northwestern
Geophys. 153, 655–663. South America, CASA GPS observations, J. S. Am. Earth Sci. 15,
Mitchell, B. J., Y. Pan, J. Xie, and L. Cong (1997). Lg coda variation across no. 2, 157–171.
Eurasia and its relation to crustal evolution, J. Geophys. Res. 102, van der Hilst, R. D., and P. Mann (1994). Tectonic implications of tomo-
22,767–22,779. graphic images of subducted lithosphere beneath northwestern South
Mukhopadhyay, S., and J. Sharma (2010). Attenuation characteristics of America, Geology 22, 451–454.
Garwhal–Kumaun Himalayas from analysis of coda of local earth- Vargas, C. A., C. Alfaro, L. A. Briceño, I. Alvarado, and W. Quintero (2009).
quakes, J. Seismol. 14, doi: 10.1007/s10950-010-9192-9. Mapa Geotérmico de Colombia, 2009, in Proceedings of X Simposio
National Hydrocarbons Agency of Colombia (2010). Total Bouguer Anoma- Bolivariano Exploración Petrolera en Cuencas Subandinas, Colombia
lies Map (MABT) of Colombia, National Hydrocarbons Agency of (in Spanish).
Colombia (ANH). Vargas, C. A., P. Mann, C. Gomez, L. A. Briceño, and C. Rey (2010). Trans-
O’Doherty, K. B., C. J. Bean, and J. McCloskey (1997). Coda wave imaging Andean mega-regional seismic reflection line extending from the
of the Long Valley caldera using a spatial stacking technique, Geophys. Caribbean coast to Cordillera Oriental of Colombia: Implications for
Res. Lett. 24, 1545–1550. hydrocarbon exploration, AAPG Annual Convention: Unmasking the
Ojeda, A., and J. Havskov (2001). Crustal structure and local seismicity in Potential of Exploration & Production, New Orleans, Abstract 90104.
Colombia, J. Seismol. 5, no. 4, 575–593. Vargas, C. A., L. G. Pujades, and L. A. Montes (2007). Seismic structure of
Oliver, M. A., and R. Webster (1990). Kriging: A method of interpolation south-central Andes of Colombia by tomographic inversion, Geofísc.
for geographical information system, Int. J. Geogr. Inf. Syst. 4, Int. 46, no. 2, 117–127.
no. 3, 313–332. Vargas, C. A., A. Ugalde, L. G. Pujades, and J. A. Canas (2004). Spatial
Ordoñez, C. O. (2001). Caracterização isotópica Rb-Sr e Sm-Nd dos variation of coda wave attenuation in northwestern Colombia,
principais eventos magmáticos nos Andes Colombianos, Ph.D. Thesis, Geophys. J. Int. 158, 609–624.
Instituto de Geociências, Universidade de Brasilia, Brasilia, 197 pp. (in Xie, J. (2002). Lg Q in the eastern Tibetan Plateau, Bull. Seismol. Soc. Am.
Portuguese). 92, 871–876.
2046 C. A. Vargas and P. Mann

Yang, Y., D. W. Forsyth, and D. S. Weeraratne (2007). Seismic attenuation Department of Earth and Atmospheric Sciences
near the East Pacific Rise and the origin of the low-velocity zone, Earth University of Houston
Planet. Sci. Lett. 258, 260–268, doi: 10.1016/j.epsl.2007.03.040. 4800 Calhoun Boulevard
Zarifi, Z. (2006). Unusual subduction zones: Case studies in Colombia and Houston, Texas 77004
pmann@uh.edu
Iran, Ph.D. Thesis, University of Bergen, Norway, 78 pp.
(P.M.)

Department of Geosciences
Universidad Nacional de Colombia–Sede Bogotá
Carrera 45 No 26-85–Edificio Manuel Ancizar
Bogotá D.C.–Colombia, 111321
cavargasj@unal.edu.co
(C.A.V.) Manuscript received 3 November 2012

You might also like