Sie sind auf Seite 1von 24

Tectonophysics 460 (2008) 134–157

Contents lists available at ScienceDirect

Tectonophysics
j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / t e c t o

Structural imprints at the front of the Chocó-Panamá indenter: Field data from the
North Cauca Valley Basin, Central Colombia
F. Suter ⁎, M. Sartori, R. Neuwerth, G. Gorin
Department of Geology–Paleontology, University of Geneva, 13 rue des Maraîchers, 1205 Geneva, Switzerland

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

Article history: The northern Andes are a complex area where tectonics is dominated by the interaction between three major
Received 4 January 2008 plates and accessory blocks, in particular, the Chocó-Panamá and Northern Andes Blocks. The studied Cauca
Received in revised form 16 July 2008 Valley Basin is located at the front of the Chocó-Panamá Indenter, where the major Romeral Fault System,
Accepted 18 July 2008
active since the Cretaceous, changes its kinematics from right-lateral in the south to left-lateral in the north.
Available online 26 July 2008
Structural studies were performed at various scales: DEM observations in the Central Cordillera between 4
and 5.7°N, aerial photograph analyses, and field work in the folded Oligo-Miocene rocks of the Serranía de
Keywords:
Distributed shear strain
Santa Barbara and in the flat-lying, Pleistocene Quindío-Risaralda volcaniclastic sediments interfingering
Intramontane basin with the lacustrine to fluviatile sediments of the Zarzal Formation.
Lateral spreading The data acquired allowed the detection of structures with a similar orientation at every scale and in all
Paleostress inversion lithologies. These families of structures are arranged similarly to Riedel shears in a right-lateral shear zone
Quindío-Risaralda volcaniclastic Fan and are superimposed on the Cretaceous Romeral suture.
Zarzal Formation They appear in the Central Cordillera north of 4.5°N, and define a broad zone where 060-oriented right-
lateral distributed shear strain affects the continental crust. The Romeral Fault System stays active and strain
partitioning occurs among both systems. The southern limit of the distributed shear strain affecting the
Central Cordillera corresponds to the E–W trending Garrapatas–Ibagué shear zone, constituted by several
right-stepping, en-échelon, right-lateral, active faults and some lineaments. North of this shear zone, the
Romeral Fault System strike changes from NNE to N.
Paleostress calculations gave a WNW–ESE trending, maximum horizontal stress, and 69% of compressive
tensors. The orientation of σ1 is consistent with the orientation of the right-lateral distributed shear strain
and the compressive state characterizing the Romeral Fault System in the area: it bisects the synthetic and
antithetic Riedels and is (sub)-perpendicular to the active Romeral Fault System.
It is proposed that the continued movement of the Chocó–Panamá Indenter may be responsible for the 060-
oriented right-lateral distributed shear strain, and may have closed the northern part of the Cauca Valley,
thereby forming the Cauca Valley Basin.
Conjugate extensional faults observed at surface in the flat-lying sediments of the Zarzal Formation and
Quindío-Risaralda volcaniclastic Fan are associatedwith soft-sediment deformations. These faults are
attributed to lateral spreading of the superficial layers during earthquakes and testify to the continuous
tectonic activity from Pleistocene to Present.
Finally, results presented here bring newinformation about the understanding of the seismic hazard in this
area: whereas the Romeral Fault Systemwas so far thought to be themost likely source of earthquakes,
themore recent cross-cutting fault systems described herein are another potential hazard to be considered.
© 2008 Elsevier B.V. All rights reserved.

1. Introduction Virginia (Fig. 1). Various hypotheses have been brought forward to
explain its origin: some authors have interpreted it as a graben
The inter Andean Cauca Valley Basin in Central Colombia corre- (Acosta, 1978; McCourt, 1984; Droux and Delaloye, 1996; MacDonald
sponds to a more than 200 km long alluvial plain between Cali and La et al., 1996) or a pull-apart basin (Kellogg et al., 1983; Alfonso et al.,
1994), whereas recent studies have shown a neotectonic compressive
activity in some faults bounding this floodplain (López and Moreno,
⁎ Corresponding author. Present address: Area de Ciencias del Mar, Universidad 2005; López et al., 2005).
EAFIT, A.A. 3300, Medellín, Colombia. Fax: +41 22 379 32 10.
E-mail addresses: Fiore.Suter@terre.unige.ch, fsuter@eafit.edu.co (F. Suter),
This research aims at explaining the mechanisms which led to the
Mario.Sartori@terre.unige.ch (M. Sartori), Ralph.Neuwerth@terre.unige.ch northern closure of this sedimentary basin. Syntectonic deposits have
(R. Neuwerth), Georges.Gorin@terre.unige.ch (G. Gorin). been used to reconstruct the kinematic evolution of this area since the

0040-1951/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.tecto.2008.07.015
F. Suter et al. / Tectonophysics 460 (2008) 134–157 135

2. Geodynamical setting

The northern part of the Colombian Andes displays a complex


structural pattern, resulting from the interaction of three major
converging tectonic plates (Fig. 2). With respect to the South American
Plate, the Caribbean Plate moves east to southeast at a velocity between
10 and 22 mm/y, whereas the Nazca Plate moves eastwards at a velocity
of 50 to 78 mm/y (Pennington, 1981; Freymueller et al., 1993; Ego et al.,
1996; Mountney and Westbrook, 1997; Gutscher et al., 1999; Taboada
et al., 2000; Trenkamp et al., 2002; White et al., 2003). Based on shallow
to deep seismicity and seismic tomographic images, various 3-D models
of the lithospheric structure in the Northern Andes have been produced
(Pennington, 1981; Van der Hilst and Mann, 1994; Gutscher et al., 1999;
Taboada et al., 2000; Cortes and Angelier, 2005). Although the geometry
of subducted slabs is still controversial in northwestern Colombia, these
authors agree that both the Caribbean and Nazca slabs are subducting
the South American Plate, the former with a low angle in an ESE to SE
direction, and the latter with a high angle in an ESE direction.
Somewhere north of 5°N, these two subducting plates overlap. Taboada
et al. (2000) define the southern limit of the subducting Carribean slab at
5.2°N, in an area where the intermediate seismicity distribution beneath
the Eastern Cordillera suggests a right-lateral E–W trending transform
shear zone (TSZ), whereas Pindell et al. (2005) locate the present
position of the Caribbean Plate southern edge at 4°N.
In the convergence zone between these three major plates, three
distinct blocks are moving and being deformed in order to
accommodate the resulting stress; i.e., the Chocó-Panamá, North
Andes, and Maracaibo blocks (Fig. 2).
The Chocó-Panamá Block (CPB) is a volcanic island arc with its
associated oceanic crust (Restrepo and Toussaint, 1988). It collides into
NW South America in an E to ESE direction, and is limited by the
transpressive, left-lateral Uramita fault zone to the east and the right-
lateral Istmina fault zone to the south (Duque-Caro, 1990; Taboada et al.,
2000). The latter lays slightly west of, and is parallel to the Garrapatas
fault, which displays neotectonic activity (Taboada et al., 2000). The
onset of the collision is not precisely dated, but it ranges from the Early
Miocene to Early Pliocene (Pennington, 1981; Restrepo and Toussaint,
1988; Duque-Caro, 1990; Mann and Corrigan, 1990; Van der Hilst and
Mann, 1994; Taboada et al., 2000; Trenkamp et al., 2002). Because of its
buoyancy the CPB does not subduct below South America. Therefore, it is
considered as a rigid indenter, which induces deformations north of 5°N
reaching the lowlands of the Eastern Cordillera some 600 km east. This
corresponds to a horizontal shortening exceeding 150 km (Audemard,
2002; Trenkamp et al., 2002). In addition, this collision is considered to
be responsible for the latest and major phase of uplift in the Colombian
Andes which corresponds to the Andean tectonic phase that affected the
three cordilleras (Taboada et al., 2000; Cortes et al., 2005).
The North Andes Block (or the Northern Andean Block, Cline et al.,
1981a) corresponds to the highly deformed portion of territory laying
between the three major tectonic plates and the CPB (Fig. 2). South of
4°N, it is limited westwards by the trench where the Nazca plate
subducts beneath the South American Plate, whereas north of 4°N it is
bounded westwards by the southern and eastern limits of the CPB. Its
eastern limit corresponds to the Santa Marta-Bucaramanga Fault (SMBF)
and the Eastern Frontal Fault System (EFFS), which borders the eastern
foothills of the Eastern Cordillera. South of 3.5°N the eastern boundary of
Fig. 1. Digital elevation model (DEM, USGS, 2005) of western central Colombia showing the North Andean Block changes strike from SSW to SW along the
the course of the Cauca River. The study area is located upstream of La Virginia town, at Algeciras transpressive right-lateral fault system (AFS). The latter is
the northern termination of the Cauca Valley Basin. Downstream of La Virginia, the located slightly west of the EFFS and continues SW down to the Gulf of
Cauca River regime changes from meandering and low energy to turbulent and erosive, Guayaquil in Ecuador (Velandia et al., 2005). To the south this block has a
and is confined to a canyon.
triangular shape and is squeezed between the Nazca and South
onset of the Chocó-Panamá Block (CPB) collision into the Western American plates. Studies based on shallow to deep earthquake focal
Cordillera (Fig. 2). mechanisms and/or GPS data demonstrated that this portion of the
The results of this study should contribute to a better under- Colombian Andes is moving towards the NNE (Pennington, 1981;
standing of the seismic hazard and improve the assessment of seismic Freymueller et al., 1993; Trenkamp et al., 2002). Trenkamp et al. (2002)
risks in this tectonically active area. estimated the velocity of motion at 6 ± 2 mm/y, Kellogg (1985; in:
136 F. Suter et al. / Tectonophysics 460 (2008) 134–157

Fig. 2. Geodynamics of NW South America: velocities and senses of motion for the different plates and blocks with respect to South America (Pennington, 1981; Kellogg et al., 1983;
Freymueller et al., 1993; Trenkamp et al., 2002); tectonic data modified after Gutscher et al. (1999), Taboada et al. (2000), Cortes and Angelier (2005) DEM from USGS, (2005).

Freymueller et al., 1993) at 10 mm/y with a direction of 55°, and undergoing right-lateral transpression induced by the slightly oblique
Freymueller et al. (1993) at 16 ± 5 mm/y with a direction of 35° ± 25. The Nazca Plate subduction with respect to the North Andes Block. The area
latter implies a transpressive right-lateral kinematics of the EFFS and studied in this research is located between 4 and 5°N, at the front of the
AFS. However, recent studies integrating field and subsurface structural Chocó-Panamá Block, east of its southeastern boundary (Figs. 1 and 3).
data and/or paleostress determinations (Branquet et al., 2002; Cortes
et al., 2005) suggest that, since the onset of the Andean tectonic phase, 3. Geologic setting
the part of the North Andes Block north of 4–5°N is undergoing
shortening in a direction perpendicular to the main fault trends rather 3.1. Geology
than right-lateral transpression. Therefore, although the main faults had
a transpressive right-lateral kinematics before Mio-Pliocene times, the North of La Virginia (Fig. 3), the Cauca Depression terminates
latter was converted into thrusting following the onset of the abruptly. South of this town, the latter is bounded to the east by the
convergence of North and South Americas and the subsequent Serranía de Santa Barbara (SSB), which forms a topographic barrier
indentation of the Chocó-Panamá Block (Cortes et al., 2005). According between this depression and the Quindío-Risaralda Basin to the east
to the latter authors, the North Andes Block is submitted to the influence (James, 1986; Suter et al., 2008) (Figs. 1, 3, 4 and 5).
of the Caribbean Plate rather than to that of the Nazca plate, at least The SSB is constituted by three continental sedimentary forma-
north of 4°N. This is in agreement with the southern limit of the tions: the Oligocene Cartago Formation (Rios and Aranzazu, 1989), the
subduced Caribbean slab at 5.2°N as suggested by Taboada et al. (2000). Miocene, syntectonic La Paila Formation, and La Pobreza Formation
In summary, the Northern Andes are divided into two distinct (Figs. 3 and 4), which is considered by some authors to be equivalent
structural domains. The first one, north of 4–5°N, is undergoing the in age and lithological composition to the La Paila Formation (Van der
strain generated by the subducted Caribbean plate and the indentation Hammen, 1958; McCourt, 1984; Keith et al., 1988; Rios and Aranzazu,
of the Chocó-Panamá Block. The second domain, south of 4–5°N, is 1989). These formations were deposited on the emerged part of the
F. Suter et al. / Tectonophysics 460 (2008) 134–157 137

Fig. 3. Compilation of the five 1:100'000 geological maps of INGEOMINAS (Caballero and Zapata, 1983; Parra, 1983; McCourt et al., 1984; Nivia et al., 1995; Estrada and Viana, 1998)
which cover the study area.

accretionnary prism very close to its contact with the South American Quindío-Risaralda volcaniclastic Fan (Fig. 3; Guarín et al., 2004);
Shield, west of the Romeral Fault System (RFS, materialized by the detrital material from the Western Cordillera and SSB; intercalated,
Cauca-Almaguer Fault in Fig. 4). autochthonous lacustrine diatomite-rich sediments of the Zarzal
The folding of the SSB began during Late Oligocene–Early Miocene Formation (Van der Hammen, 1958; Cardona and Ortiz, 1994;
times. It is documented by the syn-tectonic deposition of the La Paila Neuwerth et al., 2006; Suter et al., 2008). These interdigitated
Formation (Keith et al., 1988). It coincides with the change of sediments unconformably overlie all older units mentioned above.
convergence direction between the oceanic plate and the continental Van der Hammen (1958) suggested a Pliocene age for the Zarzal
margin in a W–E direction following the break up of the Farallon plate Formation without any palynological evidence. However, palynologi-
into the Nazca and Cocos plates (Taboada et al., 2000). The SSB is a fold cal investigations carried out in this study (Neuwerth et al., 2006;
and thrust mountain range with a NNE–SSW trending strike. The main Suter et al., 2008) have shown the presence of Alnus pollengrains in
segment shows cylindrical folds. Its central axis is characterized by a clays of the Zarzal Formation. This tree migrated southwards from
basement-involved pop-up structure (Suter, 2003). North America after the formation of the Panamá isthmus and its first
Since Pleistocene times, the two sub-basins separated by the SSB occurrence in Colombia dates back to less than 1 my (Hooghiemstra
were infilled by sediments with different sources: volcaniclastic, and Cleef, 1995). Consequently, a large part of the Zarzal Formation is
reworked material from the Central Cordillera and forming the considered Pleistocene in age.
138 F. Suter et al. / Tectonophysics 460 (2008) 134–157

Fig. 4. Simplified cross-section across the Cauca depression and the Quindío-Risaralda Basin, no vertical exaggeration, see Fig. 3 for location. After Alfonso et al. (1994), INGEOMINAS,
(1999), Neuwerth et al. (2006), Suter et al. (2008).

The deposits of the Zarzal Formation and Quindío-Risaralda Ansermanuevo faults are the most prominent structures belonging to
volcaniclastic Fan show numerous soft-sediment deformations this system (Fig. 6). In this zone they are left-lateral (James, 1986; Nivia
which testify of the continuous seismic activity of the study ar- et al., 1995; Guzmán et al., 1998; Paris et al., 2000).
ea during the Pleistocene (Neuwerth et al., 2006). These depo- On the other side of the Cauca Valley, in the SSB foothills and its
sits are essentially flat-lying, but can be locally slightly folded: the surroundings, the Quebradanueva faults system corresponds to west
Zarzal Formation near Obando (Cardona and Ortiz, 1994) and the and east verging thrusts where the Oligocene Cartago Formation
Quindío-Risaralda volcaniclastic Fan sediments in their western part overlies the Miocene La Paila Formation (Fig. 4). These faults were
(G. Paris, pers. comm.). Moreover, numerous extensional fractures defined as transpressive left-lateral by Guzmán et al. (1998). They
were encountered in these flat-lying, but locally deeply eroded de- bound the SSB pop-up to the east and west (Figs. 4 and 6).
posits (Pardo et al., 1994; Suter et al., 2008) (Fig. 3). The recent Further east, in the western foothills of the Central Cordillera, the
alluvial sediments of the Cauca River are now infilling the Cauca Cauca-Almaguer Fault System crosses the Quindío-Risaralda Fan and
Valley. outlines the suture between oceanic rocks to the west and continental
basement to the east (Figs. 3 and 4) (Cline et al., 1981b; McCourt and
3.2. Tectonics Aspden, 1983; McCourt, 1984; Restrepo and Toussaint, 1988; Restrepo-
Pace, 1992; Paris and Romero, 1994; MacDonald et al., 1996; Nivia,
The fault pattern in the study area and its surroundings is 1996; Cortes et al., 2005; Taboada et al., 2000). The model illustrated
dominated by the Romeral Fault System (RFS), which crosses the in Fig. 4 is the preferred one in the literature, but the geometry of the
Northern Andes from Guayaquil up to the Caribbean (e.g. Paris and suture may be complex. This is why an alternative model has been
Romero, 1994; Paris et al., 2000). This SSW–NNE to S–N trending fault proposed where the Romeral fault is a low angle and west dipping
system is inherited from a complex sequence of accretions since the fault (Bourgois et al., 1987; Toto and Kellogg, 1992; Nivia, 1996). The
Cretaceous. At a local scale, this fault system is represented by series of kinematics of the Montenegro, Cauca-Almaguer, Armenia and Silvia-
parallel to subparallel fault segments, sometimes anastomosed Pijao faults are still a matter of debate (Table 1, Fig. 6).
(MacDonald et al., 1996) and forming releasing and restraining The Ibagué fault is the most documented among the “non-
bends (Restrepo and Toussaint, 1985). Between 4°N and 5°N, its Romeral” faults (e.g., Paris and Romero, 1994; Vergara, 1999; Paris
kinematics changes from right-lateral in the south to left-lateral in the et al., 2000; Taboada et al., 2000; Marquínez, 2001; Audemard, 2002;
north (Ego et al., 1995, 1996; Taboada et al., 2000). Some authors have Bohórquez et al., 2005; Gallego et al., 2005; Montes et al., 2005a;
observed a segmentation of the RFS by some NW to SW striking faults Montes et al., 2005b; Diederix et al., 2006). Morphotectonically, the
(James, 1986; Guzmán et al., 1998; INGEOMINAS, 1999; Bohórquez Ibagué fault is characterised by a series of “en-échelon” synthetic
et al., 2005; Gallego et al., 2005; Ospina, 2007). riedels in the Ibagué Fan (Diederix et al., 2006; Montes et al., 2005b).
In the study area and surroundings, published kinematics data of This right-lateral wrench fault has a strong inverse component and
individual faults is complex. For example, in the Quindío department, dips northwards with a high angle at the surface and a lower angle at
fault studies of the RFS in basement rocks show a right-lateral depth (Marquínez, 2001). It cuts across the eastern flank of the Central
kinematics, whereas neotectonic studies indicate a left-lateral kine- Cordillera, where it shows a 29 km long dextral displacement (Figs. 5
matics of the same system. Often the authors do not agree and and 6). Quaternary mafic volcanism is encountered at the junction of
published data are contradictory as shown in Table 1. the Ibagué fault with the eastern segment of the Pericos fault (Núñez
In the study area, the RFS was divided into three sub-systems: The et al., 2001). It may indicate a locally extensive regime and that the
Cali-Patía Fault System on the eastern foothill of the Western Ibagué fault is a crustal structure (Montes et al., 2005b).
Cordillera, the Quebradanueva Fault System in the SSB and its
foothills, and the Cauca-Almaguer Fault System on the western 3.3. Seismicity
foothill of the Central Cordillera.
The Cali-Patía fault system (Fig. 6) follows the western border of The study area is a zone of high seismicity, at least from Pleistocene
the Cauca Valley Basin (Paris and Romero, 1994; Nivia et al., 1995) times (Neuwerth et al., 2006), and numerous earthquakes affected this
down to southern Colombia. Close to the study area, the Toro and region during historical times (Espinosa, 2003). The last strong
F. Suter et al. / Tectonophysics 460 (2008) 134–157 139

Fig. 5. 90-meter resolution DEM based on radar photographs (USGS, 2005) showing lineaments in the Central Cordillera and its surroundings, between 4°N and 5.75°N. Compilation
of data after Guzmán et al. (1998), Bohórquez et al. (2005) and Guarín F. (PhD Thesis, in preparartion), and new observations.

earthquake which affected the area was the 6.2 Mw Armenia 4. Methods for data collection
earthquake of January 25th, 1999. According to Gallego et al. (2005)
and Monsalve and Vargas (2002), its rupture propagated to the north At large scale, lineaments were observed on a 90 m DEM (USGS,
and occurred at a depth of 18 km in an oblique left-lateral normal fault 2005). In the study area, faults and lineaments were observed on a
dipping 67° to the east. Other authors (e.g., Lalinde, 2004) suggest that 30 m DEM (USGS, 2005) and on aerial photographs (flights M-1109,
it might have occurred along an E–W striking right-lateral fault. Based C-1474, C-2483, C-2539, C-2297 and C-2575 of the Agustin Codazzi
on the aftershocks, Gallego et al. (2005) calculated a transpressive Geographical Institute). Orientations of fractures, faults and fault
stress at a depth greater than 10 km (σ1, σ2 and σ3 with respective stria were collected in the field wherever they were encountered.
orientations 308/35, 108/52 and 211/10) and an extensive stress above Detail about data acquisition methods is given in each section
(σ1, σ2 and σ3 with respective orientations 287/72, 145/14, 052/07). below.
Despite these geophysical data, the state of knowledge about the
kinematics, plane strike and dip of faults in the study area remains 5. Results
controversial as shown in Table 1 (e.g., McCourt and Aspden, 1985;
González and Núñez, 1991; Paris, 1997; Guzmán et al., 1998; 5.1. Regional structural data
INGEOMINAS, 1999; Paris et al., 2000; Montes and Sandoval, 2001a;
Montes and Sandoval, 2001b; Vergara et al., 2001; Botero et al., 2004a; A lineament map was drown in a zone located in between the
Vargas et al., 2005). eastern flank of the Western Cordillera and the Magdalena River
140 F. Suter et al. / Tectonophysics 460 (2008) 134–157

Table 1
Compilation of published literature about fault characteristics in the study area and its surroundings. These faults are located in Fig. 6.

Fault Author (s) Kinematics Strike Dip azimuth Site of determination


N to NNE striking faults (Calí-Patía, Quebradanueva and Cauca-Almaguer fault systems)
Toro Caballero y Zapata (1983) Inverse N 20 E W
Ansermanuevo James (1986) Left-lateral N 20 E 75° to the W
Pardo et al. (1994) Right-lateral N 15–20 E
Nivia et al. (1995) Left-lateral Map
Guzmán et al. (1998) Left-lateral NNE to N E
Minor normal
Paris et al. (2000) Left-lateral N 6.6 E E
Bohórquez et al. (2005) Inverse NNE 60° to the W
Quebradanueva Caballero and Zapata (1983) Inverse High angle to the E
James (1986) Inverse N 15 to 25 E 60° to 70° to the E
Left-lateral E
Guzmán et al. (1998) Inverse NNNE E with back-thrusts
to the W
Left-lateral
Pardo et al. (1994) Right-lateral N 15–20 E
Bohórquez et al. (2005) Inverse NNE 60° to the E
Potrerillos Guzmán et al. (1998) Inverse W, back-thrust of In the Quindío department
Quebradanueva fault
Lalinde (2004): ALCALA 1: W side down N 10 E Vertical Hacienda San Felipe
ALCALA 2: Normal NS 74° to the E Relleno Sanitario de Pereira
Sevilla González and Núñez (1991) Approximate
Cardona and Ortiz (1994) N15 E
Montenegro Paris (1997) Left-lateral N 20 E NW
Inverse
Guzmán et al. (1998) Left-lateral N W
Inverse
INGEOMINAS (1999) Inverse N 10 E E
Right-lateral
Paris et al. (2000) Left-lateral N 25 E High angle to the W
Normal
Montes and Sandoval (2001a) Right-lateral N 10 E E
Inverse
Vergara et al. (2001) Left-lateral N 20 E High angle (quite vertical)
Normal to the SE
Cauca-Almaguer Guzmán et al. (1998) Inverse ? NNE E
INGEOMINAS (1999) Inverse N 25 E Vertical
Montes and Sandoval (2001a) Inverse N 25 E Vertical
Bohórquez et al. (2005) Inverse N 50° to the E
Left-lateral
Armenia James (1986) N 35 E Vertical?
Paris (1997) Left-lateral N 20 E NW
Guzmán et al. (1998) NE to N W
INGEOMINAS (1999) Right-lateral N 10 E W
Inverse
Paris et al. (2000) Left-lateral N 23 E High angle to the W
Normal
Montes and Sandoval (2001a) Left-lateral N 10 E W to Vertical
Vergara et al. (2001) Left-lateral NE High angle (quite vertical)
to the SW
Normal
Lalinde (2004) Inverse N 50 E Vertical Alto el Roble sector
Bohórquez et al. (2005) Inverse NNE 60° to the NW
Silvia-Pijao Paris (1997) Inverse High angle to the E
Right-lateral
McCourt et al. (1984) MAP Inverse N 20 E High angle
Right-lateral
González and Núñez (1991) Right-lateral High angle to the E
Inverse
Guzmán et al. (1998) N to NNE
INGEOMINAS (1999) Inverse N to NE E
Right-lateral
Montes and Sandoval (2001a) Inverse N to NE E
Right-lateral
Vergara et al. (2001) Left-lateral N 15 E SE
Normal
Botero et al. (2004a) Vertical? N 15 to 25 E Between Calarcá and Pijao,
Minor right-lateral western foothills of the CC
Ospina (2007) Left-lateral N to NNE High angle to the W N of Calarcá, western foothills
Normal of the CC
Bohórquez et al. (2005) Inverse NNE 60° to the E
Navarco James (1986) Eastern side N 35 E Vertical
higher
Paris et al. (2000) Left-lateral N 18.5 E Vertical
F. Suter et al. / Tectonophysics 460 (2008) 134–157 141

Table 1 (continued)
Fault Author (s) Kinematics Strike Dip azimuth Site of determination

N to NNE striking faults (Calí-Patía, Quebradanueva and Cauca-Almaguer fault systems)


San Jerónimo James (1986) Inverse N 15 E Vertical to 80° to the E
Left-lateral?
González and Núñez (1991) Inverse Map High angle to the E
Right-lateral
INGEOMINAS (1999) Inverse N5W 75° to the E
Left-lateral
Paris (1997) Right-lateral Map
Montes and Sandoval (2001a) Inverse NE 75° to the E
Left-lateral
Vergara et al. (2001) Left-lateral N 15 E High angle to the E
Normal (quite vertical)
Bohórquez et al. (2005) Inverse NNE 65° to the E

ENE to EENE striking faults (Ibagué/Garrapatas type)


Río verde Cardona and Ortiz (1994) N 83 E
Botero et al. (2004a) Likely right-lateral N 85 E Quindío deparment, south
of Calarcá town
Santa rosa James (1986) Left-lateral (?) N 70 E 75° to the SE
Normal (?)
Cardona and Ortiz (1994) N 65 to 70 E 75° to the E
Guzmán et al. (1998) Normal observed NE NW
Inverse right-lateral
Expected
Bohórquez et al. (2005) Left-lateral N 70 E 75° to the SE
Normal
Ibagué Paris (1997) Right-lateral ENE Wrench fault, south side up
Taboada et al. (2000) Right-lateral ENE
Paris et al. (2000) Right-lateral N 67.9 E ± 11° Vertical, south-side up
Slightly oblique
Montes and Sandoval (2001a) N 80 E Vertical, south-side up
Marquinez (2001) Right-lateral Map Steep at the surface,
Inverse decreasing at depth
Audemard (2002) Right-lateral ENE
Bohórquez et al. (2005) Inverse NE 70° to the NW
Right-lateral
Montes et al. (2005a) Right-lateral ENE
Montes et al. (2005b) Right-lateral N 70 E
Garrapatas Taboada et al. (2000) Right-lateral ENE
Paris et al. (2000) Inverse, probably N 60.8 E ± 14° 50° to the NW
right-lateral
Audemard (2002) Right-lateral ENE

WWNW striking faults (Salento type)


Salento González and Núñez (1991) Right-lateral N 81.5 W Vertical
Guzmán et al. (1998) Normal EW 80° to the N
Right-lateral
INGEOMINAS (1999) Normal N 42 W 80° to the N
Right-lateral
Montes and Sandoval (2001a) Normal
Right-lateral
Bohórquez et al. (2005) Normal
Right-lateral

NW striking faults (Otún type)


Rio Arma Guzmán et al. (1998) Normal NW 70° to the NE
Right-lateral
Bohórquez et al. (2005) Left-lateral
Otún James (1986) Lineament N 40 W Vertical (?)
Cardona and Ortiz (1994) Lineament N 56 W
Guzmán et al. (1998) Lineament
Consota James (1986) Lineament N 40 W Vertical (?)
Guzmán et al. (1998) Normal NW 65° to the SW
Right-lateral
INGEOMINAS (1999) NW
Bohórquez et al. (2005) Normal NW 65° to the SW
Right-lateral
Guacaica Guzmán et al. (1998) Normal NW 70° to the NE
Right-lateral
Bohórquez et al. (2005) Normal
Right-lateral

NE striking faults (palestina type)


Palestina Feiniger (1970) Right-lateral N to NNE
Page (1986) Likely left-lateral
during Quaternary
James (1986) Left-lateral since late Tertiary N 50 E Vertical

(continued on next page)


142 F. Suter et al. / Tectonophysics 460 (2008) 134–157

Table 1 (continued)
Fault Author (s) Kinematics Strike Dip azimuth Site of determination
NE striking faults (palestina type)
Palestina Paris (1997) Left-lateral inverse N 10 E
Right lateral to the north (in Antioquia)
Guzmán et al. (1998) Inverse NE 75° to the NW
Left-lateral
Paris et al. (2000) Inverse N 17.8 E ± 11° Moderate to high angle to the W
Left-lateral
Bohórquez et al. (2005) Inverse NE 75° to the NW
Left-lateral
Ospina (2007) Left- Lateral NNE
Rio Dulce James (1986) N 35 E Vertical
Bohórquez et al. (2005) Right-lateral N 30 to 45 E
Rio Roble Guzmán et al. (1998) Inverse NE 55° to the NW
Agua Bonita Guzmán et al. (1998) Inverse NE 55° to the NW
Tapias Guzmán et al. (1998) Inverse NE 40° to the NW
Bohórquez et al. (2005) Inverse NE 40° to the NW
Rio San Juan Guzmán et al. (1998) Inverse NE 40° to the SE
Pericos Montes et al. (2005b) Right-lateral Map

Valley (Fig. 5). It includes the faults and lineaments observed by The Río Verde fault appears as a lineament on the geological map
Bohórquez et al. (2005), Guzmán et al. (1998) and Guarín (in of the Quindío Department (González and Núñez, 1991). It strikes
preparation) and those observed in the SSB (this study, see below). WWSW–EENE and forms the core of a right-stepping, “en-échelon”
Numerous unpublished lineaments could be observed on the radar- system which links the south-western end of the Quindío-Risaralda
based DEM (Fig. 5). The criteria used were the continuity and Fan with the Ibagué Fan. This system comprises, from west to east, the
alignment of the lineaments. The raw data of Fig. 5 are summarized Espejo and Argentina lineaments, the Río Verde fault and the
in Fig. 6 A, highlighting the most important features in this area. Caicedonia lineament (Fig. 6). The Río Verde fault and Caicedonia
Several groups of lineaments have been identified according to their lineament are less easy to follow in the Central Cordillera and seem to
strike. The strike of each lineament was measured and plotted on a be a linear succession of shorter segments (Fig. 5).
rose diagram (Fig. 6 B). The few curved lineaments were divided into North of the Ibagué fault, numerous ENE to EENE striking
smaller segments and two or three strikes were measured. Based on structures are observable in the Central Cordillera (Figs. 5 and 6).
this diagram, eight conspicuous orientations have been identified and Three more families of faults and lineaments are present in
plotted (Fig. 6 C). Finally, the lineaments corresponding to known and the study area and its surroundings. They are sriking NE (Palestina
already described faults were named (Fig. 6). type), WWNW to W (Salento type), and NW (Otún type) (Figs. 5, 6, and
The RFS was divided into two distinct families (Fig. 6); the faults Table 1):
located south of the Ibagué Fault and west of the Quebradanueva Fault
1) The Palestina fault crosses the Central Cordillera in a NE direction
(in pink) have a SSW–NNE strike, whereas the faults located north of
from the north-eastern end of the Quindío-Risaralda Fan. It passes
the Ibagué Fault (in black) have a NS strike.
through the Nevado del Ruiz volcano and bends towards the north
The second group of dominant lineaments is a series of ENE to
in a NNE direction (G. Paris, pers. comm.). It is made of several
EENE striking “Ibagué type” lineaments (in red, Fig. 6). Some of these
parallel and/or aligned segments. North of the Ibagué fault,
lineaments correspond to faults described in the literature (e.g., the
numerous faults and lineaments with a similar strike are present
Garrapatas and Ibagué Faults, Fig. 6), whereas others are only derived
(e.g., the Río Dulce, Río Roble, Agua Bonita, Tapias and Pericos faults;
from the DEM and so far are not proven faults. Locally, they form right-
Fig. 6).
stepping “en-échelon” systems. They crosscut the Western and
2) The Salento fault is right-lateral with a normal component
Central Cordilleras and seem to crosscut all pre-existent structures.
(González and Núñez, 1991; Guzmán et al., 1998; INGEOMINAS,
On the rose diagram of Fig. 6, they show two dominant orientations,
1999; Montes and Sandoval, 2001a; Bohórquez et al., 2005) (Fig. 6,
one EENE similar to that of the Ibagué fault, and one ENE similar to
Table 1). There are numerous lineaments with the same orienta-
that of the Santa Rosa Fault. From a morphotectonic point of view, the
tion in the Central Cordillera north of the Ibagué fault (Fig. 6).
best expressed faults or lineaments are, from west to east, the
3) The Otún trend is made of numerous lineaments in the Central
Garrapatas, the Santa Rosa, the Río Verde and the Ibagué Faults.
Cordillera north of the Ibagué fault (e.g., the Río Arma, Consota, and
The right-lateral Garrapatas Fault crosses the Western Cordil-
Guacaica faults; Fig. 6). Some of these lineaments are observed
lera and outlines the south-eastern end of the CPB. It is curved
faults (dark blue ones in Fig. 6A; Bohórquez et al., 2005). Their
and its strike becomes NNE on the eastern flank of the Western
sense of shear is questionably right-lateral.
Cordillera.
At the eastern foothills of the Western Cordillera, a lineament Finally, numerous WNW to WWNW short lineaments could be
aligned with the Santa Rosa fault (SR in Fig. 6A) takes over from the observed on the DEM (in yellow, Fig. 6). They are not described in the
Garrapatas Fault forming an apparent eastern prolongation of the published literature and were called here “Ocaso” type.
latter. From Pereira, the SRF appears to be continued eastwards by a
series of aligned lineaments with the same strike up to Mariquita in 5.2. Structural data in the SSB and surroundings
the Magdalena Valley (Fig. 6). Although its continuity beneath the
Cauca Valley is not demonstrated, this succession of lineaments 5.2.1. Lineaments
cutting across the CC might be the surface expression of the Santa Rosa Aerial photographs covering parts of the SSB at different scales
Fault. The kinematics of the latter is not well known (Table 1), al- allowed the detection of 397 lineaments affecting the Tertiary and
though Guzmán et al. (1998) suggest that it may be right-lateral and Quaternary Formations (Fig. 7). The criteria used were the alignment
inverse. of the drainage segments, the high degree of incision of drainage with
F. Suter et al. / Tectonophysics 460 (2008) 134–157 143

Fig. 6. A: summary of Fig. 5, where the principal lineaments affecting this area are grouped into families according to their strike. Those corresponding to published faults are named,
as well as the lineaments mentioned in the text. B: Quantity-dependent rose-diagram illustrating the strikes of each family. Their average strike is simplified in C. TSZ: Transform
Shear Zone after Taboada et al. (2000). For names of cities and volcanoes, refer to Fig. 5. The Cali-Patía fault was located after Nivia et al. (2001).

respect to the short travel distance, and the anomalous orientation of NNE, and the “Ibagué” type (Fig. 7). Note that the WNW striking
some drainage segments (i.e. following neither the steeper slopes nor “Ocaso”, the WWNW striking “Salento”, and NW striking “Otún”
the strike of sedimentary beds). families also stand out in this rose diagram.
In the SSB, the dominant lineament orientation corresponds to Thus, the trend pattern of the SSB lineaments resembles to the one
both the “RFS” type, with dominant orientations closer to N than to observed at large scale on the 90-meter DEM (Fig. 6).
144 F. Suter et al. / Tectonophysics 460 (2008) 134–157
F. Suter et al. / Tectonophysics 460 (2008) 134–157 145

Fig. 8. Examples of fractures with no indication of relative displacement; A and B: in conglomerates of La Paila Fm near Obando town, west of the SSB; C: in a Quaternary volcaniclastic
debris-flow deposit of the Quindío-Risaralda volcaniclastic Fan near Pueblo Tapao, east of the SSB.

5.2.2. Fractures In the overlying sediments of the Zarzal Formation and Quindío-
In outcrop, fractures with no indication of relative displacement Risaralda volcaniclastic Fan, the N striking “RFS” fracture set is clearly
were observed in all the sedimentary succession, from the folded dominant, followed by the “Ocaso”, “Ibagué”, “Otún” and “Salento”
Oligo-Miocene rocks of the SSB to the overlying Quaternary Zarzal families. The NNE sriking “RFS”, “Palestina” and “Santa Rosa” types are
Formation and Quindío-Risaralda volcaniclastic Fan (Fig. 8). A the less represented.
maximum of locations were sampled with respect to bad outcrop Therefore, in both the SSB and the Zarzal Formation and Quindío-
conditions and accessibility. Dip and dip azimuth of 639 fracture Risaralda volcaniclastic Fan, there is a good correlation between the main
planes were measured in the field and plotted in rose diagrams fracture sets and the lineament families observed in the 90-meter DEM at
(Fig. 9). The latter allow the comparison between the orientation of larger scale (Figs. 5 and 6) and on aerial photographs (Fig. 7) in the SSB.
the fracture sets and those of the eight lineament families observed at
large scale on the 90-meter-resolution DEM (Fig. 6 C). Although 5.3. Faults in Oligo-Miocene folded rocks
fractures with strike ranging from 0 to 360° were encountered, some
dominant fracture orientations stand out in the diagrams. In the field, 33 sites were encountered where rocks present striated
In the SSB, despite the great dispersal of the strike values, the fault planes indicative of relative movement. The kinematics of each
“Palestina” family is the most represented type, followed by the fault was determined through crystallisation fibres, steps impressed in
“Salento” type, the NNE sriking “RFS” family and the “Otún” type. the plane by striator objects, riedel criteria and stylolithic stairs
Fractures belonging to the “Ibagué”, “Santa Rosa” and N striking “RFS” (Hancock, 1985; Petit, 1987). Because of the thick vegetal cover and
trends are less represented than in the large scale DEM observations, strong weathering through tropical climate, the best outcrop sites
suggesting that these fault systems are more localised. were encountered in river beds during the dry season (Fig. 10).

Fig. 7. 30-meter resolution, radar-based DEM (USGS, 2005) covering the study area. The principal structural features of the SSB are shown in white, and the lineaments observed on
aerial photographs in black. The rose diagram allows the comparison of their orientation with the different families observed in Fig. 6. Numbers correspond to the stereoplots of Fig. 11.
146 F. Suter et al. / Tectonophysics 460 (2008) 134–157

Fig. 9. Dip and strike of the 639 fractures measured, which do not exhibit any evidence of relative displacement. Fractures measured in the folded Oligo-Miocene SSB rocks (A) and in
the Quaternary Zarzal Formation and the Quindío-Risaralda volcaniclastic Fan (B) are compared with the orientations of the lineament families observed at large scale (see Fig. 6).

All together, 488 fault planes and their slickensides were measured eter Ф = (σ2 − σ3)/(σ1 − σ3) and the average misfit angle between
(Fig. 11). observed and calculated striae.
In order to determine the paleostress axis orientations, the “direct For each fault and stria dataset, four factors were taken into
inversion method” (Angelier and Goguel, 1979) was used (“Tectonics account to evaluate the quality of the result after inversion: 1) in
FP” software, Ortner et al., 2002). For each paleostress tensor, the terms of paleostress axis orientation, the result of the direct
direction of the major, intermediate and minor stress axis (σ1, σ2 and inversion method has to give results comparable with the graphical
σ3 respectively) were calculated, as well as the ellipsoid form param- inversion method (Angelier and Mechler, 1977); 2) the second

Fig. 10. A: Site in the SSB where faulted rocks display good slickensides; B: Fault plane: s: stylolithic stairs; f: crystallization fibres. The black arrow shows the motion of the missing
compartment with respect to that in the picture.
F. Suter et al. / Tectonophysics 460 (2008) 134–157
Fig. 11. Stereoplots representing the dip, dip azimuth, and kinematics of fault planes at each site numbered in Fig. 7, as well as the orientation of their calculated paleostress tensors (Wulff stereograms, lower hemisphere); black dots: σ1; white

147
squares: σ2; black triangles: σ3. A total of 488 fault planes were measured.
148 F. Suter et al. / Tectonophysics 460 (2008) 134–157

Table 2
Parameters of the 29 calculated stress tensors, with the name of the site, the orientations of σ1, σ2 and σ3, the corresponding ellipsoid form parameter (Ф) and the stress regime the
tensor belongs to. The number of faults used for the calculation appears in the “n” column; Var (°) indicates the average misfit angle after the final calculation. A quality criterion
between 1 (good) and 3 (bad) is given for the result.

N° Site σ1 σ2 σ3 Ф Stress regime n Var (°) Quality


σ3 steep, compressional tensors
1 Palmarito 340/11 070/01 164/79 0.91 Radial compression 20 6.35 2
3 El Vergel 316/02 047/07 209/83 0.4 Pure compression 15 7.4 1
5 Cruzes 293/06 202/14 046/75 0.27 Pure compression 23 5.61 2
6 NE Obando 270/06 179/10 031/78 0.38 Pure compression 16 20.46 2
7 El Infierno 098/11 005/14 223/72 0.65 Pure compression 12 7.17 2
11 La Cascada Sur 112/10 205/13 346/73 0.47 Pure compression 15 6.93 1
12 SE Villarodas ramal E 008/10 271/36 111/52 0.94 Radial compression 7 3.57 2
15 Nicaragua N Playa Azul II 296/31 190/26 068/48 0.7 Pure compression 5 3 3
16 Nicaragua N Playa Azul I 274/19 013/24 149/58 0.58 Pure compression 7 6.29 3
18 Playa Azul Sol y JuanK 202/06 111/07 330/81 0.88 Radial compression 4 0 3
19 Nicaragua II 162/20 254/04 356/69 0.3 Pure compression 7 7.71 3
20 Nicaragua Rhyolite 246/05 336/01 075/85 0.86 Radial compression 4 0 3
21 Nicaragua Grès 255/11 346/07 108/77 0.81 Radial compression 18 8.22 1
22 La Palma-Areneros 267/13 358/04 106/76 0.78 Radial compression 19 4.95 1
23 "Sevilla" 242/15 149/10 028/72 0.76 Radial compression 9 5.78 1
25 La Olanda "La Maria" 194/05 104/03 341/85 0.89 Radial compression 14 3.79 2
26 La Olanda Quebrada 1 129/02 039/17 226/73 0.38 Pure compression 8 4.38 1
27 Barcelona 5 013/09 281/13 137/74 0.83 Radial compression 11 5.64 1
28 Barcelona 4 248/07 158/00 067/83 0.91 Radial compression 15 4.27 2
29 Barcelona 2 324/05 055/20 219/69 0.9 Radial compression 10 2.5 1

σ2 steep, strike-slip tensors


2 Chimichagua 117/20 338/65 213/15 0.15 Transpression 9 6.33 2
8 Agua Dulce 278/05 174/70 010/20 0.36 Pure strike-slip 9 5.67 1
9 Mona Recochadora 109/22 225/47 002/34 0.25 Transpression 4 0 3
14 SE Villarodas ramal W 199/04 079/83 290/06 0.86 Transtension 8 9.13 2
17 Playa Azul Murciélagos y Quebrada 132/01 223/49 042/41 0.05 Transpression 4 0 3

σ1 steep, extensional tensors


4 Entre Cruzes y Zaragoza 031/71 196/19 287/05 0.95 Transtension 18 5.06 1
10 La Cascada N 140/42 009/37 257/27 0.07 Radial extension 10 2.3 2
13 La Cascada más abajo 321/74 202/08 110/13 0.87 Transtension 9 3.44 2
24 La Ilusión I y II 153/74 033/08 301/14 0.03 Radial extension 14 4 2

factor is the stability of the result with respect to particular faults.


Some faults have a strong influence on the result when added
or removed from the dataset. In this case the quality criterion
would be bad; 3) the third factor is the misfit angle mean for one
dataset. Individual faults with angles bigger than 20° were removed
from the dataset for the final calculation. If too many faults had to
be removed, the result was considered as bad; 4) the last factor is
the number of faults used for the inversion calculation. A dataset
with less than 8 faults would automatically give a low quality
result.
Based on these four factors, a quality value between 1 (good)
and 3 (bad) was attributed to each calculated tensor (Table 2), and
four datasets were discarded from the data because they showed
too bad results. The 29 remaining sites, together with the main
structural features of the SSB, are located in Fig. 7. They are
numbered in relation to the corresponding stereoplots shown in
Fig. 11.
Knowing the ellipsoid form parameter Ф= (σ2 − σ3)/(σ1 − σ3) and
the steeper strain vector among σ1, σ2 and σ3, it was possible to classify
each site into one of seven tectonic regimes (Fig.12, Table 2) (Champagnac
et al., 2003). Out of the 29 selected sites, eleven show radial compression,
nine pure compression, seven strike-slip (transpressive, pure or
transtensive) and two radial extension. Finally, the repartition and
orientation of σ1, σ2 and σ3 in the SSB were plotted on maps (Fig. 13).
The most interesting results can be summarized as follows: 1)
most of the σ1 are horizontal and have a dominant WNW–ESE
direction (in the case of radial compression they could be permuted
with σ2, because σ1 and σ2 are both horizontal and of similar
Fig. 12. Histogram representing the number of tensors belonging to each of the 7
intensity, Fig. 14); 2) most of the σ3 are vertical and their calculated existing stress regimes versus their respective ellipsoid form parameter Ф = (σ2 − σ3)/
tensors indicate radial or pure compression; 3) most of the striated (σ1 − σ3). Most of the tensors occur in compressive stress regimes.
F. Suter et al. / Tectonophysics 460 (2008) 134–157 149

Fig. 13. Distribution of the 29 calculated palaeostress tensors.

fault planes have strikes similar to essentially the “RFS” family, and Widespread soft-sediment deformations affect these sediments
secondarily the “Palestina”, “Santa Rosa” and “Otún” families (Fig. 11). all over the study area (Neuwerth et al., 2006). At some locations,
clastic injections in relation with these faults or load casts in
5.3.1. Faults in flat-lying Pleistocene deposits the underlying strata highlight the relationship of these exten-
In the flat-lying interpenetrating sedimentary successions of the sional faults with the widespread soft-sediment deformations
Zarzal Formation and Quindío-Risaralda volcaniclastic Fan, only (Fig. 15). At other locations, this relationship could not be observed
extensional faulting was observed. All these conjugate extensional (Fig. 16).
faults are small in size. In cross section, they reach lengths from Plots of the fracture plane poles show that the faults belong to
centimetres to tens of metres. Fault planes are mainly straight, conjugate systems (Fig. 17). In such a case, the mean vector projection is a
forming small grabens and horsts, sometimes slightly or strongly good estimation of the local σ3. The results show dominant NS (±10°) and
curved; their inclination varies from vertical to about 45°. They affect EW (±20°) directions for extensional vectors (local σ3). The remaining
all kind of sediments and are sometimes filled by fault breccias extensional vectors range between these two predominant orientations.
(Fig. 15). With respect to the strike of these conjugate extensional faults
150 F. Suter et al. / Tectonophysics 460 (2008) 134–157

South of the Quindío-Risaralda volcaniclastic Fan, a 7 km long–3 km


wide reverse “L” shaped alluvial plain contrasts with the surrounding
deposits (Fig. 18 C). This alluvial plain is sharply cut on its edges and is
closed at its outlet. Consequently it must have been opened after the
deposition of the southern end of the Quindío-Risaralda volcaniclastic
Fan. One possible mechanism allowing such a basin to open in the
compressive state of stress known for the area would be the right-
lateral kinematics of the right-stepping “en-échelon” Argentina-Río
Verde-Caicedonia system (Fig. 6), thereby forming a pull-apart basin. It
is named here the Caicedonia pull-apart basin. Nevertheless, although
the shape of this alluvial plain supports a local Quaternary extensional
or transtensional tectonic activity, field data documenting the right-
lateral motion of these lineaments is lacking.

6. Discussion

6.1. Common strain pattern at different scales

Because of the high number of family trends, one has to be cautious


when making correlations between lineament families at large and
small scales. Nevertheless the orientation of lineaments in the SSB
Fig. 14. Projections of the 29 σ1 stress axis obtained by the inversion calculations. They show a similar angular pattern to those observed at larger scale (Figs. 6
are divided according to the type of tensor they belong to. These σ1 stress axes suggest and 7). Additionally, the outcrop-scale fracture sets in the SSB, those in
a WNW–ESE trend. This trend would be enhanced if the four σ1 stress axes symbolised the Quaternary interpenetrating Zarzal Formation and Quindío-
by a cross would be rotated of 90°.
Risaralda volcaniclastic Fan (Fig. 9), as well as the faults used for the
calculation of the paleostress tensors (Fig. 11), also show the dominant
(Fig. 17), most of them have an orientation similar to that of the “RFS”, orientations. In order to illustrate this similarity, the angular gaps of
“Ibagué” and “Otún” families observed on the 90-meter DEM (Fig. 6). each rose diagram in which no (or very few) lineaments, fractures or
faults appear were determined and compared (Fig. 19). These gaps
5.4. Neotectonics occur between NW and N strikes, rotating clockwise. In other words,
no features develop between the “Otún” and “RFS” strike families.
Northwest of the SSB, a prolongation of the west verging This similarity in lineament and fracture orientation, at each scale
Quebradanueva thrust belonging to the N striking “RFS” (Figs. 4, 6 and in each formation, suggests that a similar strain pattern may affect
and 18) forms a step and clearly outlines the mountain front of this the area delimited by the Western Cordillera eastern foothills and the
small fold and thrust range. On the 30-meter DEM, it is easy to follow Magdalena Valley between 4.5°N and at least 5.5°N.
its strike trough the Cartago Fan (Fig. 18 A and B). This is the clearest The relative differences in occurrence frequencies of the fractures
indication of Quaternary activity of this fault. in each unit cannot be interpreted as time dependent, genetic var-
North and south of the Cartago Fan, the mountain front sinuosity iability. Conversely, they are due to variability in the spatial dis-
(Keller and Pinter, 2002) was calculated. A straight mountain front tribution of fractures. The predominance of N striking RFS fractures in
develops when the tectonic activity dominates upon the erosion rate. the younger Zarzal Formation and Quindío-Risaralda Fan, and its low
The following relation compares the sinuosity of the mountain front frequency in the underlying formations (Fig. 9) can only be explained
with a theoretical straight line: Smf = Lmf/Ls, where Smf is the mountain in this way.
front sinuosity, Lmf the length of the mountain front at the slope break at
the foot of the mountain, and Ls the straight-line length of the mountain 6.2. Paleostress inversions — WNW–ESE trending maximum horizontal
front. North of the Cartago Fan, the Smf is 1.3 and south of this Fan, it has stress
a value of Smf = 2.2. A Smf between 1 and 1.6 characterizes mountain
fronts associated with active bounding faults. A less active mountain The paleostress inversions obtained from striated fault sets in the
front but still reflecting active tectonics would have a Smf ranging from folded Oligo-Miocene rocks of the SSB gave a mean WNW–ESE
1.4 to 3 (Keller and Pinter, 2002). Therefore, the northern mountain front trending maximum horizontal stress (Fig. 14), perpendicular to the
is apparently more active than the southern one. fold axial traces, with 69% of compressive tensors (Figs. 12 and 13).
Furthermore, using aerial photographs, drainage inversions were This result is consistent with the paleostress determinations obtained
observed in the distal part of the Quindío-Riseralda Fan (Fig. 18 B). by the previous authors in this area (Guzmán et al., 1998;
They are located on the western termination of this fan onto the SSB. INGEOMINAS, 1999; Paris et al., 2000; Monsalve and Vargas, 2002;
The geomorphology highlights that the thickness of the fan is low and Lalinde, 2004; Bohórquez et al., 2005; Cortes and Angelier, 2005;
it is easy to guess the strike of the underlying Oligo-Miocene layers. In Cortes et al., 2005; Gallego et al., 2005; Montes et al., 2005b; Vargas
this part of the Fan, the general drainage direction is WNW. et al., 2005). Considering that these data were taken exactly at the CPB
Nevertheless, slightly west of the Potrerillos fault trace, much of the indentation front, they are coherent with an E to ESE motion direction
tributaries alimenting the principal rivers flow in a SE to NE direction. of the CPB with respect to South America.
These three indices of tectonic activity on the northern termina- Because features generated by this stress pattern are observed in rocks
tion of the SSB show that this fold and thrust range was still uplifted of Pleistocene age, this pattern is assumed to have been active at least
after the deposition of the Cartago Fan and distal part of the Quindío- from Lower Miocene to Pleistocene and probably it is still active today.
Risaralda Fan. At this distal location with respect to the rest of the Fan,
the drainage incision is not so sharp indicating a younger age. Thus, 6.3. Romeral fault system (RFS), change of strike and neotectonic activity
the uplift of the northern end of the SSB associated with the
Quebradnueva thrust and Potrerillos back-thrust (Fig. 4) has been The RFS is inherited from a long history of accretions. In the study
continuously active during Quaternary times. area, many authors have noted a segmentation of this fault system by
F. Suter et al. / Tectonophysics 460 (2008) 134–157 151

Fig. 15. Examples of extensional features observed in the Zarzal Formation on both sides of the SSB; numbers 3, 5a and 6 in black circles refer to sites in Fig. 17: A: double sand injection
associated with extensional faulting, white arrows indicate the sense of injection, geologist and coin for scale; B: white arrows indicate stratigraphic markers allowing the evaluation
of the apparent fault heave. These normal faults are associated with load casts; field book for scale; C: clastic injection following the fault plane; white arrows indicate the direction of
injection, and fault breccia. Pens for scale.

some NW to SW striking faults (James, 1986; Guzmán et al., 1998; with probably small strike-slip components (e.g. Quebradanueva thrust
INGEOMINAS, 1999; Bohórquez et al., 2005; Gallego et al., 2005; Ospina, and Potrerillos back-thrust; Figs. 4 and 18).
2007). The strike of the RFS is perpendicular to subperpendicular to the This study permitted the detection of a change in strike of the
calculated maximal horizontal stress (Fig. 19). From a neotectonic point Romeral faults: north of the EW line marked by the Garrapatas, Río
of view, it is active. It behaves principally as thrusts and back-thrusts Verde and Ibagué right-lateral active faults, they strike N ± 10° (black
152 F. Suter et al. / Tectonophysics 460 (2008) 134–157

Fig. 16. Example of extensional dislocations not showing any relationship with soft-sediment deformations, hammer for scale. Number 4 in black circle corresponds to location in Fig. 17.
F. Suter et al. / Tectonophysics 460 (2008) 134–157 153

Fig. 18. Neotectonic overview of the study area: A: mountain front sinuosity (Smf = Lmf/Ls) of the SSB north and south of the Cartago Fan. The lengths (Lmf) and straight-line lengths
(Ls) of the mountain front are drawn in white and black respectively, their values are in quantity of pixels. The Quebradanueva thrust and Potrerillos back-thrust are highlighted by
the white arrows; B: drainage inversion due to recent uplift of the northern end of the SSB; C: Interpretation of the right-lateral Caicedonia pull-apart basin.

lineaments, Fig. 6); south of this line, their strike ranges between 020 family with an angle of 66° would be the antithetic riedels (R'). In-
and 040 (pink lineaments, Fig. 6). between lie the “Salento” and “Ocaso” fault families, which are parallel
Fractures, lineaments and faults striking similarly to the RFS are to subparallel to the maximum horizontal stress σ1. The “Palestina”
present at all scales and in each formation studied in this study fault system forming an angle of 18° anticlockwise with respect to the
(Figs. 6, 7, 9, 11, 17 and 19). “Santa Rosa” family could be interpreted as synthetic P-shears.

6.4. Angular relationship between “non-Romeral” structures 6.5. Distributed shear strain in the Central Cordillera north of 4.5°N

The angular setting of the strike families not related to the RFS All these “non-Romeral” lineaments occur in the Central Cordillera
(Fig. 19) resembles strongly to the angular arrangement of faults in a north of 4.5°N (Fig. 6). They are not the consequence of any major
right-lateral shear system (Tchalenko, 1970; Hancock, 1985; An and particular strike-slip discontinuity but respond to a right-lateral
Sammis, 1996; Schreurs and Colletta, 2002; Rovida and Tibaldi, 2005; distributed shear strain.
Van der Pluijm and Marshak, 2004). Although good field data at the Various laboratory simulations of shear zones have been published
front of the CPB N of 4°N are lacking (Table 1, Figs. 5 and 6), this (e. g., Tchalenko, 1970; An and Sammis, 1996; An, 1998; Schreurs and
angular setting suggests that these families are associate faults, where Colletta, 2002). They show that shearing is a dynamic process and
the ENE striking right-lateral “Santa Rosa” fault family indicates the demonstrate that the angles between the principal shears (R, R′,
shear direction and corresponds to the main fault orientation. and P) vary slightly from one experiment to other as a function of the
Consequently, the “Ibagué” fault family forming a 19° angle with the material used and its water content (An and Sammis, 1996). The
main fault would be the synthetic riedels (R) and the “Otun” fault angular setting of shears observed at different scales in this study fits

Fig. 17. 30-meter resolution DEM (USGS, 2005) with location of the sites where conjugate normal faults planes were measured in the Quaternary Zarzal Formation and Quindío-Risaralda
volcaniclastic Fan. The black dots in the plots (Wulff stereonets, lower hemisphere) are projections of the fault plane poles. The plunge of their mean vector (calculated after Wallbrecher
(1986)) indicating the direction of elongation (local σ3) is represented by the red asterisks. The dip and dip azimuth of the latter appears besides each plot. The values given for the sites
where only one single fault plane could be measured (numbers 2, 6, 7, 9; black arrows on map) correspond to the dip and dip azimuth of the fault planes. The directions of local σ3 are
symbolized on the DEM by the arrows. The rose diagram shows the strikes of these 85 normal faults. C: Cartago; Q: Quimbaya; M: Montenegro; O: Obando; T: La Tebaida; Z: Zarzal.
154 F. Suter et al. / Tectonophysics 460 (2008) 134–157

Fig. 19. Summary of the data obtained in this study and comparison with the theoretical fault pattern developed under right-lateral shear. A) The gaps of the rose-diagrams obtained
for the fractures, lineaments and faults, at each scale and in each lithology, have been stacked with some degree of transparency in order to evidence their main angular range. Data of
the normal faults affecting the superficial Pleistocene sedimentary rocks are not included here because the quantity of measured faults is statistically not sufficient (85 over 12 spots).
The grey arrows show the direction of the maximum horizontal stress (σ1) obtained by the inversion methods. B) Classical angular organisation of Riedel shears in a right-lateral
shear system (after Tchalenko (1970), Harding (1974), Hancock (1985), An and Sammis (1996), Schreurs and Colletta (2002), Schreurs (2003)). If one adds the observed gaps with the
Romeral Fault interval, the resulting angular range corresponds to the theoretical gap existing within a set of Riedel shear fractures in an ENE trending right-lateral shear zone.

with the angle ranges obtained in these experiments. Moreover, the have a right-stepping “en-échelon” arrangement, which is an ap-
orientation of the calculated maximum horizontal stress bisects the propriate condition to form right-lateral negative flower structures.
presumed synthetic and antithetic Riedels (R and R′, i.e., Ibagué and From a neotectonic point of view, the right-lateral Ibagué and
Otún familes; Fig. 19), which is in agreement with these models. Garrapatas faults are active (Paris et al., 2000; Taboada et al., 2000;
During the dynamic process of shearing, the synthetic P-shears Montes et al., 2005b), and the Caicedonia pull-apart basin is a proof of
always appear after the R and R′, at a slightly more evolved stage of the Quaternary activity south of the Quindío-Risaralda volcaniclastic Fan
shear. Here, at large scale (Fig. 6), the density of “Palestina” lineaments (Fig. 18). The Río Verde fault seems to be active as well (Cardona and
seems to decrease slightly towards the north. Therefore, it is possible Ortiz, 1994; Botero et al., 2004b; Lalinde, 2004).
that shearing in the north is less developed than in the south, and that North of the E–W line marked by these right-stepping “en-
the onset of this distributed shear strain migrates towards the north échelon” features, the Riedel shears are more abundant (Fig. 6); the
with time. During shear evolution, the angular range where faults do RFS changes its strike and is segmented. This E–W line probably
not appear stands between the R' and the P-shears (Tchalenko, 1970; corresponds to the southern limit of the area affected by the dis-
An and Sammis, 1996; An, 1998; Schreurs and Colletta, 2002). In the tributed shear strain.
data obtained here, this angular gap is narrower because the
accretion-inherited faults of the RFS are present (Fig. 19) and still 6.8. Geodynamic hypothesis
active.
The processes of oblique subduction and accretion which led to the
6.6. Simultaneous activity of Romeral and “non-Romeral” fault systems existence of the RFS as a Cretaceous suture cannot explain the pre-
sence in the Central Cordillera, north of 4.5°N, of such a distributed
Both the Romeral and “non-Romeral” fault systems have neotec- shear strain. To overprint the inherited N to NNE striking Romeral
tonic expression (see section 5.3), thereby suggesting that there is faults, the Riedel shears must be younger and respond to another
strain partitioning between these “non-Romeral” shear faults and the major geological constraint.
accretion-inherited RFS. Following the convergence of South America and North America,
The RFS is a long Cretaceous suture extending from Guayaquil up which began during the Eocene, the underplating of the Caribbean
to the Carribean Sea, whereas “non-Romeral” faults are present in Plate below South America in an E to ESE direction led to the collision
the Central Cordillera north of 4.5°N, at the indentation front of the and indentation of the CPB onto the Western Cordillera (Fig. 20). The
CPB (Figs. 2 and 6). Such a shear system is independent; it does not precise age of this collision is not well defined; it ranges between the
need the RFS faults to work, and in the study area, the RFS is Miocene and the Pliocene.
segmented. It is likely that the active faults of the E–W Garrapatas-Ibagué line
first developed at the initial stage of the collision, when the CPB was
6.7. E–W active right-lateral strike-slip fault zone at 4.5°N still an active volcanic arc and its eastern termination first collided
with the NW corner of South America. Because the convergence
The Garrapatas, Ibagué and Río Verde faults, together with the between North and South America continued and the buoyancy of the
Espejo, Argentina, and Caicedonia lineaments form an E–W fractured CPB prevented its subduction, the latter began to deform so as to
zone at 4.5°N, crosscutting the Western and Central Cordilleras as well acquire a lazy-S shape. This generated a left-lateral shear zone in
as the Cauca and Magdalena Valleys (Fig. 6). These right-lateral faults eastern Panamá (Mann and Corrigan, 1990). Because the indentation
F. Suter et al. / Tectonophysics 460 (2008) 134–157 155

volcaniclastic Fan. Their occurrence seems odd in such a tectonic


context.
Nevertheless, translatory slides triggered by earthquakes could
occur (Hansen et al., 1965). They affect horizontal layers limited on one
side by a cliff or a steep slope, when these are underlain by a layer with
high water content and low shear strength. During tremors, this layer
looses its cohesion causing the extensional break up of the overlying
layers. This phenomena is called lateral spreading and is often
accompanied by soil liquefaction and soft-sediment deformations,
such as sand volcanoes or contorted layers (Audemard and De Santis,
1991; Gonzalez et al., 2004; Rastogi, 2004; Audemard et al., 2005).
Geomorphology shows that the subhorizontal sediments of the
Zarzal Formation and Quindío-Risaralda volcaniclastic Fan were
eroded after their deposition, leading to small shallow valleys bounded
by low-relief cliffs (Fig. 3). A significant seismicity took place during the
Pleistocene as shown by the widespread soft-sediment deformations
affecting these sediments (Neuwerth et al., 2006).
These observations suggest that in eroded flat-lying deposits
containing layers prone to liquefaction, the lateral spreading is the
most probable mechanism able to break out the superficial part of the
sedimentary sequence in the Cauca Depression and Quindío-Risaralda
Basin.
Most of the extensional faults observed strike similar to the main
lineament and fault families observed at all scales and in all lithologies
(Fig. 17). Therefore, it is likely that pre-existent fracture planes linked
to the regional shearing processes have been reactivated as normal
faults.

7. Conclusions

The existence of syntectonic deposits in the Cauca Valley Basin has


permitted the reconstruction of its kinematic evolution from Late
Oligocene to Recent times.
The inherited Cretaceous faults of the Romeral Fault System played
a fundamental role in the development of the Serranía de Santa
Barbara at least since Oligo-Miocene times. The folding and thrusting
in the Serranía are still active today and are evidenced by the N to NNE
striking Quebradanueva thrust and Potrerillos back-thrust.
At 4.5°N, the younger, major E–W right-lateral active Garrapatas-
Ibagué Fault zone crosscuts the Western and Central Cordilleras from
the Pacific coastline up to the Magdalena Valley. This set of EENE
trending, right-lateral, right-stepping, “en-échelon” active faults
Fig. 20. Simplified kinematic reconstruction of the northwestern corner of South clearly segments and affects the Romeral Fault System, which changes
America from Lower Miocene to present times illustrating the collision of the Chocó- strike at 4.5°N.
Panamá Block (after Mann and Corrigan (1990), Taboada et al. (2000), Moreno and Pardo
(2002)). Following the onset of the collision, while the shortening between North and
North of this E–W trending fault zone, the polymetamorphic rigid
South Americas continues, the Chocó-Panamá Block bends up instead of subducting. core of the Central Cordillera as well as the overlying Cenozoic
This produces a left-lateral shear zone in eastern Panamá (Mann and Corrigan, 1990) and sediments are affected by lineaments and faults having an angular
a right-lateral distributed shear strain (DSS) in the rigid polymetamorphic Central arrangement typical of Riedels in a right-lateral shear system. The
Cordillera of Colombia at the indentation front of the CPB. The pairs of double black half
distributed right-lateral shear strain strikes at around 060, as well as
arrows are oriented according to the distributed shear direction.
the Santa Rosa fault. The northern termination of this sheared portion
of crust is not defined yet. These shear faults and lineaments are
superimposed upon and crosscut the Romeral Fault System. They
of the CPB went on together with the underplating of the Caribbean probably formed consecutive to the collision of the Chocó-Panamá
Plate, the rigid metamorphic core of the Central Cordillera began to Block.
shear under the effect of an ENE oriented distributed right-lateral The paleostress inversion calculations gave a WNW–ESE trending
shear strain (Fig. 20). maximum horizontal stress, which is (sub)-perpendicular to the
The Riedel shears generated by this process were superimposed on Romeral thrusts and bisects the antithetic and synthetic Riedels of the
the pre-existent and still active N to NNE striking Romeral Fault right-lateral shear system.
System, thus subsequently segmenting it. This shear system was active The kinematic reconstruction of the study area indicates that the
at least until the Middle Pleistocene and it is probably still active Cauca Valley Basin may have been formed by the northern closing up
today. of the valley following the indentation of the Chocó Panamá Block and
the eastward shift of the Western Cordillera. Subsequently, it would
6.9. Extensional faulting in the Quaternary sediments have been rapidly filled with alluvial sediments.
In contrast to the compressive regional tectonic context, the super-
Extensional features were observed at surface in the subhorizon- ficial sediments of the Zarzal Formation and Quindío-Risaralda volcani-
tal soft sediments of the Zarzal Formation and Quindío-Risaralda clastic Fan are affected by extensional faulting due to soil liquefaction and
156 F. Suter et al. / Tectonophysics 460 (2008) 134–157

lateral spreading. These phenomena are linked to the continuous Cortes, M., Angelier, J., 2005. Current states of stress in the northern Andes as indicated
by focal mechanisms of earthquakes. Tectonophysics 403, 29–58.
seismicity which affected the study area throughout its evolution. Cortes, M., Angelier, J., Colletta, B., 2005. Paleostress evolution of the Northern Andes
The results presented here bring new information about the (Eastern Cordillera of Colombia); implications on plate kinematics of the South
understanding of the seismic hazard in this area; Importantly and in Caribbean region. Tectonics 24 (TC1008), 1–27.
Diederix, H., Audemard, F., Osorio, J., Montes, N., Velandia, F., Romero, J., 2006.
addition to the Romeral Fault System that is thought to be the most likely Modelado morfotectónico de la falla transcurrente de Ibagué, Colombia. Revista de
source of earthquakes, the more recent cross-cutting, overprinted fault la Asociación Geológica Argentina 61–4, 492–503.
systems described here in are another potential hazard to be considered. Droux, A., Delaloye, M., 1996. Petrography and geochemistry of Plio-Quaternary calc-
alkaline volcanoes of Southwestern Colombia. Journal of South American Earth
Sciences 9, 27–41.
Acknowledgements Duque-Caro, H., 1990. The choco block in the northwestern corner of South America:
Structural, tectonostratigraphic, and paleogeographic implications. Journal of South
American Earth Sciences 3, 71–84.
The authors are indebted to A. Espinosa for his support in the field and
Ego, F., Sebrier, M., Yepes, H., 1995. Is the Cauca-Patia and Romeral fault system left- or
in the lab, to F. Audemard and J.-M. Cortés for fruitful discussions, to H. right-lateral? Geophysical Research Letters 22, 33–36.
Echeverri for his technical support during fieldwork, to Jacques Metzger Ego, F., Sebrier, M., Lavenu, A., Yepes, H., Egues, A., 1996. Quaternary state of stress in the
for his graphical work, to Gabriel Paris, Martín Cortés and Camilo Montes Northern Andes and the restraining bend model for the Ecuadorian Andes.
Tectonophysics 259 (1–3), 101–116.
for their reviews which considerably helped to improve the manuscript, Espinosa, A., 2003. Historia Sísmica de Colombia 1550–1830. Academia Colombiana de
and to J. I. Martínez. This research is supported by the Swiss Ciencias Exactas, Físicas y Naturales.
National Science Foundation (grants no. 21-67080.01 and 20-107866.05). Estrada, J.J., Viana, R., 1998. Mapa geológico de Colombia, Plancha 205 — Chinchiná.
INGEOMINAS, Medellín.
Feiniger, T., 1970. The Palestina Fault, Colombia. Geological Society of America Bulletin
81, 1201–1216.
References Freymueller, J.T., Kellogg, J.N., Vega, V., 1993. Plate motions in the North Andean region.
Journal of Geophysical Research, B, Solid Earth and Planets 98, 21,853–21, 863.
Acosta, C.E., 1978. El graben interandino Colombo-Ecuatoriano (Fosa Tectónica del Gallego, A., Ospina, L.M., Osorio, J., 2005. Sismo del Quindío del 25 de enero del 1999,
Cauca Patía y el corredor Andino-Ecuatoriano). Boletin de geología, Universidad evaluación morfotectónica y sismológica. Boletin de Geología, Universidad Industrial
Industrial de Santander 12, 63–199. de Santander (UIS) 27, 133–150.
Alfonso, C.A., Sacks, P.E., Secor, D.T., Rine, J., Perez, V., 1994. A tertiary fold and thrust belt González, H., Núñez, A., 1991. Mapa geológico generalizado del departamento del
in the Valle del Cauca Basin, Colombian Andes. Journal of South American Earth Quindío. INGEOMINAS, Bogotá.
Sciences 7, 387–402. Gonzalez, J., Schmitz, M., Audemard, F.A., Contreras, R., Mocquet, A., Delgado, J., De
An, L.J., 1998. Development of fault discontinuities in shear experiments. Tectonophysics Santis, F., 2004. Site effects of the 1997 Cariaco, Venezuela earthquake. Engineering
293, 45–59. Geology 72, 143–177.
An, L.J., Sammis, C.G., 1996. Development of strike-slip faults; shear experiments in Guarín, F., Gorin, G., Espinosa, A., 2004. A Pleistocene stacked succession of
granular materials and clay using a new technique. Journal of Structural Geology 18, volcaniclastic mass flows in central Colombia: the Quindío-Riseralda fan. Acta
1061–1077. Vulcanologica 16 (1–2), 109–124.
Angelier, J., Mechler, P., 1977. Sur une méthode graphique de recherche des contraintes Guarín, F., (in preparation). Sedimentological and tectonic studies in the Quindío-
principales également utilisable en tectonique et en séismologie: la méthode des Risaralda Fan. PhD Thesis, Terre et Environnement, Université de Genève (in
dièdres droits. Bulletin de la Société géologique de France 7, 6. French).
Angelier, J., Goguel, J., 1979. Sur une methode simple de determination des axes Gutscher, M.A., Malavieille, J., Lallemand, S., Collot, J.Y., 1999. Tectonic segmentation of
princepaux des contraintes pour une population de failles. Comptes Rendus the North Andean margin; impact of the Carnegie Ridge collision. Earth and
Hebdomadaires des Seances de l'Academie des Sciences, Serie D: Sciences Planetary Science Letters 168, 255–270.
Naturelles 288, 307–310. Guzmán, J., Franco, G., Ochoa, M., Paris, G., Taboada, A., 1998. Proyecto para la mitigación del
Audemard, F.A., 2002. Ruptura de los grandes sismos históricos venezolanos de los riesgo sísmico de Pereira Dosquebradas y Santa Rosa de Cabal: Evaluación neotectónica.
siglos XIX y XX revelados por la sismicidad instrumental contemporánea. XI Informe Final, Corporación Autónoma Regional de Risaralda, Pereira, Colombia.
Congreso Venezolano de Geofísica, edited by S.V.d.I. Geofísicos, pp. 8, Caracas. Hancock, P.L., 1985. Brittle microtectonics; principles and practice. Journal of Structural
Audemard, A.F., De Santis, F., 1991. Survey of liquefaction structures induced by recent Geology 7 (3–4), 437–457.
moderate earthquakes. Bulletin of the International Association of Engineering Hansen, W.R., Schmidt, R.A., Bedford, B., Calderwood, K.W., Ganopole, G., Hamilton, J.A.,
Geology 44, 5–16. Helmuth, D.N., Moening, H.J., Richter, D.H., 1965. Effects of the earthquake of March 27,
Audemard, F.A., Gomez, J.C., Tavera, H.J., Orihuela, G.N., 2005. Soil liquefaction during 1964, at Anchorage, Alaska. U. S. Geological Survey Professional Paper, pp. A1–A68,
the Arequipa Mw 8.4, June 23, 2001 earthquake, southern coastal Peru. Engineering U. S. Geological Survey, Reston, VA, United States.
Geology 78, 237–255. Harding, T.P., 1974. Petroleum traps associated with wrench faults. AAPG Bulletin 58,
Bohórquez, O.P., Monsalve, M.L., Velandia, F., Gil, C.F., Mora, H., 2005. Marco tectónico de 1290–1304.
la cadena volcánica más septentrional de la Cordillera Central de Colombia. Boletin Hooghiemstra, H., Cleef, A.M., 1995. Pleistocene climatic change and environmental and
de Geología, Universidad Industrial de Santander (UIS) 27, 55–80. genetic dynamics in the North Andean montane forest and páramo. In: Churchill, S.e.a.
Botero, P., García, L., Hernández, C., Sepúlveda, L., Torres, E., 2004a. Cartografía de depósitos (Ed.), Biodiversity and conservation of neotropical montane forests. Botanical Garden,
cuaternarios y evaluación morfotectónica de las fallas Silvia-Pijao y Córdoba entre los New York, pp. 35–49.
municipios de Calarcá y Pijao (Quindío): un enfoque hacia la paleosismicidad. Tesis de INGEOMINAS, 1999. Terremoto del Quindío (Enero 25 de 1999), informe técnico-
Grado, Universidad de Caldas, Manizales, Colombia (in Spanish) científico. INGEOMINAS, Bogotá, Colombia.
Botero, P., García, L., Hernández, C., Sepúlveda, L., Torres, E., 2004b. Evaluación de la James, M.E., 1986. Estudio sismotectónico en el area del Viejo Caldas. Informe nr. 2008.
Actividad Tectónica Reciente del Frente Montañoso Occidental de la Cordillera INGEOMINAS, Medellín, Colombia. 113 p.
Central entre los municipios de Calarcá y Pijao (Quindío). Boletín de Historia de las Keith, J.F., Rine, J.M., Sacks, P.E., 1988. Frontier Basins of Colombia, Valle del Cauca, Field
Geociencias en Venezuela 94. report, pp. 267. Earth Sciences and Resources Institute, University of South Carolina.
Bourgois, J., Toussaint, J.-F., Gonzalez, H., Azema, J., Calle, B., Desmet, A., Murcia, L.A., Keller, A., Pinter, N., 2002. Active Tectonics, Earthquakes, Uplift, and landscape. Prentice
Acevedo, A.P., Parra, E., Tournon, J., 1987. Geological history of the Cretaceous Hall, New Jersey, United States, p. 362.
ophiolitic complexes of northwestern South America (Colombian Andes). Tecto- Kellogg, J., Godley, V.M., Ropain, C., Bermudez, A., 1983. Gravity anomalies and tectonic
nophysics 143, 307–327. evolution of northwestern South America. Caribbean Geological Conference 10th,
Branquet, Y., Cheilletz, A., Cobbold, P.R., Baby, P., Laumonier, B., Giuliani, G., 2002. Andean pp. 18–31. Cartagena, Colombia.
deformation and rift inversion, eastern edge of Cordillera Oriental (Guateque- Lalinde, C., 2004. Evidencias Paleosísmicas en la región Pereira-Armenia, Colombia. Msc
Medina area), Colombia. Journal of South American Earth Sciences 15, 391–407. thesis, 182 pp., Universidad EAFIT, Medellín, Colombia (in Spanish).
Caballero, A., Zapata, G., 1983. Mapa geológico de Colombia, Plancha N°224 — Pereira. López, M.C., Audemard, F.A., Velásquez, A., 2005a. Evidencias geomorfológicas y
INGEOMINAS, Medellín. estratigráficas de compresión Holocena en el Valle del Cauca, Colombia. X Congreso
Cardona, J.F., Ortiz, M., 1994. Aspectos estratigráficos de las unidades del intervalo colombiano de geología, Bogotá. Colombia.
Plioceno Holoceno entre Pereira y Cartago. Propuesta de definición para la López, M.C., Moreno, M., 2005b. Tectónica y sedimentación en el piedemonte occidental
Formación Pereira. Tesis de Grado, Universidad de Caldas, Manizales, Colombia. de la Cordillera Central de Colombia. Un ejemplo en la cantera El Vínculo. X Congreso
(in Spanish) colombiano de geología, Bogotá, Colombia.
Champagnac, J.D., Sue, C., Delacou, B., Burkhard, M., 2003. Brittle orogen-parallel MacDonald, W.D., Estrada, J.J., Sierra, G.M., González, H., 1996. Late Cenozoic tectonics
extension in the internal zones of the Swiss Alps (South Valais). Eclogae geol. Helv. and paleomagnetism of North Cauca Basin intrusions, Colombian Andes; dual
96, 325–338. rotation modes. Tectonophysics 261, 277–289.
Cline, K.M., Hutchings, L., Page, W.D., Jaramillo, J., 1981a. Quaternary tectonics of Mann, P., Corrigan, J., 1990. Model for late Neogene deformation in Panama. Geology
northwest Colombia. Revista CIAF 6, 113–114. (Boulder) 18, 558–562.
Cline, K.M., Page, W.D., Gilliam, M.L., Cluff, L.S., Arias, L.A., Benalcázar, L.G., López, J.H., Marquínez, G., 2001. Modelamiento del Abanico de Ibagué con métodos geofísicos,
1981b. Quaternary activity on the Romeral and Cauca Faults, northwest Colombia. implicaciones hidrogeológicas y estructurales, Departamento de Tolima, Colombia.
Revista CIAF 6, 115–116. VIII Congreso Colombiano de Geología, Manizales, Colombia.
F. Suter et al. / Tectonophysics 460 (2008) 134–157 157

McCourt, W.J., 1984. The geology of the Central Cordillera in the departments of Valle Tectonic controls on basin development in Proto-Caribbean margins. Geological
del Cauca, Quindío and NW Tolima. INGEOMINAS, Brit. Geol. Surv., Cali, Colombia, Society of America, Special Paper, 394, pp. 7–52.
p. 37. Petit, J.P., 1987. Criteria for the sense of movement on fault surfaces in brittle rocks.
McCourt, W.J., Aspden, J.A., 1983. Modelo tectónico de placas para la evolución Journal of Structural Geology 9, 597–608 5–6.
fanerozóica de Colombia central y del sur. INGEOMINAS, Cali, Colombia, p. 36. Rastogi, B.K., 2004. Damage due to the M (sub w) 7.7 Kutch, India earthquake of 2001.
McCourt, W.J., Aspden, J.A., 1985. Modelo tectónico de placas para la evolución Tectonophysics 390, 85–103 1–4.
Fanerozoica de Colombia central y del sur. Memorias — VI Congreso Latinoamer- Restrepo, J.J., Toussaint, J.F., 1985. Cuencas de tracción sinistrales en la falla de Minas del
icano de Geologia, Bogotá, Colombia, 6, pp. 1–35. Sistema Cauca-Romeral, en las cercanías de Medellín, Colombia. Facultad de
McCourt, W.J., Mosquera, D., Nivia, G., Núñez, T., 1984. Mapa geológico de Colombia, Ciencias, Universidad Nacional de Colombia, Medellín, Colombia.
Plancha N°243 — Armenia. INGEOMINAS, Cali, Colombia. Restrepo, J.J., Toussaint, J.F., 1988. Terranes and continental accretion in the Colombian
Monsalve, H., Vargas, C.A., 2002. El sismo de Armenia, Colombia (Mw = 6.2) del 25 de Andes. Episodes 11, 189–193.
enero de 1999: Un análisis telesísmico de ondas de cuerpo, observaciones de campo y Restrepo-Pace, P.A., 1992. Petrotectonic characterization of the Central Andean Terrane,
aspectos Sismotectónicos. Revista Geofísica del Instituto Panamericano de Geografía e Colombia. Journal of South American Earth Sciences 5, 97–116.
Historia (IPGH) 57, 21–57. Rios, P.A., Aranzazu, J.M., 1989. Analysis litofacial del intervalo Oligoceno-Mioceno en el
Montes, N., Sandoval, A., 2001a. Base de datos de fallas activas de Colombia. INGEOMINAS, sector noreste de la subcuenca del Valle del Cauca, Departamento del Valle,
Bogotá, Colombia. Colombia. Tesis de Grado, 257 pp., Universidad de Caldas, Manizales, Colombia (in
Montes, N., Sandoval, A., 2001b. Monografía: Fuentes sismogénicas que contribuyen a la Spanish).
Amenaza sísmica en el Eje Cafetero. Compendio geológico, neotectónico y sismológico Rovida, A., Tibaldi, A., 2005. Propagation of strike-slip faults across Holocene volcano-
de la zona del Eje Cafetero. INGEOMINAS, Bogotá, Colombia. sedimentary deposits, Pasto, Colombia. Journal of Structural Geology 27, 1838–1855.
Montes, C., Hatcher, J., Robert, D., Restrepo-Pace, P.A., 2005a. Tectonic reconstruction of Schreurs, G., 2003. Fault development and interaction in distributed strike-slip shear
the northern Andean blocks: Oblique convergence and rotations derived from the zones; an experimental approach. Geological Society of London, London, United
kinematics of the Piedras-Girardot area, Colombia. Tectonophysics 399, 221–250. Kingdom.
Montes, N., Velandia, F., Osorio, J., Audemard, F., Diedrix, H., 2005b. Interpretación Schreurs, G., Colletta, B., 2002. Analogue modelling of continental transpression. Journal
morfotectónica de la falla Ibagué para su caracterización paleosismológica. Boletín of the Virtual Explorer (online) 7, 67–78.
de Geología, Universidad Industrial de Santander (UIS) 27, 95–114. Suter, F., 2003. Géologie de la région de Playa Azul, partie occidentale distale du fan
Moreno, M., Pardo, A., 2002. Historia geológica del Occidende colombiano. Geo-Eco- fluvio-volcanique du Quindío (Serranía de Santa Barbara, Quindío et Valle del
Trop 26, 91–113. Cauca, Colombie), Msc thesis, 133 pp., Université de Genève, Genève (in French).
Mountney, N.P., Westbrook, G.K., 1997. Quantitative analysis of Miocene to Recent Suter, F., Neuwerth, R., Guzman, C., Gorin, G., 2008. Depositional model of (Plio-)
forearc basin evolution along the Colombian convergent margin. Basin Research 9, Pleistocene sediments in a tectonically active zone of Central Colombia. Geologica
177–196. Acta 6 (3), 231–249.
Neuwerth, R., Suter, F., Guzman, C.A., Gorin, G.E., 2006. Soft-sediment deformation in a Taboada, A., Rivera, L.A., Fuenzalida, A., Cisternas, A., Philip, H., Bijwaard, H., Olaya, J.,
tectonically active area: the Plio-Pleistocene Zarzal Formation in the Cauca Valley Rivera, C., 2000. Geodynamics of the Northern Andes; subductions and intraconti-
(Western Colombia). Sedimentary Geology 186, 67–88. nental deformation (Colombia). Tectonics 19, 787–813.
Nivia, A., 1996. The Bolivar mafic-ultramafic complex, SW Colombia: the base of an Tchalenko, J.S., 1970. Similarities between shear zones of different magnitudes.
obducted oceanic plateau. Journal of South American Earth Sciences 9, 59–68. Geological Society of America Bulletin 81, 1625–1639.
Nivia, A., Galvis, N., Maya, M., 1995. Mapa geológico de Colombia, plancha 242 — Zarzal. Toto, E.A., Kellogg, J.N., 1992. Structure of the Sinu-San Jacinto fold belt, an active
INGEOMINAS, Cali, Colombia. accretionary prism in northern Colombia. Journal of South American Earth Sciences
Nivia, A., McCourt, W.J., Aspden, J.A., De Armas, M., Aucott, J.W., 2001. Geología del 5, 211–222.
departamento del Valle del Cauca. INGEOMINAS, Bogotá, Colombia. Trenkamp, R., Kellogg, J.N., Freymueller, J.T., Mora, H.P., 2002. Wide plate margin
Núñez, A., Gómez, J., Rodríguez, G., 2001. Vulcanismo básico al sureste de la ciudad de deformation, southern Central America and northwestern South America, CASA
Ibagué, departamento del Tolima — Colombia. VIII Congreso colombiano de GPS observations. Journal of South American Earth Sciences 15, 157–171.
geología, V Conferencia colombiana de geología ambiental, Manizales, Colombia. U.S.G.S., 2005. http://edc.usgs.gov/products/elevation.html.
Ortner, H., Reiter, F., Acs, P., 2002. Easy handling of tectonic data; the programs Van der Hammen, T., 1958. Estratigrafía de Terciario y Maestrichtiano continentales y
TectonicVB for Mac and TectonicsFP for Windows. Shareware and freeware in the tectogenesis de los Andes colombianos. INGEOMINAS. Boletin Geológico 67–128 VI.
geosciences; II, A special issue in honour of John Butler. Pergamon, New York- Van der Hilst, R., Mann, P., 1994. Tectonic implications of tomographic images of
Oxford-Toronto, International. subducted lithosphere beneath northwestern South America. Geology (Boulder) 22,
Ospina, L.M., 2007. Morphotectonique des dépôts quaternaires dans la région de 451–454.
Calarcá, Quindío (Colombie centrale). Msc thesis, 96 pp., Université de Genève, Van der Pluijm, B., Marshak, S., 2004. Earth Structure. Norton, New York, p. 656.
Genève, Switzerland (in French). Vargas, C., Monsalve, H., Valdés, M., Ochoa, L., Espinosa, A., Castillo, L., Montes, L.,
Page, W.D., 1986. Seismic geology and seismicity of northwestern Colombia. Integral Ordóñez, L., Nieto, M., Camacho, G., Chicangana, G., Kammer, A., 2005. Geología
Ing. Consult.. Medellin, Colombia, p. 198. estructural, respuesta morfotectónica y modelamiento numérico del sistema de
Pardo, A., Moreno, M., Gómez, A., 1994. Evidencias de actividad neotectónica en la fallas de Romeral dentro del Abanico del Quindío, sector de Armenia. X Congreso
carretera Cartago-Ansermanuevo (Valle del Cauca, Colombia). 6 pp., Universidad de Colombiano de Geología, p.116, Bogotá, Colombia.
Caldas, Manizales, Colombia. Velandia, F., Acosta, J., Terraza, R., Villegas, H., 2005. The current tectonic motion of the
Paris, G., 1997. Fallas potencialmente sismogénicas que pueden afectar las obras del Northern Andes along the Algeciras Fault System in SW Colombia. Tectonophysics
complejo vial y cruces a desnivel de la 2a con Avenida Bolivar y de la Cejita en la 399, 313–329.
Avenida República del Líbano. Alcaldía de Armenia, Armenia, Colombia. Vergara, H., 1999. Actividad neotectónica de la falla de Ibagué, Colombia. V Congreso
Paris, G., Romero, J.A., 1994. Fallas activas en Colombia. Boletín Geológico — Ingeominas Colombiano de Geología, vol. I, pp. 147–167, Bucaramanga, Colombia.
34, 3–25. Vergara, H., Espinosa, A., Jiménez, E., 2001. Análisis sismotectónico y tectónica activa de
Paris, G., Machette, M.N., Dart, R.L., Haller, K.M., 2000. Map and database of Quaternary la región epicentral del sismo del Quindío. VIII Congreso Colombiano de Geología y
faults and folds in Colombia and its offshore regions. U. S. Geological Survey. Reston, V Conferencia de Geología Ambiental, Manizales, Colombia.
VA, United States. 61 pp., 1 sheet. Wallbrecher, E., 1986. Tektonische und gefügeanalytische Arbeitsweisen. Enke-Verlag,
Parra, E., 1983. Mapa geológico de Colombia, Plancha N°223 - El Cairo (Valle del Cauca), Stuttgart, Germany, p. 244.
INGEOMINAS, Cali, Colombia. White, S.M., Trenkamp, R., Kellogg, J.N., 2003. Recent crustal deformation and the
Pennington, W.D., 1981. Subduction of the eastern Panama Basin and seismotectonics of earthquake cycle along the Ecuador-Colombia subduction zone. Earth and
northwestern South America. Journal of Geophysical Research. B 86, 10753–10770. Planetary Science Letters 216, 231–242.
Pindell, J., Kennan, L., Maresch, W.V., Stanek, K., Draper, G., Higgs, R., 2005. Plate-
kinematics and crustal dynamics of circum-Caribbean arc-continent interactions:

Das könnte Ihnen auch gefallen